Volleyball: A New Twist On Terminal Velocity Thomas W. Cairns, Ph. D. Department of Mathematical Sciences The University of Tulsa



Similar documents
Project: OUTFIELD FENCES

Ground Rules. PC1221 Fundamentals of Physics I. Kinematics. Position. Lectures 3 and 4 Motion in One Dimension. Dr Tay Seng Chuan

XI / PHYSICS FLUIDS IN MOTION 11/PA

Mathematical Modeling and Engineering Problem Solving

When the fluid velocity is zero, called the hydrostatic condition, the pressure variation is due only to the weight of the fluid.

1 of 7 9/5/2009 6:12 PM

The Viscosity of Fluids

The Viscosity of Fluids

Mechanics 1: Conservation of Energy and Momentum

Chapter 3 Falling Objects and Projectile Motion

The Physics and Math of Ping-pong and How It Affects Game Play. By: Connor Thompson & Andrew Johnson

1.3.1 Position, Distance and Displacement

The Physics of Kicking a Soccer Ball

Experiment 3 Pipe Friction

Rotation: Moment of Inertia and Torque

Acceleration Introduction: Objectives: Methods:

Natural Convection. Buoyancy force

Parachute Jumping, Falling, and Landing David C. Arney, Barbra S. Melendez, Debra Schnelle 1

F = ma. F = G m 1m 2 R 2

In order to describe motion you need to describe the following properties.

Swissmetro travels at high speeds through a tunnel at low pressure. It will therefore undergo friction that can be due to:

Dimensional analysis is a method for reducing the number and complexity of experimental variables that affect a given physical phenomena.

du u U 0 U dy y b 0 b

Ch 7 Kinetic Energy and Work. Question: 7 Problems: 3, 7, 11, 17, 23, 27, 35, 37, 41, 43

AP Physics C. Oscillations/SHM Review Packet

Chapter 6 Work and Energy

Lecture L2 - Degrees of Freedom and Constraints, Rectilinear Motion

Numerical Solution of Differential Equations

W i f(x i ) x. i=1. f(x i ) x = i=1

A Determination of g, the Acceleration Due to Gravity, from Newton's Laws of Motion

Viscous flow in pipe

Chapter 2. Derivation of the Equations of Open Channel Flow. 2.1 General Considerations

Lab 2: Vector Analysis

Hydraulic losses in pipes

FRICTION, WORK, AND THE INCLINED PLANE

Calculating average acceleration from velocity change and time

Newton s Laws. Physics 1425 lecture 6. Michael Fowler, UVa.

FLUID MECHANICS IM0235 DIFFERENTIAL EQUATIONS - CB _1

Introduction to COMSOL. The Navier-Stokes Equations

Physics 2A, Sec B00: Mechanics -- Winter 2011 Instructor: B. Grinstein Final Exam

Speed, velocity and acceleration

MULTIPLE CHOICE. Choose the one alternative that best completes the statement or answers the question.

Chapter 28 Fluid Dynamics

The Bullet-Block Mystery

NUMERICAL ANALYSIS OF THE EFFECTS OF WIND ON BUILDING STRUCTURES

Soccer ball lift coefficients via trajectory analysis

A. Hyll and V. Horák * Department of Mechanical Engineering, Faculty of Military Technology, University of Defence, Brno, Czech Republic

Lecture L6 - Intrinsic Coordinates

Behavioral Animation Simulation of Flocking Birds

Newton s Laws. Newton s Imaginary Cannon. Michael Fowler Physics 142E Lec 6 Jan 22, 2009

LAB 6: GRAVITATIONAL AND PASSIVE FORCES

Acceleration levels of dropped objects

Experimental investigation of golf driver club head drag reduction through the use of aerodynamic features on the driver crown

Chapter 10: Linear Kinematics of Human Movement

Acceleration Due to Gravity

Free Fall: Observing and Analyzing the Free Fall Motion of a Bouncing Ping-Pong Ball and Calculating the Free Fall Acceleration (Teacher s Guide)

Pipe Flow-Friction Factor Calculations with Excel

4.What is the appropriate dimensionless parameter to use in comparing flow types? YOUR ANSWER: The Reynolds Number, Re.

