Sports Technology Publication details, including instructions for authors and subscription information:

Similar documents
Dissipation Function in Hyperbolic Thermoelasticity

A determination of the dynamic response of softballs

DESCRIBING THE PLASTIC DEFORMATION OF ALUMINUM SOFTBALL BATS

Gilles Bourgeois a, Richard A. Cunjak a, Daniel Caissie a & Nassir El-Jabi b a Science Brunch, Department of Fisheries and Oceans, Box

PLEASE SCROLL DOWN FOR ARTICLE

Use and Abuse of Regression

Diatom Research Publication details, including instructions for authors and subscription information:

Online publication date: 30 March 2011

Communications in Algebra Publication details, including instructions for authors and subscription information:

Full terms and conditions of use:

The Homogeneous Markov System (HMS) as an Elastic Medium. The Three-Dimensional Case

Guangzhou, P.R. China

Full terms and conditions of use:

University, Tempe, Arizona, USA b Department of Mathematics and Statistics, University of New. Mexico, Albuquerque, New Mexico, USA

Published online: 05 Oct 2006.

Testing Goodness-of-Fit for Exponential Distribution Based on Cumulative Residual Entropy

Characterizations of Student's t-distribution via regressions of order statistics George P. Yanev a ; M. Ahsanullah b a

OF SCIENCE AND TECHNOLOGY, TAEJON, KOREA

Dresden, Dresden, Germany Published online: 09 Jan 2009.

Erciyes University, Kayseri, Turkey

Precise Large Deviations for Sums of Negatively Dependent Random Variables with Common Long-Tailed Distributions

Online publication date: 01 March 2010 PLEASE SCROLL DOWN FOR ARTICLE

The American Statistician Publication details, including instructions for authors and subscription information:

George L. Fischer a, Thomas R. Moore b c & Robert W. Boyd b a Department of Physics and The Institute of Optics,

Geometrical optics and blackbody radiation Pablo BenÍTez ab ; Roland Winston a ;Juan C. Miñano b a

Version of record first published: 01 Sep 2006.

FB 4, University of Osnabrück, Osnabrück

Park, Pennsylvania, USA. Full terms and conditions of use:

G. S. Denisov a, G. V. Gusakova b & A. L. Smolyansky b a Institute of Physics, Leningrad State University, Leningrad, B-

The Fourier transform of the unit step function B. L. Burrows a ; D. J. Colwell a a

Open problems. Christian Berg a a Department of Mathematical Sciences, University of. Copenhagen, Copenhagen, Denmark Published online: 07 Nov 2014.

Derivation of SPDEs for Correlated Random Walk Transport Models in One and Two Dimensions

Nacional de La Pampa, Santa Rosa, La Pampa, Argentina b Instituto de Matemática Aplicada San Luis, Consejo Nacional de Investigaciones Científicas

Ankara, Turkey Published online: 20 Sep 2013.

CCSM: Cross correlogram spectral matching F. Van Der Meer & W. Bakker Published online: 25 Nov 2010.

PLEASE SCROLL DOWN FOR ARTICLE

Finite Element Modeling of a Baseball

Online publication date: 22 March 2010

University of Thessaloniki, Thessaloniki, Greece

Tore Henriksen a & Geir Ulfstein b a Faculty of Law, University of Tromsø, Tromsø, Norway. Available online: 18 Feb 2011

Acyclic, Cyclic and Polycyclic P n

To cite this article: Edward E. Roskam & Jules Ellis (1992) Reaction to Other Commentaries, Multivariate Behavioral Research, 27:2,

University, Wuhan, China c College of Physical Science and Technology, Central China Normal. University, Wuhan, China Published online: 25 Apr 2014.

PLEASE SCROLL DOWN FOR ARTICLE. Full terms and conditions of use:

Discussion on Change-Points: From Sequential Detection to Biology and Back by David Siegmund

PLEASE SCROLL DOWN FOR ARTICLE

PLEASE SCROLL DOWN FOR ARTICLE

PLEASE SCROLL DOWN FOR ARTICLE

Vertical bounce of two vertically aligned balls

Geometric View of Measurement Errors

35-959, Rzeszów, Poland b Institute of Computer Science, Jagiellonian University,

MSE Performance and Minimax Regret Significance Points for a HPT Estimator when each Individual Regression Coefficient is Estimated

Xiaojun Yang a a Department of Geography, Florida State University, Tallahassee, FL32306, USA Available online: 22 Feb 2007

Fractal Dimension of Turbulent Premixed Flames Alan R. Kerstein a a

Tong University, Shanghai , China Published online: 27 May 2014.

