11.06.2014 Views

Impulsive Manipulation - The Robotics Institute - Carnegie Mellon ...

Impulsive Manipulation - The Robotics Institute - Carnegie Mellon ...

Impulsive Manipulation - The Robotics Institute - Carnegie Mellon ...

SHOW MORE
SHOW LESS

You also want an ePaper? Increase the reach of your titles

YUMPU automatically turns print PDFs into web optimized ePapers that Google loves.

<strong>Impulsive</strong> <strong>Manipulation</strong><br />

Wesley H. Huang<br />

August 1997<br />

CMU–RI–TR–97–29<br />

Submitted in partial fulfillment of the<br />

requirements for the degree of<br />

Doctor of Philosophy in <strong>Robotics</strong><br />

<strong>The</strong> <strong>Robotics</strong> <strong>Institute</strong><br />

<strong>Carnegie</strong> <strong>Mellon</strong> University<br />

Pittsburgh, Pennsylvania 15213<br />

Copyright c1997 Wesley H. Huang. All rights reserved.<br />

This research was funded in part by NASA through the Graduate Student Researchers<br />

Program and through grant NAGW 1175 and by the NSF under grants IRI–9318496and IRI-<br />

9114208. <strong>The</strong> views and conclusions contained in this document are those of the author and<br />

should not be interpreted as representing the official policies, either expressed or implied,<br />

of the NSF, NASA, or the U.S. Government.


Abstract<br />

Impact is a complicated phenomenon. A collision lasts for only a short period of time but<br />

results in tremendous forces and accelerations during that time. <strong>The</strong> fact that so many<br />

sports and games involve impact reflects the both the power and the difficulty in using it.<br />

This thesis studies the use of impulsive forces to manipulate objects in order that the<br />

methods, techniques, and strategies of this mode of manipulation can be transferred to<br />

robotic manipulators. <strong>The</strong> particular form of impulsive manipulation that I have studied<br />

in this thesis is tapping planar objects which then slide on a support surface, coming to rest<br />

due to friction.<br />

<strong>The</strong> mechanics of impact and friction are first analyzed in order to plan single taps<br />

that achieve some desired translation and rotation. This solution is then extended to plan<br />

multiple tap sequences and to characterize the controllability of objects under this form of<br />

manipulation. <strong>The</strong> analysis culminates with the study of vibratory manipulation — the<br />

use repeated impacts for manipulation, particularly a high frequency low amplitude series<br />

of impacts.<br />

Experiments are a major part of this work. <strong>The</strong> experimental effort started with the design<br />

of specialized tapping devices that deliver a single controlled repeatable impact. Over<br />

the course of this work, a number of different devices were designed, built, and tested; this<br />

thesis describes these devices and the issues that were important in their design.<br />

<strong>The</strong> first experiments were single tap experiments, designed to test how well the models<br />

predict object motion for a variety of different objects and support surfaces. <strong>The</strong> next<br />

experiments demonstrated positioning via tapping. <strong>The</strong>se experiments required the adaptation<br />

of the basic planning method developed in the analysis in order to explicitly consider<br />

practical issues such as errors and tapping device limitations. <strong>The</strong> resulting planning<br />

methods lie somewhere in between traditional planners and feedback control systems. <strong>The</strong><br />

culmination of the experimental effort was a demonstration that it is possible to position<br />

an object via tapping more precisely than the tapping device itself is positioned.


Acknowledgements<br />

I am indebted to many people for their help along the way.<br />

I would like to thank Fritz Morgan, Costa Nikou, Kevin Dowling and the Medical<br />

<strong>Robotics</strong> and Computer Assisted Surgery Lab for assistance and use of their OptoTrak system;<br />

Arthur Quaid and the Microdynamic Systems Laboratory for assistance and use of<br />

their laser interferometer; Ben Brown for many years of advice and assistance on mechanisms<br />

and mechanical design; Al Rizzi for numerous suggestions, particularly on the planning<br />

methods and sensitivity analysis; and Jason Powell for building and helping design<br />

the tapping devices.<br />

I have had a great group of peers in the <strong>Manipulation</strong> Lab: Kevin Lynch, Srinivas<br />

Akella, Nina Zumel, Tammy Abel, Yan-Bin Jia, Garth Zeglin, and Mark Moll. I have also<br />

been fortunate to be a part of the Vision and Autonomous Systems Center (VASC).<br />

I have had the benefit of interaction with many other students in the <strong>Robotics</strong> PhD<br />

program. In particular, I have had valuable interactions with Brad Nelson, Dan Morrow,<br />

and Rich Voyles.<br />

Thanks to Jeff Trinkle and Yangsheng Xu, two of my committee members, and to Mike<br />

Erdmann and Takeo Kanade for their contributions to this work; also to Garth Zeglin for<br />

comments on a draft of this dissertation.<br />

Finally, I would like to thank my advisors Matt Mason and Eric Krotkov for their years<br />

of support, advice, faith, and patience.


Contents<br />

1 Introduction 1<br />

1.1 Motivations ..................................... 1<br />

1.1.1 Potentialapplications ........................... 2<br />

1.2 <strong>The</strong>sisoutline .................................... 3<br />

1.3 Assumptions..................................... 4<br />

1.3.1 Impact .................................... 4<br />

1.3.2 Friction.................................... 5<br />

1.4 RelatedWork .................................... 5<br />

1.4.1 <strong>Impulsive</strong>manipulation.......................... 5<br />

1.4.2 Friction.................................... 6<br />

1.4.3 Impact .................................... 6<br />

1.4.4 Nonprehensilemanipulation ....................... 7<br />

1.4.5 Impact-baseddynamicsimulation.................... 7<br />

1.4.6 Minimalisminrobotics .......................... 7<br />

1.4.7 Partsfeeding ................................ 8<br />

1.4.8 Roboticjuggling&ballbouncing..................... 8<br />

2 <strong>The</strong> Inverse Sliding Problem 9<br />

2.1 Axisymmetriccase ................................. 9<br />

2.1.1 Forceandtorqueduetofriction ..................... 10<br />

2.1.2 Equationsofmotionanddisplacementfunctions............ 11<br />

2.1.3 Monotonicityofthedisplacementfunctions .............. 12<br />

2.1.4 Levelcurvesofdisplacementfunctions ................. 14<br />

2.1.5 Solvingforinitialvelocities ........................ 17<br />

2.1.6 Additionalaxisymmetricproperties ................... 19<br />

2.2 Nonaxisymmetriccase ............................... 20<br />

2.2.1 Generalsolution .............................. 21<br />

2.2.2 Anexample ................................. 21<br />

2.2.3 Apracticalsolution............................. 23<br />

3 <strong>The</strong> Impact Problem 25<br />

3.1 Impactmodels.................................... 25<br />

3.2 Impactproblemconstructs ............................ 27<br />

3.2.1 Fromimpulsetovelocities......................... 28<br />

i


ii<br />

CONTENTS<br />

3.2.2 Directedvelocityratiosets&impactcones ............... 28<br />

3.2.3 Fromvelocitiestoimpulse......................... 29<br />

3.3 Examples....................................... 29<br />

3.3.1 Circularobject................................ 29<br />

3.3.2 Squareobject ................................ 29<br />

4 Planning & Controllability 31<br />

4.1 Axisymmetriccase ................................. 31<br />

4.1.1 Planning................................... 32<br />

4.1.2 Controllability ................................ 33<br />

4.2 Nearlyaxisymmetriccase ............................. 33<br />

4.2.1 Planning................................... 34<br />

4.2.2 Controllability ................................ 38<br />

5 Generating Impact 39<br />

5.1 Backgroundandphilosophy............................ 39<br />

5.1.1 Methodsofgeneratingmotionandimpact ............... 40<br />

5.1.2 Otherimpactgeneratingdevicesformanipulation........... 40<br />

5.1.3 Preliminarytappingexperiments..................... 41<br />

5.1.4 Designgoals,criteria,andconstraints .................. 41<br />

5.2 <strong>The</strong>tappingdevices................................. 42<br />

5.2.1 Firstspring-loadedtappingdevice.................... 42<br />

5.2.2 Pneumatictappingdevice......................... 43<br />

5.2.3 Electromagnetictappingdevice...................... 44<br />

5.2.4 Wreckingballtapper............................ 45<br />

5.2.5 Secondspring-loadedtappingdevice .................. 45<br />

5.3 Issuesintappingdevicedesign.......................... 46<br />

6 Single Tap Experiments 51<br />

6.1 Experimentalmaterials,setup,andprocedures ................. 52<br />

6.1.1 Measurementissues ............................ 53<br />

6.1.2 Measuringmaterialparameters...................... 54<br />

6.2 Plexiglasdiskexperiment ............................. 54<br />

6.2.1 Evaluationoftheslidingmodel...................... 55<br />

6.2.2 Evaluationoftheimpactmodel...................... 58<br />

6.3 Aluminumsquareexperiment........................... 62<br />

6.3.1 Evaluationoftheslidingmodel...................... 63<br />

6.3.2 Evaluationoftheimpactmodel...................... 64<br />

6.4 Aluminumdiskexperiment............................ 68<br />

6.4.1 Evaluationoftheslidingmodel...................... 70<br />

6.4.2 Evaluationoftheimpactmodel...................... 70<br />

6.5 Experimentalobservations............................. 72<br />

6.5.1 Strikingheight ............................... 72<br />

6.5.2 Objectoverhang .............................. 73<br />

6.5.3 Objectpath ................................. 74


CONTENTS<br />

iii<br />

6.5.4 Breakingstaticfriction........................... 74<br />

6.5.5 Threedimensionaleffects ......................... 76<br />

6.5.6 Variance in e withvelocity......................... 76<br />

6.6 Conclusions ..................................... 76<br />

7 Positioning Experiments 79<br />

7.1 Preliminaries..................................... 79<br />

7.1.1 ExperimentalSetup............................. 79<br />

7.1.2 Errormodels ................................ 80<br />

7.1.3 Positioningfigures ............................. 81<br />

7.2 Planningmethods.................................. 82<br />

7.2.1 Exacttwo-tapplans ............................ 82<br />

7.2.2 Conservativetwo-tapplans........................ 82<br />

7.2.3 Multi-tapplans ............................... 84<br />

7.3 Highprecisionpositioning............................. 88<br />

7.3.1 Experiment ................................. 88<br />

7.3.2 SensitivityAnalysis ............................ 90<br />

7.3.3 Discussion.................................. 93<br />

8 Vibratory <strong>Manipulation</strong> 95<br />

8.1 Vibratorymanipulationinonedimension.................... 96<br />

8.1.1 Intermittenttapping ............................ 96<br />

8.1.2 Continuoustapping ............................ 97<br />

8.1.3 Vibratorymanipulationandballbouncing ............... 98<br />

8.1.4 Roboticjuggling .............................. 101<br />

8.2 Vibratorymanipulationintwodimensions ................... 101<br />

8.2.1 Limitingcaseofintermittenttapping .................. 102<br />

8.2.2 Limitingcaseofcontinuoustapping................... 103<br />

8.3 Comparisonwithpushing............................. 106<br />

8.4 Examplesforthelimitingcases .......................... 107<br />

8.5 Discussion...................................... 108<br />

8.5.1 Limitingcaseofintermittenttapping .................. 108<br />

8.5.2 Continuoustappinglimitingcase .................... 109<br />

8.5.3 Trackingversusfollowingtheobject................... 109<br />

9 Conclusions 111<br />

9.1 Contributions .................................... 111<br />

9.2 Futurework ..................................... 112<br />

A Impact Analysis 115<br />

A.1 Two dimensional rigid-body collisions with friction . . . ........... 115<br />

A.1.1 Equationsofmotion ............................ 115<br />

A.1.2 Impulsespaceconstructs ......................... 116<br />

A.1.3 Impactprocess............................... 117<br />

A.1.4 End of the collision . . .......................... 118


iv<br />

CONTENTS<br />

A.1.5 Impulsesolutions.............................. 118<br />

A.2 Generatingimpulsewithinafrictioncone.................... 119<br />

A.2.1 Generalcase................................. 120<br />

A.2.2 Generalizedcentralimpacts........................ 121<br />

A.3 SpecialCases..................................... 121<br />

A.3.1 Sphericalstriker............................... 123<br />

A.3.2 Sphericalstrikerandcircularobject ................... 123<br />

A.3.3 Longthinrodstriker............................ 123<br />

A.3.4 Longthinrodstrikerwithcircularobject ................ 124


List of Figures<br />

2.1 Notationfortheaxisymmetricinverseslidingproblem ............ 10<br />

2.2 Exampledisplacementfunctions ......................... 12<br />

2.3 Levelcurvesofthedisplacementfunctions ................... 15<br />

2.4 Comparing two points on an xf levelcurve................... 16<br />

2.5 <strong>The</strong>angularvelocitytrajectoriescross....................... 16<br />

2.6 <strong>The</strong>translationalvelocitytrajectoriescross. ................... 17<br />

2.7 Solvingfortheinitialvelocities .......................... 18<br />

2.8 Relationshipbetweenvelocityanddisplacementratios ............ 20<br />

2.9 <strong>The</strong> x f displacement function for the two-dimensional barbell and a graph<br />

ofitslevelcurves................................... 22<br />

2.10 <strong>The</strong> f displacement function (net rotation) for the two-dimensional barbell<br />

andagraphofitslevelcurves. .......................... 22<br />

2.11 <strong>The</strong> y f displacement function for the two-dimensional barbell and a graph<br />

itslevelcurves. ................................... 23<br />

3.1 Notationfortheimpactproblem ......................... 27<br />

3.2 Directedvelocityratiomappings ......................... 30<br />

4.1 Planningforaxisymmetricobjects ........................ 32<br />

4.2 Small time local controllability of axisymmetric objects . ........... 33<br />

4.3 <strong>The</strong>reachableconfigurationsforanearlyaxisymmetricobject ........ 35<br />

4.4 States with the same orientation that can reach the goal in one tap . . . . . . 35<br />

4.5 States with a different orientation that can reach the goal in one tap . . . . . 36<br />

4.6 An illustration of the spiral staircase ....................... 37<br />

4.7 <strong>The</strong>spanningdistance ............................... 37<br />

4.8 <strong>The</strong> spiral staircase and reachable displacement cone projected onto the plane. 37<br />

4.9 Intersection between the spiral staircase and the reachable displacement cone 38<br />

5.1 <strong>The</strong>firstspring-loadedtappingdevice. ..................... 42<br />

5.2 Double impact example for the first spring-loaded tapping device. . . . . . 43<br />

5.3 <strong>The</strong>pneumatictappingdevice .......................... 44<br />

5.4 <strong>The</strong>electromagnetictappingdevice ....................... 45<br />

5.5 <strong>The</strong>“wreckingball”tapper ............................ 46<br />

5.6 <strong>The</strong>secondspring-loadedtappingdevice .................... 47<br />

v


vi<br />

LIST OF FIGURES<br />

5.7 Position and velocity profiles for the second spring-loaded tapping device. . 47<br />

6.1 Setupforthesingletapexperiments ....................... 52<br />

6.2 Objectalignmentmethods............................. 53<br />

6.3 Intended sampling of the space of initial conditions for runs A–G of the plexiglasdiskexperiment.<br />

............................... 54<br />

6.4 Netdisplacementsfortheplexiglasdiskexperiment. ............. 56<br />

6.5 InitialvelocitiesforRunsA–Doftheplexiglasdiskexperiment ....... 57<br />

6.6 Initial velocities for Runs E, C, F, and G of the plexiglas disk experiment . . 57<br />

6.7 Plexiglasdiskvelocityprofiles—goodfit.................... 59<br />

6.8 Plexiglasdiskvelocityprofiles—acceptablefit................. 60<br />

6.9 Plexiglasdiskvelocityprofiles—unacceptablefit ............... 61<br />

6.10 Displacements and initial velocities for the aluminum square experiment. . 63<br />

6.11Forceandtorquefunctionsonasquareobject.................. 64<br />

6.12 Calculated velocity profiles for the aluminum square experiment, unscaled<br />

torque......................................... 65<br />

6.13 Comparison of velocity profiles for the aluminum square experiment . . . . 67<br />

6.14 Measured displacements and initial velocities for the aluminum disk experiment..........................................<br />

69<br />

6.15 Comparison of velocity profiles for the aluminum disk experiment . . . . . 71<br />

6.16Spikesinvelocityprofilesduetoobjectoverhang. ............... 74<br />

6.17Curvatureinobjectpaths.............................. 75<br />

6.18Representativetrialsfromthetripoddiskexperiment ............. 77<br />

7.1 Experimentalsetupforpositioningexperiments................. 80<br />

7.2 <strong>The</strong> error projection and the effect of maximum ellipse radii on the set of<br />

reachablestates.................................... 81<br />

7.3 <strong>The</strong>exacttwo-tapplanningmethod ....................... 82<br />