Physics 111: Lecture 4: Chapter 4 - Forces and Newton s Laws of Motion. Physics is about forces and how the world around us reacts to these forces.

Forces. Definition Friction Falling Objects Projectiles Newton s Laws of Motion Momentum Universal Forces Fluid Pressure Hydraulics Buoyancy

FREE FALL. Introduction. Reference Young and Freedman, University Physics, 12 th Edition: Chapter 2, section 2.5

Physics Labs with Computers, Vol. 2 P38: Conservation of Linear Momentum A

Practice Problems on Boundary Layers. Answer(s): D = 107 N D = 152 N. C. Wassgren, Purdue University Page 1 of 17 Last Updated: 2010 Nov 22

HEAVY OIL FLOW MEASUREMENT CHALLENGES

Notes on Polymer Rheology Outline

Chapter 11 Equilibrium

LAB 6 - GRAVITATIONAL AND PASSIVE FORCES

Chapter 4. Forces and Newton s Laws of Motion. continued

Lecture 5 Hemodynamics. Description of fluid flow. The equation of continuity

Calculus 1st Semester Final Review

1. Fluids Mechanics and Fluid Properties. 1.1 Objectives of this section. 1.2 Fluids

The aerodynamic center

Figure 1.1 Vector A and Vector F

Physics Notes Class 11 CHAPTER 3 MOTION IN A STRAIGHT LINE

Lecture L5 - Other Coordinate Systems

Resistance in the Mechanical System. Overview

MECHANICAL PRINCIPLES HNC/D MOMENTS OF AREA. Define and calculate 1st. moments of areas. Define and calculate 2nd moments of areas.

NACA Nomenclature NACA NACA Airfoils. Definitions: Airfoil Geometry

Motion Graphs. It is said that a picture is worth a thousand words. The same can be said for a graph.

Peter M. Arronax Consultants, Ltd S. Quid Street, Captainee, MO 61057

FLow pattern around cable and its aerodynamic characteristics in critical Reynolds number range

Worksheet 1. What You Need to Know About Motion Along the x-axis (Part 1)

Laboratory Report Scoring and Cover Sheet

Fluid Dynamics Basics

Chapter 4. Polynomial and Rational Functions. 4.1 Polynomial Functions and Their Graphs

Lecture L14 - Variable Mass Systems: The Rocket Equation

FLUID MECHANICS. TUTORIAL No.7 FLUID FORCES. When you have completed this tutorial you should be able to. Solve forces due to pressure difference.

A Comparison of Analytical and Finite Element Solutions for Laminar Flow Conditions Near Gaussian Constrictions

VELOCITY, ACCELERATION, FORCE

Review of Fundamental Mathematics

Journal of Engineering Science and Technology Review 2 (1) (2009) Lecture Note

Physical Quantities and Units

PHY231 Section 2, Form A March 22, Which one of the following statements concerning kinetic energy is true?

Chapter 4: Newton s Laws: Explaining Motion

INVESTIGATION OF FALLING BALL VISCOMETRY AND ITS ACCURACY GROUP R1 Evelyn Chou, Julia Glaser, Bella Goyal, Sherri Wykosky

Worksheet for Exploration 2.1: Compare Position vs. Time and Velocity vs. Time Graphs

Lab 1a Wind Tunnel Testing Principles & Lift and Drag Coefficients on an Airfoil

F B = ilbsin(f), L x B because we take current i to be a positive quantity. The force FB. L and. B as shown in the Figure below.

Transcription:

Volleyball: A New Twist On Terminal Velocity Thomas W. Cairns, Ph. D. Department of Mathematical Sciences The University of Tulsa Abstract The terminal velocity problem is a popular classroom example for Calculus instructors. It is often posed in a version that is simplified to the point of being unrealistic. This paper uses a volleyball as the example object and computes terminal velocities under a variety of assumptions. Actual values for the drag coefficient are used and computations are done in Mathematica. The computations illustrate how the drag crisis intervenes and complicates the mathematical analysis. The Mathematica code used is provided in an appendix. No attempt has been made to assess the efficacy of this approach in the calculus classroom. Introduction Terminal velocity problems have long been popular with instructors in calculus and differential equations. 1 They can be posed in their full aerodynamic complexity or simplified in a variety of ways that meet the needs of the moment and exemplify the mathematical ideas that are being studied. They are inherently one-dimensional and so can be posed that way with no loss of generality. The idea is that an object is dropped from a high altitude, accelerated by gravity and this acceleration is opposed by an aerodynamic force called drag. Later we'll be specific by making the object a volleyball. Gravity acts downward and drag acts in the direction opposite to the direction of motion. So all forces are vertical and a single axis is required to represent the position of the object at any later time. We'll assume here that upward is the positive direction. Drag is in some sense proportional to the square of the object's speed. One would expect that the object would be accelerated by gravity with rapidly increasing opposition from drag and would asymptotically approach a velocity that could be computed by setting the drag equal to the weight of the ball. So how does this work out in practice? We would need to determine the velocity at time from what we know about the aerodynamics of the object. Questions might arise also about position and acceleration but the main questions of interest here have to do with velocity. All computations below are done in Mathematica and the required code is given in the appendix. Zero drag case The simplest case, included in every calculus text, is the one in which the drag is assumed to be zero which means that the ball is dropping in a vacuum. In this case one begins with Newton's second law, f = ma, where f is the weight of the ball (direction down), m is its mass and a is a the acceleration due to gravity. Since gravity accelerates every object downward at about 9.8 m ë s 2, this simplifies to a = -9.8 j m ë s 2 where j is the vertical unit vector. The problem is one dimensional and so the mathematics can play out in its scalar version a =-9.8 m ë s 2. Now integrate this expression with respect to t. The units are dropped below but, of course, t is in seconds and v and v are m/s. v is ± the initial speed. If the problem is proposed in its most general form, v will be positive or negative as the direction is up or down.

2 termvelpaper5.nb Ÿ a t =-Ÿ 9.8 t v =-9.8 t + v Integrate again wit respect to t. y is the initial position. The units for y and y are meters. Ÿ v t =-Ÿ 9.8 t t + Ÿ v t y =-4.9 t 2 + v t + y A variety of problems can be posed at this point. They might involve assuming that the ball was not just dropped with an initial velocity of, but was propelled up or down. But the problem we're considering here is the terminal velocity and these generalizations provide no additional insight. So we'll assume that v = and v =-9.8 t. This is too simple a case to be really interesting because it never happens that the drag is zero and, if it did, the limiting velocity would work out to be lim tø v =-. Real world cases In the interest of realism we're dropping a volleyball which is an object we know a lot about. Common sense tells us that the effect of drag on a moving object and its terminal velocity as well depend heavily on its shape and construction, especially the nature of its surface that is presented to the airstream. Sports balls offer an interesting and useful choice. In recent decades there have been many studies on the drag and lift of sports objects, especially balls, moving through air. 2,3 In this case lift usually is a result of spin. The technical name for lift due to spin is the Magnus effect. Good examples include a topspin forehand drive in tennis and an overhand curveball from a baseball pitcher. Examples of sports objects that lift by aerodynamic effects other than Magnus effect are frisbees, javelins and ski jumpers. The balls we drop mathematically to study terminal velocity will not be spinning. If they were, a lateral force would be produced by the Magnus effect and the ball would be diverted from a vertical path. So sports balls it is and, in particular, volleyballs. Why volleyballs? We'll see that volleyballs occupy an interesting spot among the family of sports balls with respect to aerodynamic effects. So let's get to the real mathematics starting with the general case and paring it down to our special application. The equation of motion of an object moving in air is m dv = mg - 1 dt 2 rac D v 2 t + 1 2 rac L v 2 Hs ätl. 4 This comes from setting the sum of the forces to zero. Commonly used values for the entries in this equation specific to a volleyball are below. m - mass =.268 kg g - acceleration due to gravity =-9.8 m ë s 2 r-mass density of air = 1.23 kg ë m 3 A - cross sectional area =pi.214 2 M2 m 2 U.36 m 2 t - unit tangent vector in the direction of motion s - unit normal vector s ät - unit binormal vector C D - drag coefficient C L - lift coefficient Because of the assumption of no spin, C L = reduces the equation to m dv dt = mg - 1 2 rac D v 2 t dv =- 1 rac dt 2 m D v2 t + g