PLEASE SCROLL DOWN FOR ARTICLE

Published online: 17 May 2012.

Melbourne, Victoria, 3010, Australia. To link to this article:

arxiv:physics/ v1 [physics.pop-ph] 4 May 2006

Global Existence of Large BV Solutions in a Model of Granular Flow

PLEASE SCROLL DOWN FOR ARTICLE

Raman Research Institute, Bangalore, India

András István Fazekas a b & Éva V. Nagy c a Hungarian Power Companies Ltd., Budapest, Hungary. Available online: 29 Jun 2011

Projective spaces in flag varieties

PLEASE SCROLL DOWN FOR ARTICLE. Full terms and conditions of use:

EXPERIMENTAL AND NUMERICAL CHARACTERIZATION OF SOFTBALLS JOSEPH GERALD DURIS

A note on adaptation in garch models Gloria González-Rivera a a

Astrophysical Observatory, Smithsonian Institution, PLEASE SCROLL DOWN FOR ARTICLE

A Strongly Convergent Method for Nonsmooth Convex Minimization in Hilbert Spaces

Communications in Algebra Publication details, including instructions for authors and subscription information:

France. Published online: 23 Apr 2007.

P. C. Wason a & P. N. Johnson-Laird a a Psycholinguistics Research Unit and Department of

Anisotropic Hall effect in Al 13 TM 4

Fokker-Planck Solution for a Neuronal Spiking Model Derek J. Daniel a a

To link to this article:

Online publication date: 12 January 2010

Chapter 9 Linear Momentum and Collisions

Avenue G. Pompidou, BP 56, La Valette du Var cédex, 83162, France

PLEASE SCROLL DOWN FOR ARTICLE

PLEASE SCROLL DOWN FOR ARTICLE

Horizonte, MG, Brazil b Departamento de F sica e Matem tica, Universidade Federal Rural de. Pernambuco, Recife, PE, Brazil

Quan Wang a, Yuanxin Xi a, Zhiliang Wan a, Yanqing Lu a, Yongyuan Zhu a, Yanfeng Chen a & Naiben Ming a a National Laboratory of Solid State

PLEASE SCROLL DOWN FOR ARTICLE

6-1. Conservation law of mechanical energy

Analysis of Leakage Current Mechanisms in BiFeO 3. Thin Films P. Pipinys a ; A. Rimeika a ; V. Lapeika a a

PLEASE SCROLL DOWN FOR ARTICLE

T. Runka a, M. Kozielski a, M. Drozdowski a & L. Szczepańska b a Institute of Physics, Poznan University of

Marko A. A. Boon a, John H. J. Einmahl b & Ian W. McKeague c a Department of Mathematics and Computer Science, Eindhoven

Springs. A spring exerts a force when stretched or compressed that is proportional the the displacement from the uncompressed position: F = -k x

C.K. Li a a Department of Mathematics and

Online publication date: 29 April 2011

A Simple Approximate Procedure for Constructing Binomial and Poisson Tolerance Intervals

(A) 0 (B) mv (C) 2mv (D) 2mv sin θ (E) 2mv cos θ

PhysicsAndMathsTutor.com

Osnabrück, Germany. Nanotechnology, Münster, Germany

A. H. Abuzaid a, A. G. Hussin b & I. B. Mohamed a a Institute of Mathematical Sciences, University of Malaya, 50603,

Momentum Practice Problems

Móstoles, Madrid. Spain. Universidad de Málaga. Málaga. Spain

The use of the Normalized Difference Water Index (NDWI) in the delineation of open water features

Transcription:

This article was downloaded by: [Alan Nathan] On: 18 May 2012, At: 08:34 Publisher: Routledge Informa Ltd Registered in England and Wales Registered Number: 1072954 Registered office: Mortimer House, 37-41 Mortimer Street, London W1T 3JH, UK Sports Technology Publication details, including instructions for authors and subscription information: http://www.tandfonline.com/loi/rtec20 Reducing the effect of the ball on bat performance measurements Alan M. Nathan a, Lloyd V. Smith b & W. L. Faber b a Department of Physics, University of Illinois, Urbana, Illinois, USA b School of Mechanical and Materials Engineering, Washington State University, Pullman, Washington, USA Available online: 17 May 2012 To cite this article: Alan M. Nathan, Lloyd V. Smith & W. L. Faber (2012): Reducing the effect of the ball on bat performance measurements, Sports Technology, DOI:10.1080/19346182.2012.663533 To link to this article: http://dx.doi.org/10.1080/19346182.2012.663533 PLEASE SCROLL DOWN FOR ARTICLE Full terms and conditions of use: http://www.tandfonline.com/page/terms-and-conditions This article may be used for research, teaching, and private study purposes. Any substantial or systematic reproduction, redistribution, reselling, loan, sub-licensing, systematic supply, or distribution in any form to anyone is expressly forbidden. The publisher does not give any warranty express or implied or make any representation that the contents will be complete or accurate or up to date. The accuracy of any instructions, formulae, and drug doses should be independently verified with primary sources. The publisher shall not be liable for any loss, actions, claims, proceedings, demand, or costs or damages whatsoever or howsoever caused arising directly or indirectly in connection with or arising out of the use of this material.

Sports Technology, 2012, ifirst article, 1 6 Reducing the effect of the ball on bat performance measurements ALAN M. NATHAN 1 *, LLOYD V. SMITH 2, & W.L. FABER 2 1 Department of Physics, University of Illinois, Urbana Illinois, USA and 2 School of Mechanical and Materials Engineering, Washington State University, Pullman Washington, USA (Received 3 March 2011; accepted 28 April 2011) Abstract A simple physical model is presented that relates the ball bat coefficient of restitution e to the ball coefficient of restitution e 0 and dynamic stiffness k 0. The model is used to develop a technique to normalize e to values of e 0 and k 0 for a standard ball. The efficacy of this normalization technique is demonstrated by comparison with experimental data. It is shown to be vastly superior to a widely used technique that is based on the physically unjustified assumption that the ratio e/e 0, commonly referred to as the Bat Performance Factor or BPF, is independent of both e 0 and k 0. A residual, but much reduced, dependence of the normalized e on k 0 is observed and is shown through finite element simulations to be due to a dependence of the bat stiffness on the ball stiffness. Keywords: baseball, softball, performance, coefficient of restitution Introduction In recent years, work has been done to measure and regulate the performance of non-wood baseball and softball bats. The measurement technique involves projecting a ball from a high-speed cannon onto a stationary bat and measuring the speed of the ball both before and after the collision. From these measurements, a value can be derived for the ball bat coefficient of restitution (COR) e, which is a measure of energy dissipation in the ball bat system. If e is to be a meaningful metric of bat performance, it is necessary to control the properties of the balls used to measure it. One such ball property is e 0, the COR of the ball when colliding with a massive rigid object, which determines the fraction of compressional energy stored in the ball that is returned as kinetic energy. A second ball property is k 0, the effective spring constant or dynamic stiffness of the ball when colliding with a massive rigid object. For a given bat, the ball stiffness controls how the initial energy is partitioned between compressional energy stored in the ball and that stored in the bat. The larger the ball stiffness, the less compressional energy is stored in the ball, leading to less overall energy dissipation and larger e. This phenomenon is popularly known as the trampoline effect (Cross, 2011; Nathan, Russell, & Smith, 2004). Based on these general ideas, a highly-simplified theoretical model is constructed in Section 2 that describes the dependence of e on e 0 and k 0. This model is used to develop a technique to normalize e to a standard ball with COR e 0S and stiffness k 0S. In Section 3, an experiment is described to test this normalization technique, with the results and discussion presented in Sections 4 and 5, respectively. The technique is applied to a practical problem in Section 6. A summary and conclusions are given in Section 7. Theoretical considerations Two-spring model for the ball bat collision The starting point is a two-spring model for the ball bat collision, Figure 1, which was previously developed by Cross as a model for the trampoline effect in the interaction of tennis balls with the racket strings (Cross, 2000). In this model, the ball and bat are each represented as masses on linear lossy springs, with force constants k 0 and k 1, respectively. We hereafter refer to k 0 as the dynamic stiffness of the ball. The two springs mutually compress each other, converting the initial center-of-mass (CM) *Correspondence: Alan M. Nathan, Department of Physics, University of Illinois, Urbana, IL, 61801, USA. E-mail: a-nathan@illinois.edu ISSN 1934-6182 print/issn 1934-6190 online q 2012 Taylor & Francis http://dx.doi.org/10.1080/19346182.2012.663533