7.4 Exampleofanexacttwo-tapplan. ........................ 83<br />

7.5 Exampleofanexacttwo-tapplanwithreplanning................ 83<br />

7.6 <strong>The</strong>conservativetwo-tapplanningmethod................... 83<br />

7.7 Exampleofaconservativetwo-tapplanwithreplanning............ 84<br />

7.8 Example of a conservative two-tap plan with replanning. (Worst case) . . . 85<br />

7.9 Construction of a sequence of cones leading to the goal for multi-tap planning 85<br />

7.10<strong>The</strong>firstfourconesfromthegoal......................... 86<br />

7.11 <strong>The</strong> sequence of cones leading to the goal for multi-tap planning . . . . . . 86<br />

7.12 <strong>The</strong> extended reachable displacement cone and sequence of cones leading to<br />

thegoal........................................ 87<br />

7.13Exampleofamulti-tapplan............................. 88<br />

7.14Exampleofamulti-tapplan............................ 89<br />

7.15Highprecisionpositioningtrial.......................... 89<br />

7.16Notationforhighprecisionpositioninganalysis................. 90<br />

7.17 Final position in terms of x f and a ........................ 91<br />

7.18Normoffinalconfigurationjacobian....................... 93


LIST OF FIGURES<br />

vii<br />

8.1 Onedimensionalintermittenttappinglimitingcase .............. 97<br />

8.2 Onedimensioncontinuoustappinglimitingcase................ 98<br />

8.3 Illustration of regular periodic bouncing . . . .................. 99<br />

8.4 Relationship between frequency and amplitude for stable periodic bouncing 100<br />

8.5 Illustration of regular periodic bouncing for intermittent tapping . . . . . . 101<br />

8.6 ConstructionforfindingthenetCOR ...................... 102<br />

8.7 Illustrations of the state space for a sliding rotating axisymmetric object . . 104<br />

8.8 Maximum T forpushingadisk.......................... 107<br />

F<br />

8.9 Illustration of the motion constraints on a disk under pushing and the limitingcasesoftapping.................................<br />

108<br />

8.10Reachingpointsoutsidetheinitialimpactcone................. 109<br />

A.1 Impactgeometryandnotation. .......................... 117


viii<br />

LIST OF FIGURES


List of Tables<br />

6.1 Summary of materials and parameters for single tap experiments . . . . . . 53<br />

6.2 Comparisonofdisplacementsfortheplexiglasdiskexperiment ....... 58<br />

6.3 Comparison of initial velocities from sliding for the plexiglas disk experiment 61<br />

6.4 Comparison of initial velocities from impact for the plexiglas disk experiment 62<br />

6.5 Comparison of displacements for the aluminum square experiment . . . . . 66<br />

6.6 Comparison of velocities from sliding for the aluminum square experiment 66<br />

6.7 Comparison of initial velocities from impact for the aluminum square experiment<br />

......................................... 68<br />

6.8 Comparison of displacements for the aluminum disk experiment . . . . . . 70<br />

6.9 Comparison of initial velocities from sliding for the aluminum disk experiment 72<br />

6.10 Comparison of initial velocities from impact for the aluminum disk experiment 73<br />

6.11Effectofstrikerheight ............................... 73<br />

6.12Momentumcomparisonbeforeandafterimpact ................ 75<br />

6.13Energycomparisonbeforeandafterimpact................... 75<br />

6.14Variationinthecoefficientofrestitution..................... 76<br />

8.1 Motionconstraintsforadiskandaring ..................... 105<br />

A.1 Sliding modes classified by the parameters P d , P q , s ,and r ......... 119<br />

A.2 Generating impact in a friction cone in impulse space, sticking or sliding case 121<br />

A.3 Generating impact in a friction cone in impulse space, reversed sliding or<br />

slidingcase...................................... 122<br />

A.4 Generalizedcentralimpacts............................ 123<br />

ix


x<br />

LIST OF TABLES


Chapter 1<br />

Introduction<br />

<strong>Manipulation</strong> is the act of moving objects. A manipulation task might be to move a large<br />

boxfromoneplacetoanother,toassembletwoparts,toopenastubborndoor,topush<br />

a stack of papers to a corner of a desk, to dribble a basketball, to shoot a basket, to join<br />

connectors, or to flip pancakes. Even this short list of tasks shows a variety of modes of<br />

manipulation, including pushing, sliding, throwing, and bouncing. Furthermore, it hints<br />

at how people greatly increase the range of tasks they can perform by using different parts<br />

of their arms and hands and by using different manipulation strategies.<br />

<strong>The</strong> physics and the strategies behind a mode of manipulation are largely independent<br />

of the manipulator. By understanding the essence of different modes of manipulation, we<br />

can enhance the abilities of manipulators such as robotic arms and parts feeding systems.<br />

This thesis is about the use of impulsive forces to manipulate objects, a mode of manipulation<br />

I have termed impulsive manipulation. I broadly define impulsive manipulation<br />

to be any form of manipulation consisting of:<br />

a strike to an object which rapidly imparts some change in its initial velocities, followed<br />

by<br />

free motion of this object subject to forces and constraints in its environment, such as<br />

friction, drag, and collisions.<br />

This definition of impulsive manipulation includes many different forms of manipulation<br />

involving a variety of physical processes and effects. For example, kicking a soccer ball<br />

involves deformation during a kick and aerodynamics of flight, whereas playing billiards<br />

involves rigid body impact, rolling, and sliding. In order to focus my efforts, I have limited<br />

the scope of this thesis to the study of tapping planar rigid objects which then slide on a<br />

support surface.<br />

1.1 Motivations<br />

<strong>The</strong>re are many tasks in which it is required to position an object on a support surface; examples<br />

include aligning masks and silicon wafers during fabrication processes and moving<br />

1


2 CHAPTER 1. INTRODUCTION<br />

parts or assemblies from one workcell to another. Tapping is a viable method to perform<br />

these tasks, and in some instances it may be the preferred method — consider the task of<br />

aligning a vice parallel to the travel of a milling machine bed. Generally, a machinist will<br />

clamp the vice down most of the way and gently tap the vice in order to make minute<br />

changes in its orientation.<br />

Another motivation for the study of tapping is to expand the abilities of manipulators<br />

by transferring the underlying physics and strategies of tapping or other modes of manipulation.<br />

One way in which this can be done is through a minimalist approach to robotics.<br />

Instead of using a complex robot, with many actuators and sensors, we embed knowledge<br />

of the mechanics of a task in a simpler robot, either through incorporating this knowledge<br />

into a planner or through the design of this robot. For example, sufficient knowledge of<br />

the mechanics may make it unnecessary to use sensing or may allow the design of a mechanism<br />

that mechanically produces stable or convergent behavior.<br />

Tapping (and impulsive manipulation in general) is well suited for this minimalist approach<br />

for several reasons. It is a nonprehensile (nongrasping) mode of manipulation and<br />

therefore does not limit the shape and size of objects that can be manipulated. Tapping<br />

can be done using a simple actuator, and it can be very fast because a robot need neither<br />

grasp the object nor track it as it moves. (Unlike the traditional pick-and-place paradigm<br />

of robotic manipulation, moving an object does not have to be limited to the speed or the<br />

workspace of the manipulator.) High interaction forces are inherent in impulsive manipulation.<br />

This can be desirable in many situations and can be limited in others through choice<br />

of materials and impact strength.<br />

One reason that tapping is not in the repertoire of today’s robots is that impact and friction<br />

are complex nonlinear phenomena and are thus more difficult to include in a robot’s<br />

conception of its world; neither are they as predictable as the traditional pick-and-place<br />

paradigm. Furthermore, generating the controlled impact required for manipulation has<br />

not received much attention. Thus, another motivation for this thesis is to gain experience<br />

and insight in impact and friction phenomena and in the design of mechanisms to generate<br />

a controlled impact.<br />

1.1.1 Potential applications<br />

Two potential applications will serve to set the context of this thesis and to indicate some<br />

of its directions.<br />

Micropositioning<br />

Micropositioning of macroscopic objects is difficult because large forces are required to<br />

break static friction, yet small forces or forces applied over a short time duration are required<br />

in order to effect small movements. Tapping is ideal for micropositioning because<br />

an impact breaks static friction instantaneously; the object motion is then determined by<br />

the amount of impulse delivered by the tap.<br />

<strong>The</strong>re are several different ways in which robotic micropositioning through tapping<br />

could be implemented:


1.2. THESIS OUTLINE 3<br />

using a general purpose robot to position a micropositioning tapping device at points<br />

around the object as required<br />

using a number of fixed tapping devices to position an object within a micropositioning<br />

“cell”<br />

by placing a number of tapping devices on the object itself<br />

Because of the variations in impact and friction, these methods require some sort of position<br />

sensing. <strong>The</strong> experimental work in this thesis is aimed toward the first method and<br />

uses overhead cameras for position measurement.<br />

Tapping can be more accurate than the pick-and-place method for micropositioning<br />

because the act of releasing a grasp can cause object motion. In addition, pick-and-place<br />

positioning is only as as precise as the manipulator. With tapping, there is the potential<br />

of positioning the object more precisely than the manipulator that positions a tapping device;<br />

the positioning accuracy will then be limited by the impulse resolution of the tapping<br />

device. (Hence, the importance of designing specialized tapping devices to optimize performance.)<br />

Parts feeding<br />

Many of the parts feeders in use today are vibrational parts feeders. One particular system<br />

of interest is SONY’s APOS system. This parts feeder uses trays with specially designed<br />

holes. A tray is held at an angle in an APOS machine, which vibrates the tray while pouring<br />

parts over it. <strong>The</strong> holes are designed so that parts will remain in them during the vibration<br />

only if they are in the correct orientation.<br />

<strong>The</strong> interaction between parts and these parts trays is complex; friction, impact, sliding,<br />

and rolling all come into play. This interaction includes forms of impulsive manipulation<br />

beyond the scope of this thesis — particularly situations when objects become airborne<br />

in between taps. Currently, these parts trays are designed by hand, guided by experience<br />

and experiments. We hope to eventually apply a general understanding of impulsive manipulation<br />

and other modes of manipulation to analyze this system and automate the design<br />

of these parts trays from CAD models of the parts.<br />

1.2 <strong>The</strong>sis outline<br />

This thesis consists of three main parts: analysis of the mechanics, planning, and controllability<br />

of tapping; experiments; and a study of vibratory manipulation.<br />

<strong>The</strong> first problem this thesis considers is finding a solution to the inverse mechanics of<br />

tapping. <strong>The</strong> key question is:<br />

Given a start configuration (position and orientation) and a desired goal configuration,<br />

how should an object be tapped in order for it to slide from the start<br />

to the goal?


4 CHAPTER 1. INTRODUCTION<br />

Since the manipulator interacts with the object only at the instant of impact, the mechanics<br />

of tapping can be divided into two subproblems: the inverse sliding problem and the impact<br />

problem, the subject of Chapters 2 and 3 respectively.<br />

<strong>The</strong> inverse sliding problem is to determine the initial velocities that will allow an object<br />

to achieve the correct displacement (distance and rotation) during the sliding motion;<br />

the impact problem is to determine, if possible, how to generate those initial velocities by<br />

striking the object.<br />

<strong>The</strong> solution to the inverse mechanics of tapping shows how to plan for a single tap. In<br />

Chapter 4, I extrapolate from this solution to demonstrate a method for planning multiple<br />

taps; this leads naturally to a demonstration of the controllability of tapping.<br />

Experiments have been a large part of this dissertation, and in order to perform these<br />

experiments, I needed some means of generating impact. Designing a device to produce a<br />

controlled repeatable impact turned out to be a difficult problem; Chapter 5 describes the<br />

tapping devices used in this thesis and discuss some of the key issues in their design.<br />

Most of the experiments were designed to determine how well the models used for<br />

analysis reflect actual object behavior. <strong>The</strong>se experiments consisted of many single tap<br />

runs, varying the impact parameters (and thus varying the initial object velocities) between<br />

runs. <strong>The</strong> results of these experiments in regard to both the impact and sliding models are<br />

described and discussed in Chapter 6. <strong>The</strong> remainder of the experiments demonstrated<br />

positioning via tapping; the culmination of the experimental effort was a demonstration<br />

that tapping can be used to position an object more precisely than the tapping device itself<br />

is positioned. <strong>The</strong> results of these experiments and the relevant planning and control issues<br />

are described in Chapter 7.<br />

<strong>The</strong> final subject of this thesis is vibratory manipulation, which considers the behavior of<br />

a sliding object subject to a regular or servoed striker motion, particularly a high frequency<br />

low amplitude motion. In this sense, vibratory manipulation considers system level issues<br />

whereas the bulk of this thesis concentrates on issues of mechanics and planning. I discuss<br />

and develop this subject further in Chapter 8.<br />

1.3 Assumptions<br />

I have made a number of assumptions in this thesis: the support surface is flat and homogeneous,<br />

and the objects are planar, rigid, and are of known geometry, mass, support<br />

distribution, and material properties. <strong>The</strong> only forces acting upon an object during free motion<br />

are those due to Coulomb friction. I make some additional assumptions specifically<br />

relating to the impact and friction phenomena.<br />

1.3.1 Impact<br />

<strong>The</strong> classical model of impact assumes that objects are rigid in the sense that the area near<br />

the contact point which undergoes deformation during impact must be small compared to<br />

thesizeoftheobject.<br />

Using the classical model of impact, a planar collision between two rigid bodies can<br />

be characterized by:


1.4. RELATED WORK 5<br />

the geometry and relative velocities of the bodies<br />

a coefficient of restitution<br />

a coefficient of friction between the striker and the object.<br />

A more detailed discussion of impact is left to Chapter 3.<br />

1.3.2 Friction<br />

In this thesis, I assume that Coulomb friction applies between the object and the support<br />

surface and between the object and the striker. Coulomb friction provides a simple linear<br />

relationship between the normal force and the frictional force on a sliding object. However,<br />

as Feynman [24] notes:<br />

...this lawis theresultofanenormouscomplexityofeventsandisnot,fundamentally,<br />

a simple thing. If we continue to study it more and more, measuring<br />

more and more accurately, the law will continue to become more complicated,<br />

not less. ...the more deeply we study it, and the more accurately we measure,<br />

themorecomplicatedthetruthbecomes...<br />

Despite the underlying complexity of this phenomenon, the Coulomb friction model works<br />

well over a broad range of materials, velocities, and contact situations.<br />

A difficulty in calculating the friction on a body with a finite pressure distribution is<br />

that the actual pressure distribution between a body and a support surface is never known.<br />

<strong>The</strong> actual pressure distribution is determined by microscopic variations in the surface of<br />

the object and the support surface. Much of the analysis in this thesis assumes a constant<br />

uniform support distribution; the results of the single tap experiments show that this is not<br />

an unreasonable assumption.<br />

1.4 Related Work<br />

<strong>The</strong>re has been little previous work on impulsive manipulation in the context of tapping.<br />

However, this thesis does draw from research in a number of related areas, and in this<br />

section, I have tried to identify the most relevant works in each area.<br />

1.4.1 <strong>Impulsive</strong> manipulation<br />

<strong>The</strong> first instance of impulsive manipulation in the robotics literature was work done by<br />

Higuchi [29] who used an electromagnetic coil to deliver an impulsive force for linear<br />

micropositioning. More recently, Yamagata and Higuchi [65] have developed a piezoelectric<br />

device to deliver impulsive forces. This device creates an impulsive force by rapidly<br />

expanding a piezoelectric material which pushes against a reaction mass. <strong>The</strong>y have used<br />

this device for precision alignment of optical sensors (Yamagata and Higuchi [65]) and for<br />

a positioning table for a scanning tunneling microscope which operates in an ultrahigh


6 CHAPTER 1. INTRODUCTION<br />

vacuum (Yamagata et al. [67]). <strong>The</strong>y have also explored the use of thermal expansion instead<br />

of the piezoelectric effect for analogous micro electrical mechanical systems (MEMS)<br />

devices (Yamagata et al. [66]).<br />

Zhu et al. [70] have studied what they call “releasing manipulation” which is based<br />

in part on a publication of some early results of this thesis (Huang et al. [33]). Releasing<br />

manipulation is similar to the subject of this thesis in that it involves an object sliding on<br />

a support surface; it differs in that the initial velocities are not produced by impact but by<br />

smoothly accelerating an object with a manipulator and then releasing it when the required<br />

velocities have been attained. In this paper, Zhu et al. presented the mechanics, simulations,<br />

and experimental results for a rectangular object sliding on a surface. <strong>The</strong>ir experimental<br />

results, like the results of experiments I have since conducted (presented in Chapter 6),<br />

show considerable variation in the angle rotated and distance translated.<br />

This thesis builds upon the work of Voyenli and Eriksen [62] who analyzed the physics<br />

of sliding rotating disks and rings. <strong>The</strong>ir development of the state space behavior of such<br />

an object led to their observation that sliding disks and rings will come to rest only at a<br />

certain ratio of angular to translational velocity. Goyal et al. [28] developed a more general<br />

formulation of this idea and called these special velocity ratios “eigendirections,” referring<br />

to their vector representation in generalized velocity space.<br />

1.4.2 Friction<br />

<strong>The</strong> study of friction dates back to Leonardo da Vinci in the mid 1400s, although our “modern”<br />

theories of dry friction did not appear until Amontons and Coulomb published their<br />

works on the subject in 1699 and 1781 respectively.<br />

<strong>The</strong> text by Bowden and Tabor [14] is an introduction to tribology and the mechanisms<br />

of friction. Two other books by Bowden and Tabor ([12] and [13]) present an extensive<br />

study, including many experimental results.<br />

<strong>The</strong>re has been much work in studying Coulomb friction on an object resting on a<br />

support surface. Goyal et al. ([27] and [28]) developed the limit surface, a graphical representation<br />

of the relationship between the velocities of an object and the net force and torque<br />

due to friction. Howe and Cutkosky [32] examined practical approximations to the limit<br />

surface and performed experiments in the context of robotic grasping.<br />

<strong>The</strong>re has also been work in measuring the pressure distribution and center of friction.<br />

Lynch [39] studied estimation of the pressure distribution of an object though quasistatic<br />

pushes. Yoshikawa and Kurisu [68] calculated the center of friction from pushing an object<br />

with a mobile robot.<br />

Armstrong-Hélouvry [6] did a detailed study of friction in his work in controlling a<br />

robotic arm.<br />

1.4.3 Impact<br />

<strong>The</strong> study of collision between two bodies has been studied since the time of Galileo,<br />

Descartes and Newton, and it was in their time (the mid 1600s) that the first theories of<br />

impact surfaced.