termvelpaper5.nb 3 where the magnitude of the drag is.5rac D v 2 and so is a function of the square of the speed. Now introduce y(t) as the position so that v'(t) = y"(t) and evaluate.5 ra/m =.825. Then the one dimensional equation becomes y "=-.825C D»y'» y'-9.8 Even if C D were constant, it would seem futile to search for an analytic solution to this differential equation. As it is, C D is known only from experimental data. Drag coefficient for a volleyball To determine C D experimentally Beatrice Hahn and David McCulloch, two students at The University of Michigan, did a wind tunnel study in 1999 to determine the drag on a non-spinning volleyball as a senior project in aeronautical engineering under the direction of Dr. Don Geister. 5 The results were an excellent set of data that were included along with other balls in a paper by Dr. Rabi Mehta of NASA Ames and Dr. Jani Pallis of Cislunar Inc and shown in Fig. 1. 2 These data are available in numerical form in the Appendix and are used for C D in this paper. Fig. 2 contains a plot of the points. The curve through the data is the Mathematica interpolation thereof. Figure 2. C D for a Volleyball 1 4 1 1 C D 1 1.1.1.1.1 1 1 1 5 1 7 Reynolds Number The graph above is a log-log plot and the units for the horizontal axis are Reynolds numbers. Drag coefficient graphs are universally displayed in this way. The Reynolds number, typically symbolized by Re, is an extremely useful device in fluid mechanics that allows engineers to abstract out a common mathematical problem from physical situations that appear to be different. An important feature of both Reand C D is that they are

4 termvelpaper5.nb dimensionless quantities. For our purposes the Reynolds number is a scalar multiple of the ball speed given by Re = rl m v where r=1.23 kg ë m 3, mass density of air L =.214 m, diameter of the ball m = 1.79 1-5 Nsëm 2, dynamic viscosity of air vmê s = ball speed This yields the relationship Re = 1475v where v is the ball speed in m/s. Most if not all C D versus Reynolds number curves look like this one, as can be seen in Fig. 1. The phenomenon at about Re = 1, where C D drops an order of magnitude is call the drag crisis. It occurs when the boundary layer of air about 1 mm thick and next to the ball switches from laminar to turbulent. This happens at different Re for different balls as shown in Fig.1. The region of rapidly falling C D is called the critical region. The four data points in the volleyball critical region are shown in Table 1 where v is in m/s. Table 1 v C D 1.2.47 15.7.15 17..1 19.7.8 The signifigance of the C D values for a volleyball is that much of the game is played in the critical region. A spike at 3 m/s (Re = 4, L is a very hard hit. Serves at 25 m/s are hard and at 2 m/s are common. Along a path of significant length a ball typically slows by about 4 m/s. The result, well known to volleyball players, is that the ball can be made to behave eratically, much like a knuckleball pitch in baseball. This occurs mostly in serves as a result of servers learning to serve a non-spinning ball at crucial speeds vis-a-vis the critical region. It's also important to note that the Hahn-McCulloch data include all speeds that actually occur in volleyball. The data point [1,,,.12] corresponds to a speed of 68 m/s and it is likely that no volleyballs are ever hit faster than 34 m/s which is Re = 5,. Getting specific with terminal velocity Some attempts have been made to use this kind of mathematical analysis to provide volleyball coaching insight. Apparently, there is some folklore that all sports balls exhibit a C D =.2. The only way one could draw this conclusion is to presume that in practice all balls live their lives in the post-critical drag regime at which C D =.2 and neither assumption is correct for a volleyball. If we were going to simplify C D to be constant, we would either have to use the value of.47 at the precritical plateau or the postcritical value of.8. We could see what would happen in those two cases. In order to check answers, we will use two methods for computing the terminal velocity: Method 1: Set drag equal to weight and solve for the speed Method 2: Solve the ODE numerically Assume C D =.47 Recall that drag =.5 rac D v 2 and the ball weight is.268 9.8 N = 2.6264 N. Solving drag = weight using Mathematica's FindRoot[ ] produces a terminal velocity of 15.8941 m/s. This could have been solved by hand. Numerical solution of the differential equation of motion in Mathematica using r =.47 requires the function NDSolve[ ]. This function produces an interpolating function that can be evaluated at time values within the time