2 A.M. Nathan et al. spot of the bat (Nathan, 2000), then Eq. (2) can be rewritten to obtain the result of Cross (2000): e 2 ¼ re2 0 þ 1 1 þ r ð3þ Figure 1. Simplified physical model for the ball-bat collision. kinetic energy entirely into compressional potential energy. The fundamental equation for the energy dissipated in the collision is: 1 2 e 2 ¼ð1 2 e 2 0 Þf 0 þð1 2 e 2 1 Þf 1 ð1þ where f 0 and f 1 are the fraction of the initial CM energy stored in the ball and bat, respectively; the quantities ð1 2 e 2 0 Þ and ð1 2 e2 1Þ are the fraction of stored energy that is dissipated in the ball or bat; and (1 2 e 2 ) is the fraction of total CM energy that is dissipated in the collision. For linear springs, f 0 ¼ k 1 /(k 1 þ k 0 ) and f 1 ¼ k 0 /(k 1 þ k 0 ). Defining r ; k 1 /k 0, which is the ratio of energy stored in the ball to that stored in the bat, Eq. 1 can be rearranged to obtain: e 2 ¼ re2 0 þ e2 1 1 þ r ð2þ Assuming no losses in the bat (i.e., e 1 ¼ 1), a reasonable assumption for impacts near the sweet Eq. 3 is the basis for our normalization procedure. A plot of e vs. r is shown in Figure 2(a) for several different values of e 0. The limiting cases have simple physical interpretations. For r @ 1, essentially all of the CM energy is stored in the ball, none in the bat, and e approaches e 0, the value for the ball alone, independent of r. This regime is typical of wood bats and low-performing hollow bats. In the opposite regime r! 1, very little energy is stored in the ball, so that e approaches 1 (or more generally e 1 ) independent of e 0. In the intermediate range, e is larger than e 0, as some of the energy that might have been stored and mostly dissipated in the ball is instead stored in the bat. For modern hollow metal or composite bats, r is generally in the range 2 15, a range in which e depends on both ball properties, e 0 and k 0. Normalizing to a standard ball Suppose a ball of known COR e 0 and dynamic stiffness k 0 is used to measure the ball bat COR for a particular bat, obtaining e. Given that information, a technique is sought to predict the ball bat COR e S when the same bat is tested with a standard or normalizing ball with COR e 0S and dynamic stiffness k 0S. In the context of the two-spring model, an exact procedure can be obtained via Eq. 3. After some algebraic manipulation, our proposed normalization prescription is obtained: sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi r S e 2 0S two 2 spring model: e S ¼ þ 1 1 þ r S ð4þ Figure 2. (a) Plot of e vs. r (Eq. 3) for three values of e 0. (b) Plot of the ratio e/e 0, commonly called the BPF, vs. r for three values of e 0, demonstrating that the BPF is not independent of either ball COR or dynamic stiffness. The dashed, solid, and dotted curves correspond to e 0 ¼ 0.40, 0.35, and 0.30, respectively. For the bat and balls studied experimentally, 2.6, r, 5.1.