1.4. RELATED WORK 7<br />

<strong>The</strong> text by Goldsmith [26] is a standard reference on impact with friction; he reviews<br />

the classical methods of analyzing impact and presents analysis of the micro-mechanics of<br />

impact. Bowden and Tabor [13] present analysis and experiments in the micro-mechanics<br />

of impact in their book. A more recent text on impact is by Brach [15]; he discusses a larger<br />

range of phenomena such as tangential and torsional restitution, energy loss, and collision<br />

of planar linkages.<br />

Routh [54] demonstrated a graphical method for analyzing planar impact with friction<br />

which was more recently presented by Wang and Mason [63]. Stronge [57], in addition to<br />

proposing a new restitution hypothesis, gave a brief history of the development of impact<br />

theories.<br />

1.4.4 Nonprehensile manipulation<br />

Nonprehensile (nongrasping) manipulation is an important type of manipulation because<br />

of its versatility. Grippers tend to be designed to manipulate a certain size or shape of<br />

parts. In contrast, nonprehensile modes of manipulation do not need a special end effector<br />

and can be used on a larger class of objects.<br />

Much research has been focused on pushing, which was originally analyzed by Mason<br />

[44]. Many others have done additional work in pushing, including Peshkin and<br />

Sanderson ([50] and [49]), Lynch ([40] and [38]), Alexander and Maddocks [5], Goldberg [25],<br />

Mani and Wilson [42], Akella and Mason [4], Kurisu and Yoshikawa [37], and Pham et<br />

al. [51].<br />

Other forms of nonprehensile manipulation include the whole arm manipulation of<br />

Salisbury [55] and Trinkle et al.([60] and [59]) and the “palmar” manipulation of Erdmann<br />

[22], Zumel [71], Yun [69], and Paljug and Yun [48].<br />

1.4.5 Impact-based dynamic simulation<br />

Related to the study of vibratory manipulation in this thesis is the idea of impact-based<br />

dynamic simulation, as described by Mirtich and Canny [47]. This method of simulation<br />

uses only impact constraints for dynamic modeling. Noncolliding contacts, such as rolling,<br />

sliding, and resting, are modeled as a series of microcollisions. Kinematic constraints, however,<br />

pose some difficulty because of the potentially large number of collisions required to<br />

model such a constraint.<br />

1.4.6 Minimalism in robotics<br />

One theme which underlies this work is that of minimalism in robotics. Canny and Goldberg<br />

([19] and [20]) use the term “RISC robotics” (reduced intricacy in sensing and control).<br />

This is the idea of enabling simpler robots to do more complex tasks by creating more intelligent<br />

planners or “compiling” intelligence into a mechanism compensate for the use of<br />

fewer actuators and sensors.<br />

<strong>The</strong> walking machines of McGeer [45] and [46] are perhaps the earliest and most<br />

dramatic example of minimalism. <strong>The</strong>se machines have no actuators or sensors; they are<br />

designed so that the mechanism (powered by gravity) produces a stable walking gait.


8 CHAPTER 1. INTRODUCTION<br />

Other examples include the tray tilting of Erdmann and Mason [23] and the parts<br />

orienting of Akella et al.([2] and [3])<br />

1.4.7 Parts feeding<br />

Vibrational parts feeders are related to impulsive manipulation and tapping, although in<br />

these systems objects tend to become airborne between impacts with a vibrating surface.<br />

<strong>The</strong> most common type of vibratory parts feeder is the bowl feeder which is described<br />

in (Boothroyd et al. [11]). Automated design and configuration of bowl feeders has been<br />

studied by Boothroyd [10], Caine [18], Christiansen et al. [21], and Berkowitz and Canny [8].<br />

<strong>The</strong> SONY APOS system (Hitakawa [30]) is another type of vibratory parts feeding system,<br />

which has been studied by Krishnasamy [36].<br />

Other work in vibratory parts feeding includes that of Böhringer et al. [9], who oriented<br />

planar parts by taking advantage of the vibrational modes of the support surface, and<br />

Swanson et al. [58], who studied parts orienting using a vibratory juggling motion.<br />

1.4.8 Robotic juggling & ball bouncing<br />

<strong>The</strong> vibratory manipulation discussed in Chapter 8 is closely related to robotic juggling<br />

and the problem of a bouncing ball on a sinusoidally vibrating table. <strong>The</strong> most relevant<br />

work in robotic juggling studies stability of single ball (or puck) juggling, which has been<br />

done by Schaal and Atkeson [56], Bühler and Koditschek [16], Rizzi and Koditschek [53]<br />

and Burridge et al. [17].<br />

A ball bouncing on a sinusoidally vibrating table can display bounded chaotic behavior.<br />

<strong>The</strong> first works to study this problem were by Holmes [31], who did analysis of<br />

approximated mechanics, and Bapat et al. [7] who did analysis and simulation of the approximated<br />

and actual mechanics. Other analytical study has been done by Wiesenfeld and<br />

Tufillaro [64], and some experiments have been performed by Tufillaro and Albano [61].<br />

This work will be discussed in greater detail in Chapter 8.


Chapter 2<br />

<strong>The</strong> Inverse Sliding Problem<br />

<strong>The</strong> inverse sliding problem is the central problem of the mechanics of tapping. Given<br />

some initial velocities, it is fairly straightforward to do forward integration to determine<br />

where an object will slide to a stop. This chapter addresses the inverse problem, i.e. given<br />

a desired displacement, determining the required initial velocities. In Chapter 3, I address<br />

the second part of the mechanics: how to generate those initial velocities via impact.<br />

<strong>The</strong> inverse sliding problem can be formally stated as follows:<br />

A rigid planar object of known geometry, support distribution, and frictional<br />

properties slides on a uniform surface, slowing down and coming to rest only<br />

due to Coulomb friction. Given a desired displacement (translation and rotation),<br />

determine the required initial velocities.<br />

This chapter addresses separately the classes of axisymmetric and nonaxisymmetric<br />

objects. Axisymmetric objects are those objects whose pressure and mass distribution are<br />

functions of radius only. Axisymmetric objects, examples of which include rings, disks,<br />

and annuli, have the properties that:<br />

the net force and torque on the object are independent of orientation, and<br />

the net force on the object is always parallel to the translational velocity.<br />

<strong>The</strong>se properties simplify the equations of motion for axisymmetric objects.<br />

It is the coupling between translation and rotation of a sliding object that makes this<br />

problem difficult. This chapter describes a method for searching for the required initial<br />

velocities for the axisymmetric case and the application of a general numerical root finding<br />

method to the nonaxisymmetric case. This chapter also develops some additional properties<br />

of the dynamics of sliding axisymmetric objects which will be used for planning and<br />

controllability in Chapter 4.<br />

2.1 Axisymmetric case<br />

Since the net force due to friction on an axisymmetric object is always parallel to the translational<br />

velocity, axisymmetric objects will always slide in a straight line. Thus the axisym-<br />

9


10 CHAPTER 2. THE INVERSE SLIDING PROBLEM<br />

y<br />

ω<br />

x<br />

θ<br />

v<br />

θ f<br />

start<br />

x<br />

x f<br />

goal<br />

Figure 2.1: Notation for the axisymmetric inverse sliding problem. <strong>The</strong> goal configuration<br />

is at a distance x f along the ^x axis and a rotation of f . <strong>The</strong> object’s position and orientation<br />

during the sliding motion are x and , its velocities v and !.<br />

metric inverse sliding problem need only consider one dimension in translation and one<br />

dimension in rotation.<br />

Figure 2.1 shows the notation used in this section. We place the global coordinate<br />

frame at the center of mass (COM) of the object in the start configuration so that the ^x axis<br />

passes through the COM of the object in the goal configuration. <strong>The</strong> desired displacement<br />

is represented as a translation of d along the ^x axis and a rotation of about the COM. <strong>The</strong><br />

object’s configuration is represented as its position x along the axis and its rotation with<br />

respect to the start configuration. <strong>The</strong> object’s translational velocity is v, its the rotational<br />

velocity !. <strong>The</strong> object is of mass M and has a moment of inertia I about the COM.<br />

This section first develops the net force and torque load on a sliding rotating object,<br />

formulates the equations of motion, and then mathematically states the inverse sliding<br />

problem. Unfortunately, the problem cannot be solved directly because there are no analytic<br />

solutions to the equations of motion. However, reasoning about the net force and<br />

torque load and about the velocity profiles leads to a method to search for the desired<br />

initial velocities to achieve a desired displacement.<br />

2.1.1 Force and torque due to friction<br />

Under Coulomb friction, the frictional force at a point on the object directly opposes the<br />

velocity of that point but is independent of the magnitude of the velocity. <strong>The</strong> net force<br />

and torque due to friction are simply integrals over the support distribution of the object.<br />

<strong>The</strong> velocity of a point ~r on the object is given by:<br />

<strong>The</strong> direction of this velocity is:<br />

~u = v^x + !^z ~r (2.1)<br />

^u =<br />

v^x + !^z ~r<br />

jv^x + !^z ~r j = ^x + ! ^z ~r<br />

v<br />

j^x + ! ^z ~r j (2.2)<br />

v<br />

Note that the direction of this velocity is dependent only upon the the ratio of angular<br />

velocity to translational velocity, ! , and the vector ~r.<br />

v<br />

<strong>The</strong> differential force load due to friction at this point is:<br />

~df = dmg^u (2.3)


2.1. AXISYMMETRIC CASE 11<br />

(<strong>The</strong> force that acts on the object at this point is then , ~ df.) Since the net force load will lie<br />

along the x axis, we can write:<br />

<strong>The</strong> net torque load is given by:<br />

T ( ! v )= Z<br />

F ( ! v )= Z ~ df = g Z ^u dm (2.4)<br />

Z<br />

~r df ~ = g<br />

~r ^u dm (2.5)<br />

<strong>The</strong> net force and torque load are nonlinear functions and generally do not have analytic<br />

forms.<br />

Mason [43] has shown that the net force is strictly monotonic decreasing and the net<br />

torque is strictly monotonic increasing (with respect to ! ). This property will be essential<br />

v<br />

in the following subsections.<br />

Intuitively, we can understand this property from the following argument. <strong>The</strong> magnitude<br />

of the frictional force at any given point is fixed; its direction is determined by the<br />

ratio ! . For a fixed velocity v, when! =0(i.e. when ! =0), all the frictional forces oppose<br />

translation, so there will be maximum force and zero torque. As ! increases (i.e. as ! v<br />

v v<br />

increases), the frictional forces will increasingly oppose the rotational motion. This results<br />

in increasing torque and decreasing force.<br />

2.1.2 Equations of motion and displacement functions<br />

<strong>The</strong> equations of motion that govern this system are:<br />

M _v = ,F ( ! v ) (2.6)<br />

I _! = ,T ( ! v ) (2.7)<br />

_x = v (2.8)<br />

_ = ! (2.9)<br />

It is here that the coupling between translation and rotation of a sliding object becomes<br />

apparent.<br />

<strong>The</strong> total distance and angle can be treated as functions of the initial translational and<br />

rotational velocities (v 0 and ! 0 respectively). I denote these displacement functions by x f<br />

and f . <strong>The</strong>se functions are given by:<br />

x f (v 0 ;! 0 )=<br />

f (v 0 ;! 0 )=<br />

Z<br />

t f<br />

0<br />

Z t f<br />

0<br />

v(t) dt (2.10)<br />

!(t) dt (2.11)<br />

where v(t) and !(t) are solutions to the equations of motion, subject to the conditions:<br />

v(0) = v 0 !(0) = ! 0 (2.12)


12 CHAPTER 2. THE INVERSE SLIDING PROBLEM<br />

50<br />

40<br />

30<br />

wo<br />

20<br />

10<br />

0.4<br />

1<br />

xf 50<br />

0.5<br />

40<br />

0<br />

30<br />

2.0<br />

wo<br />

1.6 20<br />

1.2 0.8<br />

10<br />

vo 0.4<br />

30<br />

20<br />

10<br />

0<br />

1.2 1.6 2.0<br />

0.8 vo<br />

thf<br />

Figure 2.2: Example displacement functions x f and f for a disk of radius 0.05 meters and<br />

mass 0.1 kg; the coefficient of friction is 0.25. v 0 is in meters/sec, ! 0 in radians/sec, x f in<br />

meters, and f in radians. <strong>The</strong>se surfaces were numerically generated.<br />

and<br />

v(t f )=0 !(t f )=0 (2.13)<br />

where t f is the time at which the object comes to rest.<br />

<strong>The</strong> inverse sliding problem for axisymmetric objects can now be stated as solving:<br />

x f (v 0 ;! 0 ) = d (2.14)<br />

f (v 0 ;! 0 ) = (2.15)<br />

for ! 0 and v 0 given a desired translation d and rotation .<br />

Although x f and f cannot be computed analytically, certain properties can be deduced<br />

which will lead to a method for finding the required ! 0 and v 0 .<br />

Note that the x f function is even with respect to ! 0 and f odd, i.e. x f (v 0 ;! 0 ) =<br />

x f (v 0 ; ,! 0 ) and f (v 0 ;! 0 )=, f (v 0 ; ,! 0 ). I will often consider only positive rotations for<br />

clarity, but the results are symmetric.<br />

2.1.3 Monotonicity of the displacement functions<br />

For the remainder of this section, I assume that the initial rotational velocity ! 0 is nonnegative.<br />

Although analogous properties and methods hold for negative initial rotational<br />

velocities, the notation and discussion becomes cumbersome when handling both cases<br />

simultaneously.<br />

<strong>The</strong> displacement functions x f (v 0 ;! 0 ) and f (v 0 ;! 0 ) are monotonic in v 0 and ! 0 in<br />

the following sense: if we start at some point (v 0 ;! 0 ) these functions both increase as v 0<br />

increases (with ! 0 held constant), and vice versa. Figure 2.2 shows an example of the<br />

displacement functions.<br />

Intuitively, if the object has a greater initial translational velocity it will slide further.<br />

This means that it will take more time for the object to come to rest, so it will also rotate


2.1. AXISYMMETRIC CASE 13<br />

more. Similarly, if the object has a greater initial angular velocity, it will rotate more and<br />

also slide further.<br />

More formally, we have the following implications concerning two different trajectories.<br />

v 01 >v 02 ! 01 = ! 02 =) x f1 >x f2 (2.16)<br />

v 01 >v 02 ! 01 = ! 02 6=0 =) f1 > f2 (2.17)<br />

v 01 = v 02 6=0 ! 01 >! 02 =) x f1 >x f2 (2.18)<br />

v 01 = v 02 ! 01 >! 02 =) f1 > f2 (2.19)<br />

where x fi and fi is the displacement resulting from the initial velocities ! 0i and v 0i .<strong>The</strong>se<br />

statements can be proven by reasoning about the velocity profiles of these two trajectories.<br />