termvelpaper5.nb 5 range requested of NDSolve[ ] and the result is -15.8939 m/s. Plotting the interpolating function solution produces the graph in Fig. 3. Figure 3. Velocity for C D =.47 2 4 6 8 1 Assume C D =.8 Solving drag = weight using Mathematica produces a terminal speed of 38.5247 m/s. Solving the differential equation gives us -38.52. Figure 4. Velocity for C D =.8 2 4 6 8 1 Use C D values from the Hahn - McCulloch wind tunnel data Solving weight = drag now requires a numerical algorithm. If we use Mathematica's FindRoot[ ], we must have a function that interpolates the drag instead of the drag coefficient. This is straightforward to produce in Mathematica and is included in the Appendix as draginterp[ ]. The solution found this way is v = 38.7366. Clearly, this approach is inadequate and we should solve the differential equation for a complete set of values. In this case we ask NDSolve[ ] to call the C D interpolation function at each iteration. The solution -38.6223 m/s and the graph is shown in Fig. 4.

6 termvelpaper5.nb Figure 5. Velocity for Wind Tunnel C D 2 4 6 8 1 12 14 Plot all three cases We can now plot the three cases on the same graph. Figure 6. Velocities for Three Cases 2 4 6 8 1 12 14 Conclusions The above computations and plots illustrate what we might have reasoned. If one assumes C D =.47, the precritical value, the terminal velocity would be almost 16 m/s. But the wind tunnel data shows that the ball encounters the drag crisis prior to its velocity reaching that value. As long as the boundary layer is laminar, the velocity looks almost identical to the C D =.47 curve. The velocity easily makes it to 2 m/s at which value it has become postcritical and is behaving as if it had the constant value of C D =.8. Afterthoughts about drag Even though C D drops rapidly at the drag crisis, it's not obvious that the drag also drops. This is because the drag is a multiplicative function of the square of the speed. We might guess that the speed would be increasing rapidly enough that the drag wouldn't decrease even in the face of the drag crisis. To answer this, we can compute the drag from the wind tunnel data and it is plotted in Fig. 7. This shows that the drag actually decreases during the drag crisis.

termvelpaper5.nb 7 Figure 7. Drag on a Volleyball 1 1 Newtons.1.1 1-5 1 1 1 4 1 5 1 6 1 7 Re Bibliography 1. Long, Lyle N. and Weiss, Howard, The Velocity Dependence of Aerodynamic Drag: A primer for Mathematicians, The American Mathematical Monthly, Vol. 16, No. 2, (Feb., 1999), pp. 127-135. 2. Mehta, Rabindra D. and Pallis, Jani M., Sports Ball Aerodynamics: Effects of Velocity, Spin and Surface Roughness. Keynote paper presented at the Materials and Science in Sports Conference, Coronado, California, April 22, 21. 3. Armenti, Jr., Angelo (editor), The Physics of Sports, Springer-Verlag New York, Inc. (1992). 4. Munson, B., Young, D., & Okiishi, T., Fundamentals of Fluid Mechanics, John Wiley and Sons, Inc., (199). 5. McCulloch, David, Beatrice Hahn, Aerodynamic Forces on a Volleyball. Unpublished report at U. of Michigan, (1999). 6. Cairns, T., Farrior, D., Farley, A., Marshall, M., Perez, G., Van Lierop, K. and Sigley, A. Modeling the Path of a Volleyball. International Journal of Volleyball Research, (2) 3:1, 2-7.7. 7. Cairns, T., Modeling Spin and Drag Forces on a Volleyball, The Engineering of Sport 5. International Sports Engineering Association, (24). Biographical Information Thomas Cairns is a Professor of Mathematical Sciences at The University of Tulsa where he has been faculty since 1959. For seventeen years in the middle of that time span he was varsity volleyball coach and is published in the aerodynamics of the flight of the ball. He can be contacted at cairns@utulsa.edu.