(a) e (b) e 0.62 0.60 0.58 0.56 0.54 0.52 0.50 0.30 0.32 0.34 0.36 0.38 0.40 0.65 0.60 0.55 0.50 8 10 12 14 16 18 where r S is the ratio of bat stiffness to the stiffness of the standard ball, and is given by: r S ; k 1 k 0S ¼ k 0 k 0S 1 2 e 2 e 2 2 e 2 0 ð5þ A different normalization procedure (Brandt, 1997) is widely used and is based on the assumption that the ratio e/e 0, commonly known as the Bat Performance Factor or BPF, is a property of the bat alone and independent of both e 0 and k 0. The BPF normalization is given by the formula: e 0 k 0 (kn/cm) Figure 3. Results for the ball-bat COR e plotted as a function of (a) ball COR e 0 or (b) ball stiffness k 0. The blue open squares are the unnormalized values. The normalized values are shown as the red open circles (Eqs. 4 5), open black diamonds (Eqs. 7 8), or open black triangles (Eq. 6). The data in (a) are those with k 0 in the range 10.0-12.2 kn/cm. The dashed lines are linear fits to the data. e S;BPF ¼ e e 0S e 0 ð6þ Ball on bat performance measurements 3 However, the BPF assumption is not in general consistent with the two-spring model. Indeed, a careful inspection of Eq. 4 or Figure 2(b) shows that e/e 0 is independent of e 0 and k 0 only in the limit r @ 1, i.e., only for wood or low-performing hollow bats. Moreover, the BPF technique does not correct for differences in the ball dynamic stiffness. Experiment The bat and ball testing facility at the Sports Science Laboratory at Washington State University (Smith, 2008; Smith & Cruz, 2008) was used to study the dependence of e on the ball properties e 0 and k 0, with the specific goal of testing the normalization procedure of Eqs. 4 5. The measurements consisted of firing a softball from an air cannon at 49.2 ^ 0.4 m/s (110 ^ 1 mph) onto a stationary bat that is mounted horizontally and is free to pivot about a point on the handle 15 cm from the knob. The speeds of the incoming and rebounding ball are measured, from which the ball bat COR is derived using standard formulas (Nathan, 2003). From past experience with such measurements (Smith, 2008), the ball bat COR is determined with a root-meansquare precision of approximately 0.005. The measurements utilized 78 different standard softballs, whose COR and dynamic stiffness were determined in supplemental experiments (ASTM, 2010; Smith, Nathan, & Duris, 2010) and ranged from 0.31 0.39 and 8.9 17.5 kn/cm (5100 10,000 lb/inch), respectively. The bat studied was a high-performing non-wood softball bat, the Louisville Slugger Catalyst of length 0.86 m (34 inch) and weight 0.75 kg (26.5 oz). As we will discuss shortly, the r values for this bat and the balls used were in the range 2.6 5.1. From Figure 2, we see that in this regime the ball bat COR is a much stronger function of k 0 than of e 0, whereas the BPF is a strong function of both k 0 and e 0. This bat should therefore be particularly useful both for testing our procedure for normalizing to dynamic stiffness and for distinguishing the two techniques for normalizing to COR. The impact location was fixed at 16.5 cm (6.5 inch) from the barrel tip. The standard ball used to obtain normalized COR values had e 0S ¼ 0.36 and k 0S ¼ 11.7 kn/cm (6700 lb/inch). Results The results of the measurements are presented in Figure 3, where the plotted values represent individual impacts. Figure 3(a) shows the ball bat COR values plotted versus e 0, along with linear fits. Given the strong dependence of e on k 0, the results in Figure 3(a) are shown only over the limited range of dynamic stiffness 10.0-12.2 kn/cm. The two-spring