<strong>The</strong>orem 1 If, at some time t, two trajectories satisfy:<br />

then at time t + they will satisfy:<br />

v 1 >v 2 and ! 1 = ! 2 6=0<br />

v 1 >v 2 and ! 1 >! 2<br />

Proof Under the conditions at time t,wehave !1<br />

v 1<br />

< !2<br />

v 2<br />

. From the monotonicity properties<br />

of subsection 2.1.1, this implies that T ( !1<br />

) ! 2 while the condition v 1 >v 2 still holds.<br />

<strong>The</strong>orem 2 If, at some time t, two trajectories satisfy:<br />

v 1 = v 2 6=0 and ! 1 >! 2<br />

then at time t + they will satisfy:<br />

v 1 >v 2 and ! 1 >! 2<br />

Proof Under the conditions at time t, wehavej !1<br />

v 1<br />

j > j !2<br />

v 2<br />

j. From the monotonicity properties<br />

of subsection 2.1.1, this implies that F ( !1<br />

v 1<br />

) j _v 2 j, and by the definition of the derivative and of limits, this implies<br />

that at time t + we will have v 1 ! 2 still holds.<br />

<strong>The</strong>orem 3 If at some time t 0 we have:<br />

v 1 >v 2 ! 1 >! 2<br />

then for all t>t 0 the following condition will hold:<br />

v 1 v 2 ! 1 ! 2


14 CHAPTER 2. THE INVERSE SLIDING PROBLEM<br />

Proof <strong>The</strong> only way that this condition could be violated is if v 1 (t) crossed v 2 (t) or ! 1 (t)<br />

crossed ! 2 (t). Both of these crossings cannot happen simultaneously because the derivatives<br />

_v and _! are uniquely defined by the equations of motion. However it could be possible<br />

for v 1 (t) to cross v 2 (t) while ! 1 (t) >! 2 (t), orviceversa.<br />

But in order for v 1 (t) to cross v 2 (t), atsometimet we must have v 1 = v 2 . <strong>The</strong>n by<br />

<strong>The</strong>orem 2, at time t + we will have v 1 >v 2 .Inorderfor !1<br />

v 1<br />

to cross ! 2 (t), atsometimet<br />

we must have ! 1 = ! 2 . <strong>The</strong>n by <strong>The</strong>orem 1, at time t + we will have ! 1 >! 2 . Thus the<br />

condition will not be violated.<br />

We are now in a position to prove the implications from the beginning of this subsection.<br />

<strong>The</strong> left hand sides Equations 2.16 and 2.17 are the condition for <strong>The</strong>orem 1, so for<br />

some period of time after t =0, we will have v 1 >v 2 and ! 1 >! 2 . By <strong>The</strong>orem 3, we will<br />

have the condition v 1 v 2 and ! 1 ! 2 for the remaining time. Together, these imply:<br />

which simplifies to:<br />

Z t f 1<br />

0<br />

Z t f 1<br />

0<br />

v 1 (t) dt ><br />

! 1 (t) dt ><br />

Z t f 2<br />

0<br />

Z t f 2<br />

0<br />

v 2 (t) dt (2.20)<br />

! 2 (t) dt (2.21)<br />

x f (v 01 ;! 01 ) >x f (v 02 ;! 02 ) (2.22)<br />

f (v 01 ;! 01 ) > f (v 02 ;! 02 ) (2.23)<br />

Similar reasoning can be employed to prove equations 2.18 and 2.19.<br />

2.1.4 Level curves of displacement functions<br />

We would like to solve for the required initial condition by intersecting the level curve of x f<br />

corresponding to a translation d and the level curve of f corresponding to a rotation .<br />

However, since there is no analytic form for the displacement functions, there is no analytic<br />

form for their level curves. <strong>The</strong> monotonicity properties of the displacement functions can<br />

be used to determine some properties of these level curves which will lead to a method for<br />

finding the solution; this solution is unique and always exists.<br />

Trivial cases of the displacement functions<br />

<strong>The</strong> form of the displacement functions along the v 0 and ! 0 axes can easily be determined.<br />

In these cases, the coupling between translation and rotation disappears, the problem is<br />

reduced to elementary physics, and we get:<br />

x f (0;! 0 ) = 0 (2.24)<br />

f (v 0 ; 0) = 0 (2.25)<br />

x f (v 0 ; 0) =<br />

f (0;! 0 ) =<br />

1<br />

2 Mv2 0<br />

Mg<br />

1<br />

2 I!2 0<br />

g R j~r j dm<br />

(2.26)<br />

(2.27)


2.1. AXISYMMETRIC CASE 15<br />

50<br />

xf<br />

50<br />

thf<br />

40<br />

40<br />

wo<br />

30<br />

wo<br />

30<br />

20<br />

20<br />

10<br />

10<br />

0.4 0.8 1.2 1.6 2.0<br />

vo<br />

0.4 0.8 1.2 1.6 2.0<br />

vo<br />

Figure 2.3: Level curves of the displacement functions x f and f from Figure 2.2.<br />

Monotonicity of the level curves<br />

To find a level curve of x f at a height d, we start on the v 0 axis. <strong>The</strong>re, we can solve<br />

d = x f (v 0d ; 0) for v 0d using Equation 2.26. Imagine following the level curve of x f using a<br />

zigzag motion, repeatedly increasing ! 0 on the first step and then adjusting v 0 on the next.<br />

If we take a step upwards, increasing ! 0 , then we leave the level curve; by Equation<br />

2.16, we know that x f increases as we move upwards, parallel to the positive ! 0 axis.<br />

In order to get back on the level curve, we must adjust the v 0 coordinate. By Equation 2.17<br />

we know that x f increases parallel to the positive v 0 axis. Since we need to go downhill to<br />

the level curve, we decrease the v 0 coordinate. This implies that the level curve is monotonic<br />

decreasing — as its ! 0 coordinate increases, its v 0 coordinate decreases.<br />

A level curve of x f then starts at some point on the v 0 axis and moves upwards and<br />

to the left, asymptotically approaching the ! 0 axis. This level curve must lie between the<br />

lines v 0 = v 0d and v 0 =0.<br />

Similar reasoning can be employed to show that the level curve of f at a height <br />

begins at a point ! 0 on the ! 0 axis (where ! 0 is the solution to = f (0;! 0 ) using<br />

Equation 2.27). This curve is monotonic decreasing as it asymptotically approaches the v 0<br />

axis, and the curve must lie between the lines ! 0 = ! 0 and ! 0 =0.<br />

Figure 2.3 shows examples of the level curves of x f and f .<br />

Monotonicity along the level curves<br />

A level curve of x f and a level curve of f must intersect somewhere within the region<br />

bounded by the v 0 and ! 0 axes and the lines v 0 = v 0d and ! 0 = ! 0 , but their monotonicity<br />

properties do not guarantee a unique intersection point. However, the uniqueness of the<br />

intersection can be proven by showing that f is monotonic along a x f level curve and vice<br />

versa.<br />

Consider two different points on a level curve of x f . (See Figure 2.4.) Since the level


16 CHAPTER 2. THE INVERSE SLIDING PROBLEM<br />

x level curve<br />

f<br />

ω 0<br />

2<br />

1<br />

v 0<br />

Figure 2.4: By comparing velocity trajectories for two sets of initial conditions corresponding<br />

to two points on a level curve of x f , we can show that f increases along an x f level<br />

curve.<br />

v<br />

1<br />

2<br />

ω<br />

2<br />

1<br />

t<br />

t<br />

Figure 2.5: <strong>The</strong> angular velocity trajectories cross.<br />

curve is monotonic decreasing, we can assume:<br />

v 01 > v 02 (2.28)<br />

! 01 < ! 02 (2.29)<br />

and therefore !0 1<br />

v 01<br />

< !0 2<br />

v 02<br />

.Aslongas !1<br />

v 1<br />

< !2<br />

v 2<br />

, then we will have F 1 >F 2 and T 1


2.1. AXISYMMETRIC CASE 17<br />

v<br />

1<br />

2<br />

ω<br />

2<br />

1<br />

t<br />

t<br />

Figure 2.6: <strong>The</strong> translational velocity trajectories cross.<br />

If ! 1 and ! 2 cross (Figure 2.5), then we will have v 1 v 2 for the entire trajectory,<br />

implying that x f1 >x f2 which violates our assumption that the initial conditions were on<br />

a level curve of x f .<br />

Consequently, v 1 and v 2 must cross (Figure 2.6), which results in the condition ! 1 ! 2<br />

for the entire trajectory, so f1 < f2 . <strong>The</strong>refore, f increases monotonically along level<br />

curves of x f . (Since by assumption the two points are from a level curve of x f , the net area<br />

between the two translational velocity profiles will be zero.)<br />

Similar reasoning can be applied to show that x f increases monotonically along level<br />

curves of f . <strong>The</strong>se properties imply that any given level curve of x f and any given level<br />

curve of f will intersect exactly once.<br />

2.1.5 Solving for initial velocities<br />

<strong>The</strong> level curve of f for any given and the level curve of x f for any given d are guaranteed<br />

to have a unique intersection point whose coordinates will correspond to the required<br />

initial velocities to achieve that displacement. <strong>The</strong> intersection point is also bounded, but<br />

the question still remains of how to determine the coordinates of this point.<br />

<strong>The</strong> only way to get information about the location of the level curves is to do a forward<br />

integration on some initial conditions and compare the resulting displacement to the<br />

desired displacement. If a displacement component (either translation or rotation) is less<br />

than that of the desired displacement, then that point is “below” the corresponding level<br />

curve; if it is greater than the corresponding displacement component, then it is “above”<br />

that level curve.<br />

A variation on bisection can be used to zoom in on the coordinates of the intersection<br />

point. Starting with the rectangular region that bounds the intersection point, we integrate<br />

the initial conditions corresponding to the center of the rectangle. This divides the rectangle<br />

into four smaller rectangles. Based upon whether the center is above or below the x f<br />

and f level curves, we can eliminate at least one and possibly three of the newly formed<br />

subrectangles. Continuing in this manner, we can eliminate regions of the bounding rectangle<br />

by repeatedly subdividing rectangles and integrating points, eventually zooming in<br />

on the intersection point to any desired accuracy. Figure 2.7 illustrates this process.


18 CHAPTER 2. THE INVERSE SLIDING PROBLEM<br />

ωo<br />

ωo α<br />

vo<br />

d<br />

vo<br />

Figure 2.7: Solving for the initial velocities. By integrating points and determining whether<br />

they lie above or below the level curves, we can eliminate portions of the bounding rectangle<br />

and zoom in on the intersection point. <strong>The</strong> initial velocities corresponding to the<br />

intersection point are those that will produce the a translation d and rotation .<br />

Implementation notes<br />

I have implemented this procedure in C on a UNIX workstation. <strong>The</strong> underlying forward<br />

integration routines use Gaussian quadrature integration to numerically integrate the net<br />

force and torque due to friction. <strong>The</strong> equations of motion are integrated using a fixed step<br />

fourth order Runge-Kutta integration. My integration routines have drawn from Numerical<br />

Recipes in C [52], but I have written my own because of conditions unique to this problem,<br />

namely that once the object comes to rest, friction no longer acts upon it.<br />

For function integration, I experimented with less sophisticated routines (than Gaussian<br />

quadrature) including trapezoidal, Simpson, and Romberg integration. I have found<br />

that Gaussian quadrature integration gives virtually identical results as the other methods<br />

with far fewer function evaluations.<br />

For integration of the equations of motion, I have experimented with more sophisticated<br />

routines (than fixed-step fourth order Runge-Kutta), such as adaptive step-size fifth<br />

order Runge-Kutta and the Bulirsch-Stoer method. <strong>The</strong> fixed-step Runge-Kutta will overshoot<br />

the velocities, i.e. will improperly let the velocities change sign, so I modified the<br />

integration routine to back up and take smaller steps. <strong>The</strong> adaptive stepsize Runge-Kutta<br />

and the Bulirsch-Stoer methods never overshot; however, the step size became so small towards<br />

the end that these methods used far more derivative evaluations than the fixed-step<br />

Runge-Kutta. Again, the resulting answers were virtually identical.<br />

A substantial amount of code is required to keep track of all the squares, points, results<br />

of integration, and whether a square or point has been eliminated, but most of the running<br />

time is occupied by numerical integration. My implementation takes approximately 3 seconds<br />

to run on a Sun SPARC 5 using an accuracy tolerance for the resulting displacement<br />

that is beyond the accuracy of the high speed measurement device I have used for the single<br />

tap experiments. <strong>The</strong> linear position tolerance is set to 0.05 mm, the angular position<br />

tolerance to 0.057 degrees. It typically takes from 7 to 10 iterations to achieve this degree<br />

of accuracy. For an experiment with 70 trials, the average number of iterations was 8.6 and<br />

required integrating an average of 26 points.


2.1. AXISYMMETRIC CASE 19<br />

Convergence of this method would be linear like bisection if all but one rectangle<br />

was eliminated at each iteration. However this is not quite the case. For the previously<br />

mentioned experiment, there was an average of 1.2 rectangles at the end of the last iteration<br />

(most often 1 but sometimes 2).<br />

2.1.6 Additional axisymmetric properties<br />

In this subsection, I develop some additional properties of the dynamics of a sliding axisymmetric<br />

object; these properties will be used for planning and controllability in Chapter<br />

4.<br />

Trajectory scaling<br />

We can show through straightforward use of the chain rule, that if v(t) and !(t) are solutions<br />

to the equations of motion, then so are kv( t k ) and k!( t ). If the first pair of trajectories<br />

k<br />

produce a displacement of x f and f , then the second produce a displacement of k 2 x f and<br />

k 2 f .<br />

One can also think about this property in terms of the vector field:<br />

(v; !) 7! ( ,F ( ! v )<br />

M ; ,T ( ! v )<br />

) (2.34)<br />

I<br />

Since this field is dependent only upon the value of ! (i.e. the “angle” of the velocities),<br />

v<br />

the evolution of this system is not dependent upon the amount of energy of the initial<br />

conditions but simply the ratio ! . Thus, paths in state space can be scaled.<br />

v<br />

One implication of trajectory scaling is that a line through the origin in the space of<br />

initial velocities (constant !0<br />

v 0<br />

) can be mapped to a line through the origin in displacement<br />

space (constant xf<br />

f<br />

).<br />

Relationship between initial velocity and displacement ratios<br />

We can draw upon the monotonicity along level curves (i.e. that x f increases along a f<br />

and vice versa) and from the trajectory scaling property to show that xf<br />

f<br />

is a monotonic<br />

decreasing function of !0<br />

v 0<br />

.<br />

Consider a level curve of f as shown in Figure 2.8. Since x f monotonically increases<br />

along this level curve, the ratio xf<br />

f<br />

the f level curve can be parameterized as a function of !0<br />

v 0<br />

will intersect the level curve exactly once. <strong>The</strong> value of !0<br />

v 0<br />

also monotonically increases. Because of monotonicity,<br />

— every line from the origin<br />

monotonically decreases along<br />

(because of<br />

this level curve. Since the ratio xf<br />

f<br />

remains constant along a line of constant !0<br />

v 0<br />

trajectory scaling), we can conclude that xf<br />

f<br />

monotonically decreases as a function of !0<br />

v 0<br />

.<br />

Mapping initial velocity cones to displacement cones<br />

<strong>The</strong> trajectory scaling property showed that a line through the origin in initial velocity<br />

space maps to a line through the origin in displacement space. Now, with the monotonicity


20 CHAPTER 2. THE INVERSE SLIDING PROBLEM<br />

ω 0<br />

v 0<br />

x increases<br />

f<br />

ω 0<br />

v 0<br />

decreases<br />

θ level curve<br />

f<br />

Figure 2.8: Relationship between velocity and displacement ratios. We consider a level<br />

curve of f to show that xf<br />

f<br />

is a monotonic decreasing function of !0<br />

v 0<br />

.Sincex f increases<br />

along a f level curve, the value of xf<br />

f<br />

increases along this level curve; the value of !0<br />

v 0<br />

decreases along a level curve.<br />

relationship between the initial velocity ratio and the displacement ratio, we can show that<br />

cones map from one space to the other.<br />

<strong>The</strong> cones we consider are continuous ranges of rays starting at the origin. A cone in<br />

initial velocity space consists of a set of rays at the origin with a continuous range of slopes.<br />

From the monotonicity property between initial velocity ratios and displacement ratios,<br />

we know that this continuous range of slopes (values of !0<br />

v 0<br />

) maps to a continuous range of<br />

slopes (values of xf<br />

f<br />

) in displacement space. Furthermore, the cones can be characterized<br />

by their maximum and minimum slope, and the mapping between cones has a unique<br />

inverse.<br />

2.2 Nonaxisymmetric case<br />

For nonaxisymmetric objects undergoing translation and rotation, the net force due to friction<br />

will generally not be in the same direction as the net translational velocity. <strong>The</strong> underlying<br />

reason for this is that Coulomb friction is not dependent upon velocity magnitude.<br />

Consider the velocity vectors at every point on the object. <strong>The</strong> components of these vectors<br />

normal to the net translational velocity sum to zero; they do not after the vectors have been<br />

normalized.<br />

Since these objects do not move in a straight line, the equations of motion now have<br />

six state variables:<br />

M ~v _ = , F ~ ( ! ; ^v; )<br />

j~vj<br />

(2.35)<br />

I _! = ,T ( ! ; ^v; )<br />

j~vj<br />

(2.36)<br />

_~x = ~v (2.37)<br />

_ = ! (2.38)


2.2. NONAXISYMMETRIC CASE 21<br />

Also note that the force and torque load are now functions of the direction of the translational<br />

velocity ^v and the orientation of the object in addition to the velocity ratio !<br />

j~vj .<br />

In this section, I will discuss a general solution to the nonaxisymmetric case, show an<br />

example of how the displacement functions (x f , y f ,and f ) are different for nonaxisymmetric<br />

objects, and suggest a practical solution for this case of the inverse sliding problem.<br />

2.2.1 General solution<br />

Although the force and torque functions, for a given value of display the same monotonicity<br />

properties (with respect to<br />

j~vj ! ) as in the axisymmetric case, these functions change<br />

shape with variation in . Thus the rotation of the object precludes the use of these monotonicity<br />

properties in the same way that was used to find a solution for the axisymmetric<br />

case.<br />

A general solution to the nonaxisymmetric case must then rely upon general purpose<br />

numerical methods. This problem is related to two point boundary value problems —<br />

problems in which a solution to a set of differential equations must be found that satisfies<br />

constraints at two boundaries. In typical two point boundary value problems, the two<br />

boundaries are defined in terms of time, and the constraints at the boundaries are upon the<br />

state variables. For the nonaxisymmetric inverse sliding problem, the second boundary<br />

is determined by the state variables (i.e. when the object comes to rest) and the value at<br />

the second boundary is an integral over the trajectory (i.e. the object displacement). This<br />

problem lends itself naturally to the shooting method.<br />

Applying the shooting method to the nonaxisymmetric case is equivalent to performing<br />

multidimensional root finding on the displacement functions. Multidimensional root<br />

finding is a difficult problem and is best solved when one has some idea of where a solution<br />

lies. Various methods are generally based on the Newton-Raphson method — the jacobian<br />

is used to form a set of linear equations to solve for a step which brings all functions closer<br />

to zero simultaneously. In the shooting method, the Jacobian is estimated numerically.<br />

Neither existence nor uniqueness of a solution is guaranteed in this case.<br />

2.2.2 An example<br />

Although the prospects for a simple solution to the nonaxisymmetric case look bleak, examination<br />

of a few objects suggests that the problem is not quite as bad as it seems.<br />

This subsection examines the level curves of the displacement functions for a twodimensional<br />

barbell — two point masses on the end of a massless rigid rod. In this example,<br />

the rod is 0.1 meters long and each point mass is 0.1 kg; the coefficient of friction is 0.2.<br />

Figures 2.9, 2.11, and 2.10 show the displacement function and level curves for the x f , y f ,<br />

and f respectively. <strong>The</strong>se functions were computed for a translational velocity along the<br />

positive x axis and with the barbell initially aligned with the x axis.<br />

Note that the x f and f displacement functions are smooth, though there are of ripples<br />

in these surfaces. <strong>The</strong> level curves of these functions all appear to be monotonic. <strong>The</strong> y f<br />

displacement function is quite different — it has a number of distinct peaks.


22 CHAPTER 2. THE INVERSE SLIDING PROBLEM<br />

1200<br />

xf level curves<br />

xf (m)<br />

2<br />

1.5<br />

1<br />

0.5<br />

1200 0<br />

800<br />

400<br />

wo (deg/sec)<br />

0<br />

0<br />

1<br />

2<br />

vo (m/sec)<br />

3<br />

initial rot. velocity (deg/sec)<br />

1000<br />

800<br />

600<br />

400<br />

200<br />

0<br />

0 0.5 1 1.5 2 2.5<br />

initial trans. velocity (m/sec)<br />

Figure 2.9: <strong>The</strong> x f displacement function for the two-dimensional barbell and a graph of<br />

its level curves. <strong>The</strong> level curves are at 6 cm intervals.<br />

1200<br />

thf level curves<br />

thf (deg)<br />

1000<br />

500<br />

1200 0<br />

800<br />

400<br />

wo (deg/sec)<br />

0<br />

0<br />

1<br />

2<br />

vo (m/sec)<br />

3<br />

initial rot. velocity (deg/sec)<br />

1000<br />

800<br />

600<br />

400<br />

200<br />

0<br />

0 0.5 1 1.5 2 2.5<br />

initial trans. velocity (m/sec)<br />

Figure 2.10: <strong>The</strong> f displacement function (net rotation) for the two-dimensional barbell<br />

and a graph of its level curves. <strong>The</strong> level curves are at 13 degree intervals.


2.2. NONAXISYMMETRIC CASE 23<br />

yf (m)<br />

x 10 −3<br />

20<br />

10<br />

0<br />

1200<br />

800<br />

400<br />

wo (deg/sec)<br />

0<br />

0<br />

1<br />

2<br />

vo (m/sec)<br />

3<br />

initial rot. velocity (deg/sec)<br />

1200<br />

1000<br />

800<br />

600<br />

400<br />

200<br />

yf level curves<br />

0<br />

0 0.5 1 1.5 2 2.5<br />

initial trans. velocity (m/sec)<br />

Figure 2.11: <strong>The</strong> y f displacement function for the two-dimensional barbell and a graph its<br />

level curves. <strong>The</strong> level curves are at 1.2 mm intervals.<br />

2.2.3 A practical solution<br />

<strong>The</strong> monotonicity of the level curves of the x f and f displacement functions permits the<br />

use of the axisymmetric solution to solve for the initial velocities for a given translation<br />

along the x axis and for a given rotation. However, this assumes that the initial translational<br />

velocity lies on the x axis. <strong>The</strong>re will also be some translation along the y axis.<br />

This suggests the use of a one dimensional numerical root finding method in which<br />

the independent variable is the direction of the velocity. <strong>The</strong> required translational displacement<br />

can be projected onto the axis defined by the direction of the initial translational<br />

velocity, and the axisymmetric solution can be used to solve for this projected translation<br />

and the desired rotation. <strong>The</strong> resulting initial velocities determine the amount of translation<br />

normal to the initial velocity direction. <strong>The</strong> difference between this actual normal<br />

translation and the required normal translation to reach the goal is then taken as the value<br />

of the function whose root we wish to find.<br />

Alternatively, the translation normal to the initial translational velocity can simply be<br />

ignored. In the examples I have tried, the normal translation is usually much smaller than<br />

the translation in the direction of the initial translational velocity. For the two-dimensional<br />

barbell of the previous subsection, y f can be nearly 32% of x f (for smaller displacements),<br />

is 0.0261, which corresponds to a deviation of 1.5 degrees<br />

from the direction of the initial translational velocity — less than the typical errors due to<br />

variations in impact and friction, as the single tap experiments of Chapter 6 show. Deviations<br />

in direction of this magnitude can effectively be dealt with using planning and<br />

feedback control strategies such as those used in the positioning experiments of Chapter 7.<br />

but the average value of of jyfj<br />

x f


24 CHAPTER 2. THE INVERSE SLIDING PROBLEM


Chapter 3<br />

<strong>The</strong> Impact Problem<br />

Once the inverse sliding problem has been solved, determining the required initial velocities<br />

to achieve a desired displacement, the question that remains is how to generate those<br />

initial velocities (if possible) via impact. This is the impact problem.<br />

<strong>The</strong> solution to this problem must consider not only the mechanics of impact but also<br />

the shape of the object. This chapter begins with a discussion of the mechanics of impact<br />

and different impact models, but most of the details concerning impact have been left to<br />

Appendix A in order that this chapter can focus on the issues regarding object shape.<br />

At any given contact point on an object, there will be some range of directions in which<br />

translational velocity can be produced by an impact. Each of these directions will have a<br />

different lever arm and will thus produce a proportionately different amount of rotational<br />

velocity. In order to produce a given rotational and translational velocity, the boundary of<br />

the object must be searched for a point which has appropriate lever arm for a translational<br />

velocity in the desired direction. This chapter describes a method to represent all velocities<br />

that can be generated by striking a given object, and how the boundary search can be<br />

conducted in terms of this representation.<br />

3.1 Impact models<br />

<strong>The</strong> collision between two rigid bodies (assumed to contact at a single point) is analyzed<br />

through a progression of phases.<br />

Once the two objects have made contact, the compression phase begins; in this phase,<br />

small local (elastic and plastic) deformations occur which gradually deliver impulse to<br />

the objects until their relative normal velocity reaches zero. <strong>The</strong>n the restitution phase<br />

begins, and the “decompression” of the contact area (due to elastic deformation) adds more<br />

impulse to the bodies, propelling them apart, until the collision ends.<br />

Impact is characterized by a coefficient of restitution e which, at the simplest level,<br />

represents the amount of elasticity of the collision. When e =0, the collision is perfectly inelastic<br />

(or plastic); when e =1, the collision is perfectly elastic. This coefficient is generally<br />

assumed to be between 0 and 1 and is largely a property of the different materials involved<br />

in the collision. Although it is often presumed to be a constant, the coefficient of restitution<br />

25


26 CHAPTER 3. THE IMPACT PROBLEM<br />

does vary with the velocity of the impact.<br />

Collision was first studied for direct central impact (i.e. the COMs and the contact normal<br />

lie on the same line, and there is no relative tangential velocity). Early experiments<br />

typically involved the collision of two spheres. Newton’s law (for impact) defines the coefficient<br />

of restitution to be , v, v+<br />

, the ratio of the (normal) velocity after impact to that before<br />

impact. This law is undisputed for direct central impact, and all other proposed restitution<br />

hypotheses simplify to Newton’s law under these conditions.<br />

If the objects collide with some tangential velocity (in addition to normal velocity),<br />

then there will be tangential impulse due to some phenomena; this also affects the collision<br />

process. Most often, Coulomb friction is taken to apply at the contact, so some small relative<br />

motion during a “sliding” mode of the collision is postulated. Other effects, such as<br />

tangential and torsional restitution, are sometimes held to apply instead.<br />

Under Coulomb friction, the tangential impulse will oppose the direction of the relative<br />

sliding motion between the two objects, and the tangential impulse is related to the<br />

normal impulse by a coefficient of friction. If enough tangential impulse is delivered to<br />

the objects that the relative sliding velocity becomes zero, then the collision enters either a<br />

“sticking” mode, in which the relative tangential velocity remains zero, or a “reversed sliding”<br />

mode in which the objects will slide in the opposite direction. Whether this transition<br />

to “sticking” or “reversed sliding” occurs in the compression or restitution phase can make<br />

a difference in the outcome of the collision.<br />

Newton’s law has been shown to inappropriately yield energy increases under collisions<br />

in which enter a sticking or reversed sliding mode. An alternative to Newton’s<br />

law is Poisson’s hypothesis. Poisson defined the coefficient of restitution as the ratio of<br />

the normal impulse delivered during restitution to the normal impulse delivered during<br />

compression. Poisson’s hypothesis and Newton’s law yield identical results for impacts in<br />

which there is no slip or in which slip continues through the entire impact.<br />

Stronge [57] has proposed the internal dissipation hypothesis in which the square of<br />

the coefficient of restitution is defined to be the ratio of elastic strain energy released during<br />

restitution to the energy absorbed by deformation during compression. This hypothesis is<br />

also identical to Newton’s law for impacts in which there is no slip or in which slip continues<br />

through the entire impact. Stronge shows that Poisson’s hypothesis dissipates too<br />

much energy for perfectly elastic collisions involving a reversed sliding mode. <strong>The</strong> internal<br />

dissipation hypothesis results in a collision that terminates at some point in between where<br />

Poisson’s hypothesis and Newton’s law would terminate the collision.<br />

In this thesis, I have adopted Poisson’s hypothesis for several reasons. Since I will be<br />

applying this analysis to physical experiments, there will be no materials with a coefficient<br />

of restitution of 1. Poisson’s hypothesis also lends itself more easily to the analysis of this<br />

thesis than Stronge’s hypothesis, though not quite as easily as Newton’s law would.<br />

Although the choice of a restitution hypothesis affects the particular striker parameters<br />

to achieve a given impulse, it does not (at least among these three hypotheses) affect the<br />

range of impulses which can be achieved via impact at a given contact point.<br />

Routh [54] presented a graphical method for analyzing frictional impact that was more<br />

recently presented in Wang and Mason [63]. This method traces the progress of the impact<br />

in impulse space, using various lines to represent the relative accrual of normal and tan-


3.2. IMPACT PROBLEM CONSTRUCTS 27<br />

−1<br />

tan µ r<br />

r<br />

v so<br />

striker<br />

β<br />

Figure 3.1: Notation for the impact problem. A striker with velocity v s0 at an angle of<br />

incidence with respect to the contact normal collides with the object at a point ~r. By<br />

varying the values of v s0 and , the collision can deliver any impulse within a friction cone<br />

in impulse space. <strong>The</strong> resulting translational velocity will be in the same direction as the<br />

impulse, so it will lie within the shaded friction cone in the contact frame. (Coefficient<br />

of friction between the striker and the object is denoted by r .) <strong>The</strong> resulting rotational<br />

velocity will be determined by the impulse and the vector ~r.<br />

gential impulse in different impact modes as well as to represent constraints such as the<br />

boundary between the compression and restitution phases. Wang and Mason present results<br />

for Newton’s and Poisson’s laws of restitution.<br />

In Appendix A, I have summarized Wang and Mason’s results and use them as a<br />

starting point for showing some properties of impact.<br />

3.2 Impact problem constructs<br />

From Appendix A, this chapter takes the following result as a starting point:<br />

For any contact point on the object, an impact can generate any impulse within<br />

a friction cone in impulse space, i.e. any impulse in which the tangential component<br />

divided by the normal component is less than or equal to the coefficient<br />

of friction between the striker and the object.<br />

This is illustrated in Figure 3.1. <strong>The</strong> details of determining the actual striker parameters to<br />

generate a particular impact are left to the appendix.<br />

This result comes with two assumptions about the striker:<br />

<strong>The</strong> striker has no rotational velocity; thus there are two parameters — the initial<br />

velocity of the striker and the angle of incidence at which it collides with the object.<br />

Either the striker’s COM always lies on the contact normal or the striker has infinite<br />

mass or angular inertia.<br />

<strong>The</strong>se issues are discussed in greater depth in Appendix A.


28 CHAPTER 3. THE IMPACT PROBLEM<br />

3.2.1 From impulse to velocities<br />

<strong>The</strong> impulse generated through impact is related to the initial velocities of the object by:<br />

M~v 0 = ~P (3.1)<br />

I! 0 =(~r ~ P ) z (3.2)<br />

where ~r is the vector from the COM to the contact point. <strong>The</strong> direction of the impulse<br />

determines the direction of the linear velocity, and the contact point determines the lever<br />

arm with which this impulse acts on the object.<br />

Note that ~v 0 and ! 0 are both proportional to the magnitude of the impulse. Since the<br />

constraint on possible impulses is on the direction of the impulse, not the magnitude, it<br />

makes sense to consider the ratio of velocities:<br />

! 0<br />

j~v 0 j = M I (~r ^P ) z (3.3)<br />

which is a function of the direction of the impulse and the contact point.<br />

At a given contact point, a continuous range of ~P (such as all the vectors in a friction<br />

cone at the contact point) will result in a continuous range of !0<br />

j~v 0j<br />

. However, each of these<br />

velocity ratios pertains to a velocity in a different direction. Since we generally want to<br />

produce a velocity in a given direction, we must consider other contact points that can<br />

produce velocity in this direction; these other points will generally have different !0<br />

ratios<br />

j~v 0j<br />

for translational velocity in the desired direction.<br />

3.2.2 Directed velocity ratio sets & impact cones<br />

In order to determine the set of velocity ratios achievable in a given velocity direction, we<br />

must search the boundary of the object to find all points at which an impact can produce<br />

a translational velocity in that direction. This subsection shows how this search can be<br />

“precompiled” into a mapping which I denote by K.<br />

Let the boundary of the object be parameterized by , anddefineS() to be the set of<br />

all directions in which impulse can be generated at the point , i.e.thesetofallvectorsin<br />

the friction cone at the contact point. <strong>The</strong> set of velocity ratios for translational velocity in<br />

the direction ^v 0 is given by:<br />

K(^v 0 )= M<br />

I (~r() ^v 0) z <br />

<br />

9 ^v 0 2S()<br />

(3.4)<br />

<strong>The</strong> set K(^v 0 ) for smooth convex objects will consist of a continuous range of !0<br />

j~v 0j<br />

for any<br />

given ^v 0 . If there are any points on the boundary at which the object should not (or cannot)<br />

be tapped, they can simply be excluded from consideration when forming the K mapping.<br />

When K(^v 0 ) consists of a continuous range of possible velocity ratios, it is convenient<br />

to represent this set as an impact cone in velocity space (the ! 0 v 0 plane for translational<br />

velocity in a particular direction). This cone consists of all vectors for which !0<br />

2K(^v j~v 0j 0)<br />

and thus contains all points in velocity space that can be reached in a single tap.