4 A.M. Nathan et al. Figure 5. Comparison of the bat cross-sectional profile for simulated impacts with a ball of high stiffness (dotted curve) and low stiffness (dashed curve). The solid curve is the bat undeformed profile. Figure 4. Bat stiffness k 1 (closed points) as a function of ball stiffness k 0, as derived from the data and Eq. 5; the solid curve is from the corresponding finite element simulation. The open points are the sum k 0 þ k 1 ; the dashed curve is from the corresponding finite element simulation. Both the data and the simulation show that the bat stiffness decreases with increasing k 0 but the sum of bat and ball stiffness is approximately constant. normalization prescription, Eqs. 4 5, removes essentially all the dependence on e 0, reducing the slope of the linear fit by a factor of fifteen. On the other hand, the BPF normalization technique, Eq. 6, overcorrects for e 0, resulting in a slope larger in magnitude and opposite in sign compared with the uncorrected data. Figure 3(b) shows all 78 COR values plotted versus k 0 and demonstrates that e has an approximately linear dependence on k 0 with a slope of 0.0157 cm/kn. The slope is reduced by a factor of three by the two-spring normalization. Ideally, the normalized slope would be zero, so there is some additional dependence of e on k 0 that is not accounted for by the two-spring model. Using Eq. 5, the bat stiffness is estimated to be approximately 45 kn/m (26,000 lb/inch), so that r falls in the range 2.6 5.1. That the BPF technique works so poorly can be easily understood from Figure 2(b), given the range of r. Indeed, the experimental BPF values are far from constant, ranging from 1.4 1.8 for the 78 impacts. Discussion While the two-spring normalization technique substantially reduces the dependence of e S on k 0 and e 0,a residual dependence of e S on k 0 was nevertheless observed. A possible key to understanding the origin of this residual dependence may be found in Figure 4, which shows that the bat stiffness k 1 (Eq. 5) decreases with increasing k 0, while in the two-spring model k 1 is assumed to be independent of k 0. To study the limitations of the two-spring model, the bat ball impact was modeled numerically using finite elements. The model was constrained to simulate the bat performance test described in Section 3. The bat was modeled using a linear elastic material, where the modulus was adjusted to achieve the desired performance level (i.e., k 1 ). The ball was modeled using a linear viscoelastic material, with the viscoelastic properties adjusted to reproduce approximately the values of e 0 and k 0 used in the experiment (Smith & Duris, 2009). Simulated impacts between these balls and bats of differing stiffness were consistent with the predictions of the two-spring model (Figure 2). Namely, when r is small, e is more sensitive to k 0 than to e 0 ; when r is large, e is more sensitive to e 0 than to k 0. Details of the finite element analysis are given in the thesis of Faber (2010). From the simulated impact, the bat stiffness could be found from the ratio of the peak contact force to the corresponding deformation of the bat profile. These values are plotted as a curve in Figure 4 and show a similar decrease in bat stiffness with increasing k 0 to the experimental data. Upon closer inspection, the simulations reveal the origin of this behavior, as illustrated in Figure 5, which shows the bat deformed shape with a high-stiffness ball and low-stiffness ball. The simulations show that the high-stiffness ball is better able to maintain its spherical shape upon impact with the bat, while the bat exhibits noticeable local deformation in the form of a flat region at the center of the contact region. The net result is a smaller effective contact area and a correspondingly larger bat deformation for a given impact force, resulting in a smaller k 1. Qualitatively, the effect of nonconstant k 1 on the normalization procedure is understood as follows. As k 1 decreases with increasing k 0, the bat-to-ball stiffness ratio r decreases more rapidly with increasing k 0 than it otherwise would, so that the technique described in

Ball on bat performance measurements 5 a ball may result in lower ball stiffness (due primarily to increased ball deformation from the bat retaining its cylindrical shape) or higher ball stiffness (due primarily to increased deformation rate). These competing effects seem to offset each other, resulting in no net change in ball stiffness within the accuracy of the numerical model (Faber, 2010). All of the results are presented as scatter plots in Figure 6. This plot demonstrates clearly the spread of uncorrected values (rms ¼ 0.036), along with the improvement obtained by normalizing using e S (rms ¼ 0.014), the further improvement using e S1 (rms ¼ 0.010), and the poor results using the BPF technique (rms ¼ 0.038). Figure 6. Scatter plot of all the data showing the distribution of ball bat COR un-normalized values and various normalized values. Closely spaced points are displaced horizontally for clarity. Section 2.2 will under-correct for ball stiffness, exactly as the data show. The simulations further show that this effect is larger for high-performance than for low-performance bats. An interesting feature of Figure 4 is that both the data and the simulations show that the sum of ball and bat stiffness, k 0 þ k 1, is nearly independent of k 0 over the region 10 18 kn/cm. This result can be used to improve the effectiveness of the normalization method by providing a prescription for finding the stiffness of the bat when impacted by the standard ball: sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi r S1 e 2 0S e S1 ¼ þ 1 ð7þ 1 þ r S1 with the modified stiffness ratio: r S1 ¼ r S þ k 0 2 k 0S ð8þ k 0S where r S is given by Eq. 5. Figure 3(b) shows the values of e S1 obtained in this manner. The slope of e S1 versus k 0 is reduced by a factor of nine relative to the uncorrected values and a factor of three relative to e S. It is perhaps surprising that both the data and simulation show a ball stiffness independent of the bat stiffness; i.e. the different bat profiles described in Figure 5 might also affect the ball stiffness. A possible explanation comes from the simulations themselves. During impact, bat barrels with high stiffness will tend to better hold their round cross-section, resulting in higher deformation in the ball, both in magnitude and rate. In contrast to the bat s linearelastic response, the ball s time-dependent nature causes increased material stiffness as the strain rate increases. Thus, a barrel with high stiffness impacting A practical application While the experiment investigated a high-performance bat, a practical application of the normalization technique applies to low-performance bats, such as those mandated by the NCAA (2009). Low performance means bats with very little trampoline effect, which in the two-spring model means that r @ 1. Under such conditions, e should be insensitive to the precise value of r so that the following approximate normalizing expression can be derived: r @ 1approximation: e S < qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi e 2 þ e 2 0S 2 e2 0 ð9þ As a numerical example, consider a ball with e 0 in the range 0.45 0.49, with a stiffness ratio r S ¼ 25.8 and a normalizing COR e 0S ¼ 0.47. Over the range of e 0, the un-normalized COR varies from 0.481 to 0.518. Using the exact expression Eq. 4, e normalized to a constant value of 0.500 independent of e 0. Using the approximation expression Eq. 9, the normalized value ranges from 0.502 to 0.498, a factor of ten reduction relative to the un-normalized values. Thus, the approximation works very well. Summary and conclusion We have presented a model of the ball bat collision that explicitly demonstrates the dependence of the ball bat COR e on the COR e 0 and dynamic stiffness k 0 of the ball. We have used this model to develop a technique for normalizing e to properties of a standard ball. We have tested the model with a highperformance softball bat. We have shown that the dependence of e on e 0 is removed by the normalization. We have also shown experimentally that the ratio e/e 0, known as the BPF, depends on both e 0 and k 0, as predicted by the two-spring model. Therefore, it is not surprising that the BPF method fails to normalize performance for the bat tested. The data show that there is a strong nearly linear dependence