3.3. EXAMPLES 29<br />

3.2.3 From velocities to impulse<br />

Given some desired initial velocities ! 0 and ~v 0 , the condition:<br />

! 0<br />

j~v 0 j 2K(^v 0) (3.5)<br />

must be satisfied in order to generate those velocities via impact. <strong>The</strong> impulse that will<br />

generate these velocities is ~P = M~v 0 , however, the contact at which to apply this impulse<br />

must still be determined. This will require a search of the the boundary of the object for a<br />

point which can produce the proper !0<br />

j~v 0j<br />

in the right direction. For simple boundary shapes,<br />

it may be feasible to construct the contact point mapping:<br />

<br />

B( ! 0<br />

j~v 0 j ; ^v !<br />

0)= <br />

0<br />

<br />

j~v 0 j = M R j~r() ^v 0j<br />

(3.6)<br />

Given a contact point and the required impulse, the methods in Appendix A can be used<br />

to determine the striker parameters v s0 and .<br />

3.3 Examples<br />

This section develops the mapping used to solve the impact problem for several sample<br />

objects. Illustrations of these objects and the sets K(^v 0 ) are shown in Figure 3.2.<br />

3.3.1 Circular object<br />

For a circular object, the set K is independent of ^v 0 and is given by:<br />

K(^v 0 )=<br />

", MR <br />

p r<br />

; MR <br />

I 1+<br />

2<br />

r<br />

p r<br />

I 1+<br />

2<br />

r<br />

#<br />

(3.7)<br />

where r is the coefficient of friction between the striker and the object. If we parameterize<br />

the boundary by an angle from some reference axis and the direction of ^v 0 by the angle <br />

from the same axis, then we can define the contact point mapping:<br />

3.3.2 Square object<br />

sin,1 B( ! 0<br />

j~v 0 j ;)= + + !0<br />

j~v 0 j<br />

I<br />

MR<br />

<br />

(3.8)<br />

K() =8><<br />

>:<br />

,<br />

MR<br />

sin( 3 I 4<br />

,<br />

MR<br />

sin( 5 I 4<br />

,<br />

MR<br />

sin( 7 I 4<br />

,<br />

MR<br />

sin( 9 I 4<br />

) MR<br />

, ) ; sin( 3 I<br />

+ jjtan ,1 <br />

4 r<br />

) MR<br />

, ) ; sin( I<br />

+ j , 4 2 jtan,1 r<br />

) MR<br />

, ) ; sin( , I<br />

4<br />

, ) ;<br />

MR<br />

I<br />

sin( , 3 4 )<br />

j , j tan ,1 r<br />

j , 3 2 jtan,1 r<br />

; otherwise<br />

(3.9)


30 CHAPTER 3. THE IMPACT PROBLEM<br />

Wo/|Vo|<br />

10<br />

7.5<br />

5<br />

2.5<br />

-2.5<br />

-5<br />

-7.5<br />

-10<br />

Wo/|Vo|<br />

Pi/2 Pi 3Pi/2 22 Pi Pi theta<br />

η<br />

v<br />

40<br />

20<br />

-20<br />

Pi/2 Pi 3Pi/2 22 Pi Pi theta<br />

η<br />

v<br />

-40<br />

Figure 3.2: Directed velocity ratio mappings. <strong>The</strong>se plots show the mapping K, which<br />

contains all possible velocity ratios that can be achieved by impact as a function of the<br />

direction of the resulting translational velocity (parameterized by orientation .


Chapter 4<br />

Planning & Controllability<br />

<strong>The</strong> previous two chapters focused separately on sliding and impact in order to solve the<br />

two parts of the problem for planning a single tap. In this chapter, those results are combined<br />

in order to address planning for multi-tap plans. This will naturally lead to a discussion<br />

of the controllability of tapping that serves to precisely characterize this mode of<br />

manipulation.<br />

This chapter addresses only objects with an axisymmetric support distribution. Since<br />

there seems to be no simple characterization of the possible displacements for objects with<br />

nonaxisymmetric support distributions, these objects do not lend themselves to multi-tap<br />

planning. Even with axisymmetric objects, it does not make sense to extrapolate too many<br />

steps into the future because of the inherent sensitivity of impact and of friction on a dynamically<br />

sliding object.<br />

<strong>The</strong> two parts of this chapter address axisymmetric and nearly axisymmetric objects<br />

respectively. <strong>The</strong> former class considers circular objects which obey the axisymmetric mechanics.<br />

<strong>The</strong> latter class is aimed at objects whose sliding motion is reasonably modeled<br />

with the axisymmetric mechanics but whose boundary is not circular; it is simply the combination<br />

of a more general approach to impact with the axisymmetric sliding mechanics.<br />

For each of these cases, the backprojection of the set of possible displacements is constructed<br />

and used to show that any configuration can be reached in at most two taps; then<br />

the controllability of the respective class of objects is discussed.<br />

4.1 Axisymmetric case<br />

<strong>The</strong> main construct used for planning under the axisymmetric mechanics is the reachable<br />

displacement cone. This cone exists in displacement space and contains all the displacements<br />

that can be reached in a single tap.<br />

In Section 3.2.2, the impact cone was introduced as a representation of all possible initial<br />

velocities that can be produced by impact in a given translational velocity direction.<br />

This cone consists of a continuous range of rays in initial velocity space. In Section 2.1.6 it<br />

was shown that under the axisymmetric sliding mechanics, a cone in the space of initial velocities<br />

can be mapped to a cone in displacement space. Putting these two results together<br />

31


32 CHAPTER 4. PLANNING & CONTROLLABILITY<br />

θ<br />

goal<br />

start<br />

x<br />

Figure 4.1: Planning for axisymmetric objects. <strong>The</strong> reachable displacement cones for each<br />

direction are placed at the start configuration, the backprojections at the goal configuration.<br />

(For circular axisymmetric objects, these cones are identical.) <strong>The</strong> region of intersection is<br />

the set of all points that may be used as an intermediate configuration in getting from the<br />

start to the goal in two taps.<br />

yields the reachable displacement cone.<br />

<strong>The</strong> impact cone for a circular object is symmetric about the v 0 axis (see Section 3.3.1).<br />

Since the v 0 axis maps to the x f axis (pure translational velocity results in a pure translation),<br />

the reachable displacement cone will be symmetric about the x f axis in displacement<br />

space. Because the object is circular, any displacement within this cone can be achieved in<br />

any given translational direction.<br />

<strong>The</strong> backprojection of the reachable displacement cone, i.e. the set of states from which<br />

the current state can be reached with one tap, is simply the reflection of the reachable displacement<br />

cone through the current state. Because of the symmetry of the reachable displacement<br />

cone about the x axis for this case, the backprojection for tapping in one direction<br />

is the same as the reachable displacement cone for tapping in the opposite direction.<br />

4.1.1 Planning<br />

Since the reachable displacement cone is invariant with respect to orientation, the problem<br />

of positioning an axisymmetric object in the plane can be reduced to positioning the object<br />

along the line connecting the COMs in the start and goal configurations.<br />

A single tap will be able to produce the desired translation, but may not be able to<br />

produce enough rotation. Consider the problem in displacement space; assume the origin<br />

corresponds to the start state and place the reachable displacement cone for both translation<br />

directions at the origin. Place the backprojections of the reachable displacement cones<br />

for both directions at the goal state. <strong>The</strong> intersection of these pairs of cones is the set of all<br />

states that can be used as an intermediate state to get from the start to the goal state in two<br />

taps. <strong>The</strong>se constructs are illustrated in Figure 4.1.<br />

It is also not difficult to see that the Minkowsky sum of a reachable displacement cone<br />

for translation in one direction with that for translation in the opposite direction covers the


4.2. NEARLY AXISYMMETRIC CASE 33<br />

θ<br />

x<br />

Figure 4.2: <strong>The</strong> set of configurations reachable in two taps without leaving a neighborhood<br />

of the start configuration includes a (smaller) neighborhood of the start configuration, so<br />

circular axisymmetric objects are small time locally controllable.<br />

space of possible displacements.<br />

4.1.2 Controllability<br />

We have adopted a number of definitions from nonlinear control theory in order to be<br />

precise about the capabilities of modes of manipulation. A system is controllable if it can<br />

reach any goal configuration from any start configuration. A system is small time locally<br />

controllable if, for any given neighborhood about the start configuration, the system can<br />

reach a neighborhood of the start configuration without leaving the given neighborhood.<br />

For our purposes, the “small time” part of small time local controllability does not<br />

refer to time; our interest in this property is in its characterization of the maneuverability<br />

of an object. A mode of manipulation which is small time locally controllable can follow<br />

any path arbitrarily closely.<br />

An axisymmetric object is obviously controllable since it has been shown that any<br />

configuration can be reached with at most two taps. For small time local controllability,<br />

consider the set of states that can be reached without leaving a given neighborhood by<br />

tapping once in the positive x direction and then tapping once in the negative x direction.<br />

As illustrated in Figure 4.2, this set covers the central portion of the given neighborhood,<br />

which includes a smaller neighborhood of the start state. Since this can be done in any<br />

translation direction, a circular axisymmetric object is small time locally controllable.<br />

4.2 Nearly axisymmetric case<br />

For objects that do not have a circular boundary, the K mapping, relating ^v 0 to a set of<br />

! 0<br />

j~v 0j<br />

, is more complex. As in the axisymmetric case (since we assume axisymmetric sliding<br />

mechanics for this object), a range of !0<br />

j~v 0j<br />

can be mapped to a reachable displacement cone.<br />

However, this cone will be different for different ^v 0 , and in general, it will not be symmetric<br />

about the x axis.


34 CHAPTER 4. PLANNING & CONTROLLABILITY<br />

In order to fully appreciate the possible displacements for such an object, we would<br />

need to create a rather complex construct: take a graph of K (such as those in Figure 3.2) and<br />

convert the !0<br />

f<br />

j~v 0j<br />

rangeineachverticalslicetoa<br />

j~x f j<br />

range. Wrap the resulting graph around<br />

a unit cylinder along the axis in displacement space. <strong>The</strong> set of possible displacements<br />

consists of all points along all rays emanating from the origin and passing through a point<br />

of this cylindrical graph. In addition to having a complicated shape, this set will rotate as<br />

the object rotates (as it moves up and down in the direction in displacement space).<br />

<strong>The</strong> ability to reach any configuration with at most two taps can be shown with a<br />

much more restrictive set of actions — tapping the object in a single translational velocity<br />

direction such that the impact cone for this direction includes both positive and negative<br />

values of ! 0 . This will involve tapping the object at a range of contact points in order<br />

toproducearangeof !0<br />

j~v 0j<br />

in the given velocity direction. This restriction will make the<br />

displacement space constructs considerably simpler and easier to manipulate.<br />

4.2.1 Planning<br />

<strong>The</strong> approach taken to planning is essentially the same as that for the axisymmetric case.<br />

First the set of all configurations that can reach the goal in one tap is constructed; I refer<br />

to this construct as the “spiral staircase.” <strong>The</strong> spiral staircase is then intersected with the<br />

reachable displacement cone at the start configuration. <strong>The</strong> result is a set of intermediate<br />

points which may be used to reach the goal configuration in two taps.<br />

Constructing the spiral staircase<br />

Since the object may only be tapped in one direction (relative to its orientation), the orientation<br />

of the object and its two dimensional position must now be taken into account.<br />

Figure 4.3 shows a reachable displacement cone in the three dimensional displacement<br />

space. This cone is in a plane defined by the direction in which the object can be tapped<br />

and the axis. This cone may not be symmetric because of impact geometry.<br />

Note that if the object is tapped and there is some net rotation, then the object will<br />

point in a different direction. <strong>The</strong> reachable displacement cone for the next tap will also<br />

point in that direction (and its origin will be higher or lower in the direction).<br />

Assume that the origin of the coordinate system is placed at the goal configuration<br />

and that the x axis is aligned with the direction of tapping of the object, thus defining the<br />

zero orientation.<br />

<strong>The</strong> spiral staircase will be defined for values of between 0 and ,2. A given <br />

slice of the spiral staircase contains all points at that orientation that can reach the goal<br />

configuration in one tap. First consider the =0slice; the states with the same orientation<br />

as the goal that can reach the goal in a single tap are simply those directly behind the<br />

goal, as illustrated in Figure 4.4. <strong>The</strong>se states form a line from the goal configuration off to<br />

infinity in the xy plane.<br />

Now consider the = , slice. Objects in configurations corresponding to this plane<br />

4<br />

are oriented , relative to the goal configuration, so the reachable displacement cone is<br />

4<br />

pointed in that direction. <strong>The</strong> set of states in this slice which can reach the goal configuration<br />

also lie on a line through the axis. As illustrated in Figure 4.5, the closest point is


4.2. NEARLY AXISYMMETRIC CASE 35<br />

θ<br />

x<br />

y<br />

Figure 4.3: <strong>The</strong> set of reachable configurations for a nearly axisymmetric object which can<br />

only be tapped in one direction. <strong>The</strong> reachable displacement cone lies in a plane parallel to<br />

the axis and in the direction in which the object can be tapped. <strong>The</strong> arrow in the xy plane<br />

represents the object and its orientation.<br />

θ<br />

x<br />

y<br />

Figure 4.4: <strong>The</strong> set of states with the same orientation as the goal configuration that can<br />

reach the goal configuration in a single tap are those directly behind the goal. <strong>The</strong> reachable<br />

displacement cone is shown for one of those points.


36 CHAPTER 4. PLANNING & CONTROLLABILITY<br />

θ<br />

x<br />

y<br />

∆θ<br />

∆θ<br />

Figure 4.5: <strong>The</strong> set of states with a different orientation than the goal configuration that can<br />

reach the goal in a single tap fall on a line through the origin in a plane parallel to the xy<br />

plane. Here, the reachable displacement cone is shown for the closest of these points in one<br />

slice; all points behind it can also reach the goal in one tap.<br />

the one for which the goal state lies on the upper edge of the reachable displacement cone.<br />

<strong>The</strong> rest of the points in this slide lie on the line directly behind this point.<br />

<strong>The</strong> spiral staircase can be mathematically described by the points:<br />

(x; y; ) =(k cos ;k sin ;) for 2 [0; ,2];k2 [,s;,1) (4.1)<br />

where s is the slope of the upper edge of the reachable displacement cone. This set forms a<br />

surface that spirals around the axis. See Figure 4.6 for an illustration.<br />

Solving for intermediate states<br />

Consider where the reachable displacement cone at the start state encompasses all orientations<br />

from ,2 to 0. This will happen at a certain distance from the start state which I will<br />

refer to as the spanning distance. (See Figure 4.7.)<br />

<strong>The</strong> spiral staircase will sweep through all points (x; y) at some orientation except in a<br />

vicinity of the goal state (because of the varying offset of the cone from the axis).<br />

Consider projecting both the spiral staircase about the goal configuration and the<br />

reachable displacement cone from the start configuration onto the xy plane as illustrated in<br />

Figure 4.8. <strong>The</strong> spiral staircase covers the entire plane except for a region enclosed by a spiral<br />

near the goal configuration. <strong>The</strong> reachable displacement cone becomes a ray originating<br />

from the start configuration.<br />

<strong>The</strong> reachable displacement cone and the spiral staircase must intersect at all points<br />

(x; y) along the collapsed reachable displacement cone which are further away from the<br />

start configuration than the spanning distance and which lie on the collapsed spiral staircase.<br />

<strong>The</strong> value of the actual intersection point is determined by the spiral staircase. <strong>The</strong>re<br />

may be points closer to the start configuration that are also in this intersection, but the determining<br />

these intersection points is more difficult. Figure 4.9 shows an illustration of a<br />

spiral staircase and the reachable displacement cone in three dimensions.


4.2. NEARLY AXISYMMETRIC CASE 37<br />

Figure 4.6: An illustration of the spiral staircase, the set of states with orientations from<br />

,2 to 0 which can reach the goal state in a single tap. Each ray in the spiral staircase<br />

extends to infinity, though this is not shown here.<br />

θ<br />

0<br />

-2π<br />

spanning<br />

distance<br />

Figure 4.7: <strong>The</strong> reachable displacement cone will span all orientations from ,2 to 0 beyond<br />

some minimum distance (the spanning distance) from the cone origin.<br />

spanning<br />

distance<br />

start<br />

goal<br />

Figure 4.8: <strong>The</strong> spiral staircase about the goal configuration and the reachable displacement<br />

cone from the start configuration are shown here projected into the xy plane. <strong>The</strong>se constructs<br />

must intersect at some value for all (x; y) points along the reachable displacement<br />

cone whose distance from the start position is greater than the spanning distance.


38 CHAPTER 4. PLANNING & CONTROLLABILITY<br />

Figure 4.9: An illustration of the three dimensional intersection between the spiral staircase<br />

and the reachable displacement cone. As this illustration hints, there may be intersection<br />

points closer to the start configuration than the spanning distance.<br />

Any of the points in this intersection may be used as an intermediate point to reach<br />

the goal configuration in two taps.<br />

4.2.2 Controllability<br />

This system is also obviously controllable, since any goal configuration can be reached with<br />

at most two taps. However, with a single translation direction, it is not small time locally<br />

controllable.<br />

Lynch [40] showed that a pushed object is small time locally controllable if the set<br />

of generalized velocity directions along which the object can be pushed positively span a<br />

great circle of the velocity sphere which is not the ! =0plane. Similarly, tapping a nearly<br />

axisymmetric object is small time locally controllable if the possible displacements directions<br />

positively span a great circle of the unit sphere in displacement space that does not<br />

lie in the =0plane. This condition could be satisfied with as few as two contact points or<br />

two translation directions. This condition determines the number of fixed actuators (fixed<br />

either relative to the object or relative to the world) that are required in order to position<br />

an object arbitrarily. <strong>The</strong> actual number required will depend on the object geometry and<br />

thetypeofactuator.


Chapter 5<br />

Generating Impact<br />

Designing devices to generate impact has been an important part of the experimental effort<br />

of this thesis. In the course of this experimental work, I have used a variety of different<br />

approaches and devices for tapping objects.<br />

This chapter first discusses the background and philosophy behind our efforts in tapping<br />

device design. Through consideration of existing devices to generate impact and preliminary<br />

experiments, I arrived at the goal of designing a specialized tapping device that<br />

creates a single controlled repeatable impact. <strong>The</strong> remainder of the chapter presents the<br />

designs of the five tapping devices used in this thesis, relates experimental observations of<br />

their performance, and discusses the issues I have found to be important in tapping device<br />

design.<br />

5.1 Background and philosophy<br />

Impact is a fairly common phenomena, used in many different ways. <strong>The</strong> designs and<br />

design criteria for the tapping devices used in this thesis were influenced by the methods of<br />

generating impact and applications of impact from many different fields, so a brief survey<br />

is presented here.<br />

Coarse uses of impact can be found in fields such as mining (pneumatic or percussive<br />

drills, stopers, and drifters), destruction (jackhammers and wrecking balls), and construction<br />

(pile drivers and hammer drills). In most of these areas, variations in the impulse<br />

delivered by an impact is not critical; often the objective is simply to deliver the largest<br />

possible impact in order to crush materials.<br />

Many sports and games involve impact, including golf, baseball, basketball, racquet<br />

sports, soccer, cricket, football, croquet, and billiards. In these activities, the objective of the<br />

impact is some combination between delivering a large impact and exerting fine control<br />

over the impact. Generally it is desired to control the direction of an object in addition to<br />

its velocities.<br />

Many musical instruments involve impact, e.g. the xylophone, the piano, and most<br />

percussive instruments. Here the objective is to exert fine control over the strength and<br />

contact point of the impact in order to excite vibratory modes of some resonator.<br />

39


40 CHAPTER 5. GENERATING IMPACT<br />

From this brief summary, it appears that mechanical means of propelling free flying<br />

masses towards objects to generate impact are only used for generating large impacts, and<br />

compliant linkages (mostly human limbs, possibly with some tool) are always used when<br />

precisely controlled impacts are required.<br />

One possible exception is the action of a piano. Although a person is used to press<br />

the keys, the impact is generated by a linkage that essentially throws the hammer (which<br />

rotates about a pivot) at the strings and catches it after it rebounds. Another such device is<br />

the doorbell; it produces a controlled impact using an electromagnetic coil to accelerate a<br />

mass on a spring.<br />

5.1.1 Methods of generating motion and impact<br />

Although impact is usually generated by propelling a free rigid mass at some object, there<br />

are other alternatives, such as electromagnetic forces, using a burst of air, striking an object<br />

directly with some linkage, or using a reaction mass. In regard to the subject of this thesis,<br />

the key criterion is whether initial velocities can be imparted to an object in a sufficiently<br />

short time, thus minimizing the interaction of the forces that produce these initial velocities<br />

with the forces that govern the dynamics of the object motion. Such interaction complicates<br />

analysis of the system.<br />

One implicit focus of this work is on robotic or mechanical implementation for manipulating<br />

parts, materials, or assemblies. <strong>The</strong> primary purpose of these items is generally<br />

not “to be manipulated”, so they may have limitations on how they can be fitted for manipulation<br />

processes. Consequently, I have pursued the most general approach to creating<br />

impulse: collision between a striker and an object.<br />

Striking an object directly with some linkage is undesirable because transmissions to<br />

create motion of a linkage are generally stiff; also, impact tends to destroy bearings. This<br />

brings us back to the most common method of creating impulse: propelling a striker at an<br />

object.<br />

This leads to the question of how motion is generated. Four basic categories of methods<br />

to generate motion are:<br />

electromagnetic forces (electric motors, solenoids, and such),<br />

gravitational forces (falling objects),<br />

mechanical processes ( springs and compressed substances), and<br />

chemical/electrical/mechanical processes (explosions, muscles, piezoelectrics).<br />

In the course of this experimental work, I have explored the first three categories.<br />

5.1.2 Other impact generating devices for manipulation<br />

Higuchi [29] used an electromagnetic coil to induce eddy currents in a conductive plate<br />

in order to generate impulse. A condenser was used to store electrical energy and was<br />

discharged by connecting it to the electromagnetic coil. By varying the amount of energy<br />

stored in the condenser and by varying the capacitance of the condenser, Higuchi was


5.1. BACKGROUND AND PHILOSOPHY 41<br />

able to generate one dimensional displacements of an object with a ratio of minimum to<br />

maximum displacement of approximately 1:25 1 , which corresponds to an energy ratio of<br />

1:5.<br />

Yamagata and Higuchi [65] created an impact generating device, partly based on<br />

Higuchi’s earlier work, that uses a piezoelectric element as an actuator. This device consists<br />

of two masses, a main body which rests on a support surface and a reaction mass which<br />

is attached to a side of the main body by the piezoelectric element (and does not rest on<br />

the support surface). Rapid expansion of the piezoelectric element delivers an impulse to<br />

the main body of the motion mechanism. Slow contraction of the piezoelectric brings the<br />

reaction mass back to its starting position without producing forces on the main body that<br />

exceed static friction. Yamagata and Higuchi show results from an experiment which produced<br />

displacements of 4nm and from an experiment which produced displacements of<br />

2m to5m; however, it is not clear whether these experiments were done with the same<br />

object.<br />

5.1.3 Preliminary tapping experiments<br />

Preliminary experiments with tapping suggested the use of specialized tapping devices<br />

and helped identify several key criteria in their design.<br />

<strong>The</strong>se preliminary experiments used an Adept 550 robot to generate impact. At this<br />

time, I was using plywood parts sliding on a plywood surface. <strong>The</strong> striker was a small<br />

plywood disk held in the gripper of the robot. In order to generate impact, the Adept was<br />

commanded to accelerate to a specified velocity, striking the object with the small disk.<br />

By videotaping the impact, I found that this method did not work well for striker<br />

velocities below 400 mm/sec. For these velocities, the robot appeared to push the object<br />

for a significant distance. For velocities greater than 400 mm/sec, the impact was “clean”,<br />

but resulted in large displacements.<br />

<strong>The</strong>re was significant variation in the displacement of the object, though it was not<br />

clear how much of this was due to variations in the impact and how much to variations in<br />

the sliding. Other preliminary experiments in which the striker velocity could be measured<br />

(done with the first spring-loaded tapping device described later in this chapter) revealed<br />

that there was significant variation in both the impact and sliding process.<br />

5.1.4 Design goals, criteria, and constraints<br />

<strong>The</strong> goal in designing a tapping device for impulsive manipulation is to create a repeatable<br />

device that generates a single clean controllable impact.<br />

A clean impact is important to minimize the interaction between the dynamics of the<br />

striker and the dynamics of the object. Essentially, this means making the collision time as<br />

short as possible so that the striker does not end up pushing the object during an extended<br />

collision.<br />

Multiple impacts simply complicate the impact process, making it less amenable to<br />

analysis and adding more variation to the resulting object velocities.<br />

1 I have estimated a ratio of minimum to maximum displacement of 1:25 from the graphs of (Higuchi [29]).<br />

<strong>The</strong> resulting energy ratio of 1:5 roughly corresponds to his graphs.


42 CHAPTER 5. GENERATING IMPACT<br />

motor<br />

Figure 5.1: <strong>The</strong> first spring-loaded tapping device.<br />

solenoid<br />

It is best to have a dynamic range as large as possible. For positioning tasks, one<br />

would like to reach the vicinity of the goal configuration with a few large taps and then<br />

use very small taps to achieve a high positioning accuracy. My goal was to achieve a ratio<br />

between 1:50 and 1:100 in pure translation (or in energy), which would correspond to a<br />

ratio between 1:7 and 1:10 in striker velocity. It was required that the tapping device should<br />

be able to produce displacements from 1 mm to 15 mm for the set of objects I was using.<br />

<strong>The</strong> tapping devices used in this work have only achieved translation ratios in the range<br />

of 1:15 to 1:30. <strong>The</strong> difficulty is in determining the smallest impact that provides consistent<br />

results. <strong>The</strong> impact can always be made smaller (thus increasing the translation ratio) at<br />

the cost of repeatability.<br />

<strong>The</strong> objects used in the experimental work of this thesis were short flat objects, so it<br />

was essential that the tapper tip was low enough to the surface to tap these objects.<br />

5.2 <strong>The</strong> tapping devices<br />

This section describes the tapping devices designed and built in the course of this thesis in<br />

chronological order.<br />

5.2.1 First spring-loaded tapping device<br />

<strong>The</strong> first spring-loaded tapping device, illustrated in Figure 5.1, is a two stage device. <strong>The</strong><br />

first stage is an aluminum rod with a moderately stiff spring; the second stage is wooden<br />

dowel with a fairly loose spring. A motor and leadscrew move a translation stage back<br />

and forth; a solenoid-actuated latch on this translation stage draws the aluminum rod back<br />

and releases it. <strong>The</strong> size of the tap is adjusted by varying the length of time the motor is<br />

turned on when drawing back the striker. This tapping device was designed to be operated<br />

completely under robot control, so it was important to have some automatic way to reset<br />

the device.<br />

This device uses two stages (a “power” stage and a tapper tip) because of concerns<br />

about the force of a spring acting on the object during and after impact. <strong>The</strong> energy from


5.2. THE TAPPING DEVICES 43<br />

28<br />

striker stage & object positions<br />

2000<br />

striker stage velocity<br />

position (mm)<br />

26<br />

24<br />

22<br />

object<br />

stage 2<br />

stage 1<br />

velocities (mm/sec)<br />

1500<br />

1000<br />

500<br />

0<br />

20<br />

0 0.01 0.02 0.03 0.04<br />

time (sec)<br />

−500<br />

0 0.01 0.02 0.03 0.04<br />

time (sec)<br />

Figure 5.2: Double impact example for the first spring-loaded tapping device.<br />

the power spring is transferred to the tapper tip, and the spring on the tapper tip was<br />

intended to be just strong enough to return the tip to its starting position.<br />

<strong>The</strong> choice of wood for the tapper tip may not have been a good decision; on the<br />

other hand, it did not seem to be a bad choice either. It was originally picked to match<br />

the material of the object. However, wood has slightly unusual friction and deformation<br />

properties due to its fibrous structure.<br />

<strong>The</strong>re are greater difficulties with this device, though. <strong>The</strong> tapper tip is not exactly<br />

spherical, so contact normals differed from what was expected. <strong>The</strong>re is some misalignment<br />

between the axes of the two stages, which made calibration of the tapper more difficult,<br />

and there is perhaps too much play in the in the stages, particularly in the tapper<br />

tip. <strong>The</strong>se factors are undoubtedly responsible for most of the variation in impact in the<br />

experiments with this device.<br />

In later experiments using a high speed measurement system, I found that this device<br />

does often produce double impacts. Figure 5.2 shows one such occurrence.<br />

5.2.2 Pneumatic tapping device<br />

<strong>The</strong> next tapping device, illustrated in Figure 5.3, was pneumatic. Pressurized air propels<br />

a striker along a tube until it strikes the object. A solenoid-actuated valve midway down<br />

the tube remains open in order to relieve pressure behind the striker, so that air pressure<br />

might not continue to exert force on the striker during and after the collision. <strong>The</strong> striker<br />

is then free to rebound. Both the length of a burst of air (controlled by an external solenoid<br />

valve) and the air pressure can be adjusted to affect the striker velocity. In order to reset<br />

the device, the valve is closed, and a venturi generator creates a partial vacuum behind the<br />

projectile.<br />

In practice, the burst of air was very short, in part because the air should be turned off


44 CHAPTER 5. GENERATING IMPACT<br />

6"<br />

detail of rear block<br />

(rear view, 1.5X)<br />

3.25"<br />

air for<br />

vacuum<br />

air for<br />

propulsion<br />

Figure 5.3: <strong>The</strong> pneumatic tapping device. External solenoid valves control the flow of air<br />

to the two ports on this device. One port provides air to propel the tapper tip; the other<br />

produces a partial vacuum using a venturi generator to pull the tapper tip back.<br />

by the time the the striker passes the relief valve. <strong>The</strong> pulse sent to the external solenoid<br />

valve was approximately 5 ms, and the range of air pressure used was from 20 to 40 psi.<br />

It was easy to generate a range of impulses by varying the air pressure; there was a lesser<br />

degree of control through varying the length of the burst of air. A longer tube would<br />

expand the range of velocities achievable by varying the length of the burst of air; however,<br />

there would also be greater friction as the striker slides to the front of the tube.<br />

This device has the advantage of being much more compact than the first springloaded<br />

tapper, although it requires external “plumbing” and hardware. One disadvantage<br />

of this device is that the steel striker pits the aluminum objects used, possibly invalidating<br />

the impact model.<br />

5.2.3 Electromagnetic tapping device<br />

<strong>The</strong> electromagnetic tapping device is illustrated in figure 5.4. This device is a brass tube<br />

with five coils. <strong>The</strong> striker is a short section of steel shaft which is free to move in the tube.<br />

Coils are turned on in sequence in order to accelerate the projectile as well as to draw the<br />

projectile back to the end of the tube. <strong>The</strong> front four coils are used for accelerating the<br />

striker; the back-most coil serves to position the striker in its starting position.<br />

<strong>The</strong>re is no feedback on the position of the striker within the tube, so the coils must be<br />

sequenced open loop. <strong>The</strong> timing of the coils must be adjusted so that as the striker passes


5.2. THE TAPPING DEVICES 45<br />

Figure 5.4: <strong>The</strong> electromagnetic tapping device. Five coils are used to accelerate and draw<br />

back a steel striker inside a brass tube.<br />

a coil, that coil is turned off so that it pulls the striker backwards as little as possible.<br />

5.2.4 Wrecking ball tapper<br />

<strong>The</strong> first three tapping devices were designed to be end effectors of a robot and operated<br />

under robot control. As I began a more detailed investigation of the models and mechanics<br />

used in this thesis (the results of which are in Chapter 6), it became clear that I needed a<br />

simpler more repeatable device for generating impact. Appropriateness for use as a robot<br />

end effector was not the primary concern.<br />

<strong>The</strong> resulting device, which I refer to as “the wrecking ball”, is illustrated in Figure 5.5.<br />