6 A.M. Nathan et al. of e on k 0 and have shown that the normalization technique, while not perfect, reduces that dependence by about a factor of three. We have shown through finite element simulations that the residual dependence of e S on k 0 is due to a dependence of k 1 on k 0. The simulations lead to an improved normalization technique that reduces the dependence on k 0 by factor of nine. Finally we have derived an approximate normalization expression that is valid for low-performing bats. References ASTM (2010). ASTM 2845-10 Standard Test Method for Measuring the Dynamic Stiffness (DS) and Cylindrical Coefficient of Restitution (CCOR) of Baseballs and Softballs, Retrieved from http://www.astm.org/standards/f2845.htm Brandt, R. (1997). Method and apparatus for determining the performance of sports bats and similar equipment, Retrieved from http://www.google.com/patents?hl ¼ en&lr ¼ &vid ¼ USPAT 5672809 Cross, R. (2011). Physics of Baseball and Softball. New York: Springer. Cross, R. (2000). The coefficient of restitution for collisions of happy balls, unhappy balls, and tennis balls. American Journal of Physics, 68, 1025 1031. Faber, W. L. (2010). Reducing the effect of ball variation in bat performance measurements, Master s thesis, Washington State University, USA. Retrieved from http://www.dissertations.wsu. edu/thesis/fall2010/w_faber_090310.pdf Nathan, A. M. (2003). Characterizing the performance of baseball bats. American Journal of Physics, 71, 134 143. Nathan, A. M. (2000). Dynamics of the baseball-bat collision. American Journal of Physics, 68, 979 990. Nathan, A. M., Russell, D. A., & Smith, L. V. (2004). The physics of the trampoline effect in baseball and softball bats. In M. Hubbard, R. Mehta, & J. Pallis (Eds.), The engineering of sport V, vol. 2, (pp. 36 44). Sheffield, UK: International Sports Engineering Association. NCAA (2009). National Collegiate Athletic Association standard for testing baseball bat performance, Retrieved from http://fs.ncaa.org/ Docs/rules/baseball/bats/NCAA BBCOR Protocol FINAL 5 09.pdf Smith, L. V. (2008). Progress in measuring the performance of baseball and softball bats. Sports Technology, 1, 291 299. Smith, L. V., & Cruz, C. (2008). Identifying altered softball bats and their effect on performance. Sports Technology, 1, 196 201. Smith, L. V., & Duris, J. G. (2009). Progress and challenges in numerically modeling solid sports balls with application to softballs. Journal of Sports Sciences, 27, 353 360. Smith, L. V., Nathan, A. M., & Duris, J. G. (2010). A determination of the dynamic response of softballs. Sports Engineering, 12, 163 169.