It consists of a hard steel ball bearing suspended from above. It is drawn back with a<br />

thread to a bracket; the velocity and angle of incidence of the wrecking ball can be varied<br />

by adjusting this bracket. A loop of the draw string is passed through a hole in the bracket<br />

and a metal pin used to hold it there until release.<br />

<strong>The</strong> advantages of this device are its simplicity, its repeatability, and the ability to<br />

easily and accurately vary its angle of incidence. <strong>The</strong> wrecking ball is spherical, so there<br />

is no problem with variation in the contact normal. Although there is some disturbance<br />

due to the release of the wrecking ball, the variation in initial object velocities produced by<br />

this tapping device was far less than with the previous tapping devices. Furthermore, it is<br />

possible to calculate the initial velocity of the striker based on measurements of its change<br />

in height.<br />

This device was only used to tap objects at the edge of the support surface because the<br />

wrecking ball is too large to clear the surface and strike the side of an object. <strong>The</strong> wrecking<br />

ball does slightly dent aluminum objects, but not nearly as much as the pneumatic tapper,<br />

probably because of the much larger radius of the wrecking ball.<br />

5.2.5 Second spring-loaded tapping device<br />

<strong>The</strong> final tapping device is illustrated in Figure 5.6. This spring-loaded device consists<br />

of a single stage and was designed to be reconfigurable so that minor modifications and<br />

additions could easily be made. It was intended to be used as an end effector, although<br />

not under automatic control. <strong>The</strong> striker is a rod which is pushed back and secured with<br />

a simple latch. A stop can be manually adjusted in order to vary the compression of the<br />

spring.<br />

Taking some lessons from the design of previous tapping devices, this device was built<br />

with much greater precision and features linear ball bearings. <strong>The</strong> striker is more massive


46 CHAPTER 5. GENERATING IMPACT<br />

18"<br />

Figure 5.5: <strong>The</strong> “wrecking ball” tapper. An adjustable bracket (both in angle and in length)<br />

is used to pull back a steel ball bearing suspended from above by a thread.<br />

that that of previous designs (113.8 g versus 5.5 g for the first spring-loaded tapper and<br />

29.3 g for the wrecking ball). <strong>The</strong> tip of the striker is a softened steel ball bearing, so that<br />

the contact normals would be reasonably precise.<br />

This device has proved to be very repeatable, and makes a nice sound upon impact.<br />

<strong>The</strong>re is a limit on the angle of incidence — beyond 15 to 20 degrees, the impact excites<br />

some vibrational mode of the rod which makes an unpleasant sound and produces a much<br />

smaller impulse.<br />

When calibrating this tapping device, it is important to leave some distance between<br />

the points at which the spring stops acting on the striker and the striker collides with the<br />

object. This distance should be minimized because there is friction in the linear bearings<br />

that then slows the striker. However, if this distance is too short, it is possible for the striker<br />

to rebound after the collision, bounce against the spring, and strike the object again. This<br />

has not been a problem in practice.<br />

Most likely due to the ball bearings, this striker seems to be modeled well under the<br />

assumption that it has infinite angular inertia. <strong>The</strong> position and velocity profiles for this<br />

striker, an example of which are shown in Figure 5.7, show a particularly clean impact.<br />

5.3 Issues in tapping device design<br />

This section summarizes the issues I have found to be important through background reading<br />

and through the design of the tapping devices used in this thesis.<br />

Release mechanisms<br />

For any device which is powered by stored mechanical potential energy, the design of<br />

a release mechanism will influence the repeatability of the device. In these devices, the<br />

striker is subject to some compression or gravitational forces but is prevented from moving<br />

by some (usually kinematic) constraint. <strong>The</strong> goal of a release mechanism is to remove this<br />

constraint quickly while minimizing the effect of (and variations in the effect of) the release<br />

upon the object.<br />

For kinematic constraints, this description suggests a number of standard solutions.


5.3. ISSUES IN TAPPING DEVICE DESIGN 47<br />

15"<br />

top<br />

side<br />

11 mm dia.<br />

collar<br />

foam<br />

bottom<br />

Figure 5.6: <strong>The</strong> second spring-loaded tapping device. <strong>The</strong> strength of the impact can be<br />

adjusted by moving the spring backstop. <strong>The</strong> striker rod is cocked and released by hand,<br />

using a simple latch.<br />

25<br />

striker position<br />

600<br />

striker velocity<br />

position (mm)<br />

20<br />

15<br />

10<br />

5<br />

0<br />

velocity (mm/sec)<br />

400<br />

200<br />

0<br />

−5<br />

0 0.1 0.2 0.3 0.4<br />

time (sec)<br />

−200<br />

0 0.1 0.2 0.3 0.4<br />

time (sec)<br />

Figure 5.7: Position and velocity profiles for the second spring-loaded tapping device.


48 CHAPTER 5. GENERATING IMPACT<br />

Quick release and minimizing the effect of the release suggest using two sharp edges sliding<br />

against each other. Of course, sharp edges are prone to wear, so using hard materials<br />

will improve the repeatability over time. Components of a release mechanism should be<br />

stiff to reduce variations due to the speed of the release. Clearance between parts should be<br />

reduced where appropriate to avoid unwanted motion. Friction can be reduced through<br />

choice of materials or use of lubricants. Finally, conditions where jamming or wedging<br />

might occur should be avoided. More novel release mechanisms might involve electrorheological<br />

fluids or piezoelectric elements.<br />

Another type of constraint is due to some larger force, such as electromagnetic force,<br />

keeping the striker immobilized. I considered using an electromagnet for the wrecking ball<br />

release, but I did not pursue this because I feared gradual magnetization of the parts would<br />

affect repeatability.<br />

Calibration<br />

<strong>The</strong>re are two types of calibration required for a tapping device: geometric calibration and<br />

striker velocity calibration.<br />

When the tapping device will be used as a robotic end effector, kinematic calibration is<br />

much easier if it is constructed so that the axis of the striker lies on one of the axes of the tool<br />

frame. One related design issues is whether the striker should always contact the object at<br />

the same offset from the tapping device. (Because the second spring-loaded tapping device<br />

uses an adjustable back stop, the spring stops acting on the striker at different points for<br />

different settings.)<br />

Determining the relationship between the parameters of the tapping device and the<br />

striker velocity is difficult to do without some means of measuring the striker velocity directly.<br />

For one dimensional impact, the striker velocity and coefficient of restitution cannot<br />

be distinguished only from measurements of object translation. One alternative is to use<br />

some method of generating motion that produces velocities that can easily be calculated,<br />

such as a linear spring or a pendulum.<br />

Dynamic range<br />

As mentioned Section 5.1.4, it is desirable to have as large a dynamic range as possible<br />

in order to do positioning more efficiently. In addition, a large dynamic range aids in the<br />

ability to manipulate objects of different mass on the same scale of displacements (because<br />

object velocities will be inversely proportional to object mass).<br />

Striker tip material and shape<br />

<strong>The</strong> material of the tip of the tapper in relation to the object material is another factor<br />

to consider. <strong>The</strong> friction between the striker tip and the object will affect the amount of<br />

tangential impulse that can be delivered to the object. <strong>The</strong> relative hardness of the two<br />

materials will affect the amount of plastic deformation in the impact, thus affecting the<br />

coefficient of restitution.


5.3. ISSUES IN TAPPING DEVICE DESIGN 49<br />

<strong>The</strong> shape of the striker tip is important, in part because of how it affects local deformations<br />

of the object during the collision, but primarily because impact is sensitive to the<br />

direction of the contact normal. When the angle of incidence of the striker tip is changed,<br />

a spherical tip is the simplest to compensate for.<br />

Striker Mass<br />

<strong>The</strong> mass of the striker in relation to the mass of the objects being manipulated is also<br />

important because it affects the striker velocities required. A light striker will have to be<br />

accelerated to higher velocities to carry the same momentum as a more massive striker.<br />

Inefficiency of impact<br />

Impact can be inefficient due to energy losses through deformation. Small objects can generally<br />

be treated as ideal rigid objects, but larger objects show losses from the shock wave<br />

of an impact traveling through the object. This also has implications in the design of multistage<br />

tapping devices.<br />

Duration of impact<br />

In order to avoid having the forces that accelerate the tapper tip also act on the object<br />

during an extended impact, it is desirable for the tapper tip to be unconstrained in at least<br />

one dimension at impact, i.e. the tapper tip should be free to rebound. However, this can<br />

potentially cause problems with double impacts, another phenomenon which complicates<br />

the impact process.<br />

Method of acceleration<br />

Of the various methods of generating striker motion, the spring-loaded method has been<br />

the easiest to build and use; however, there does not seem to be any inherent reason why<br />

any one method is better than any other.


50 CHAPTER 5. GENERATING IMPACT


Chapter 6<br />

Single Tap Experiments<br />

In order to determine how well the models used in the analysis predict object behavior, I<br />

performed a number of single tap experiments. <strong>The</strong>se experiments involved a variety of<br />

materials, object shapes, and tapping devices. This chapter presents three experiments that<br />

illustrate the development of how the models were adapted to reflect experimental results.<br />

Each experiment consists of a number of runs at different tapper settings, and each<br />

run consists of a number of trials in order to determine the amount of variation for a given<br />

setting. Object trajectories and sometimes striker trajectories were recorded using a high<br />

speed position measurement system. By calculating initial velocities from these measurements,<br />

the effect of tapping device parameters on initial velocities (the impact model) and<br />

the effect of initial velocities on resulting displacements (the sliding model) can be evaluated<br />

separately.<br />

<strong>The</strong>re are two major deviations from the assumptions made for the models used in this<br />

thesis. <strong>The</strong> first is that the support distribution of the objects used in these experiments<br />

is not uniform. Because the actual support distribution depends upon the microscopic<br />

variations in the surfaces in contact, the support distribution is unknown. My hypothesis<br />

was that the actual support distribution would be close enough to uniform that the sliding<br />

model would still be applicable. <strong>The</strong> second deviation is that objects are three dimensional,<br />

not planar. This affects both the sliding model (because it causes a bias in the support distribution<br />

during sliding) and the impact model (because a strike above or below the COM<br />

will produce a moment not parallel to the z axis). My hypothesis was that the objects used<br />

were short enough (close enough to planar) that there would be no significant deviations<br />

from the impact model.<br />

<strong>The</strong> first experiment, the plexiglas disk experiment, showed reasonable agreement between<br />

the (unmodified) sliding model and the experimental results but some discrepancy<br />

with the impact model. <strong>The</strong> second experiment, the aluminum square experiment, showed<br />

reasonable agreement between the impact model and the experimental results but required<br />

a torque scaling factor for the sliding model to reasonably predict object motion. <strong>The</strong> third<br />

experiment, the aluminum disk experiment, also required a torque scaling factor for the<br />

sliding model to match experimental results and showed some discrepancy with the impact<br />

model.<br />

51


52 CHAPTER 6. SINGLE TAP EXPERIMENTS<br />

OptoTrak<br />

OptoTrak<br />

tapping<br />

device<br />

tapping<br />

device<br />

Figure 6.1: Setup for the single tap experiments. <strong>The</strong> OptoTrak measures the position of<br />

infrared LEDs attached to the object and the support surface.<br />

6.1 Experimental materials, setup, and procedures<br />

Figure 6.1 illustrates the experimental setup used in these experiments. An OptoTrak measurement<br />

system looks downward upon the experimental setup to measure the positions<br />

of infrared LEDs attached to the object and the support surface. Two LEDs are used to<br />

measure the position and orientation of the object, three LEDs for the plane of the support<br />

surface. A LED can be attached to the striker (for the spring-loaded tapping devices) in<br />

order to measure the striker velocity. <strong>The</strong> wires from these LEDs can exert small forces on<br />

the object and the striker, so they are carefully arranged and suspended in order to affect<br />

object and striker motion as little as possible.<br />

<strong>The</strong> support surface LEDs are measured at the beginning of each run in order to determine<br />

a local frame. Each run generally consists of 10 trials. During analysis, the object<br />

positions are projected onto the surface plane to measure the direction of the object motion.<br />

This projection was also used to check for motion normal to the surface of the plane,<br />

though none greater than the accuracy of the OptoTrak was observed.<br />

<strong>The</strong> support surfaces were also equipped with some method of aligning the object at<br />

the same starting point; the two methods used are illustrated in Figure 6.2. <strong>The</strong> plexiglas<br />

disk on the aluminum surface was pushed against two pins that were mounted on the<br />

surface; a spacer between the pins and the disk was removed after alignment to keep the<br />

disk from actually touching the pins. For the other experiments, a removable guide with a<br />

number of pins was used.<br />

<strong>The</strong> angle of incidence of the striker could be adjusted easily and fairly precisely with<br />

the wrecking ball tapping device. However, with the second spring-loaded tapping device,<br />

changing the angle of incidence required re-orienting both the tapping device and the support<br />

surface. <strong>The</strong> measurements of the angle of incidence for the wrecking ball are therefore<br />

more accurate than those for the second spring-loaded tapping device. On the other hand,<br />

the striker velocity of the second spring-loaded tapping device could be measured directly


6.1. EXPERIMENTAL MATERIALS, SETUP, AND PROCEDURES 53<br />

Figure 6.2: Object alignment methods. One support surface was equipped with two pins<br />

that were used to place disks in the same start configuration. Other surfaces used a removable<br />

guide which also used several pins for object alignment.<br />

Object shape disk square disk<br />

Object material plexiglas aluminum aluminum<br />

Object mass 52.5 g 205.1 g 245.5 g<br />

Object size r=49mm s = 76.2 mm r=48mm<br />

Surface material aluminum aluminum formica<br />

0.19 0.21 0.17<br />

striker WB WB S2<br />

e 0.54 0.76 0.86<br />

Table 6.1: Summary of materials and parameters for single tap experiments. WB refers to<br />

the wrecking ball tapper, and S2 refers to the second spring-loaded tapper.<br />

by the OptoTrak system, whereas the velocity of the wrecking ball was estimated based on<br />

its change in height.<br />

Table 6.1 shows the combinations of object, surface, and striker used for the experiments<br />

in this chapter.<br />

6.1.1 Measurement issues<br />

<strong>The</strong> OptoTrak was used to measure the position of the object at 500 Hz. <strong>The</strong> stated position<br />

accuracy of the system is 0.1 mm. <strong>The</strong> “baseline” of the two markers on the object ranged<br />

from 62.9 mm to 83.4 mm, which translates to an angular accuracy ranging from 0.18 to<br />

0.14 degrees. With data measured at 500 Hz, this could lead to errors in the calculated<br />

translational velocity of 100 mm/sec and in the calculated rotational velocity from 180<br />

to 140 deg/sec (using a first order difference to compute velocities).<br />

In practice, the data are better behaved than these errors might suggest, but there can<br />

still be significant noise in velocity profiles, particularly in angular velocity during small<br />

rotations.<br />

<strong>The</strong> analysis of these data requires measurement of the initial velocities of the object.<br />

For the experiments presented in this chapter, the initial velocities were taken as the maxima<br />

of the velocity profiles (calculated by a first order difference), with the exception of the


54 CHAPTER 6. SINGLE TAP EXPERIMENTS<br />

ω 0<br />

E<br />

D<br />

C<br />

F, G<br />

B<br />

A<br />

v 0<br />

Figure 6.3: Intended sampling of the space of initial conditions for runs A–G of the plexiglas<br />

disk experiment.<br />

plexiglas disk experiment.<br />

In the plexiglas disk experiment, there was some slight problem with object overhang<br />

(discussed in Section 6.5) which resulted in a small spike in the rotational velocity profiles.<br />

<strong>The</strong> initial translational velocities were measured automatically, but the initial rotational<br />

velocities were estimated manually from the calculated rotational velocity profiles.<br />

6.1.2 Measuring material parameters<br />

<strong>The</strong> question at hand is how well the model predicts the experimental data; however, before<br />

the model can be used, some of the data are required to determine material parameters:<br />

the coefficient of friction between the object and the support surface<br />

e the coefficient of restitution between the striker and the object<br />

r the coefficient of friction between the striker and the object<br />

<strong>The</strong> values of and e are determined using data from pure translational trials. <strong>The</strong> slopes<br />

of lines fit to the translational velocity profiles were averaged in order to determine . <strong>The</strong><br />

calculated or measured striker velocity and the measured initial translational velocity are<br />

used to determine e.<br />

<strong>The</strong>re is no simple way to measure the value of r directly with the object and striker;<br />

however, r should be close to the coefficient of friction between the object and aluminum<br />

— around 0.15 to 0.20 for the object materials used here. For the circular objects and the<br />

tapping devices used in this chapter, the impact is not directly dependent upon the value<br />

of r . So long as the impulse ratio Px<br />

P y<br />

is less than r then the value of r does not affect the<br />

impact process.<br />

6.2 Plexiglas disk experiment<br />

This experiment used the wrecking ball to tap a plexiglas disk sliding on an aluminum<br />

surface. <strong>The</strong>re were seven runs (lettered A–G), each of which contained ten trials. <strong>The</strong><br />

settings of the tapping device were determined in an attempt to sample the space of initial<br />

velocities as illustrated in Figure 6.3. Although the tapper parameters were adjusted to<br />

try to achieve the desired initial velocities, there is enough variation in initial velocities


6.2. PLEXIGLAS DISK EXPERIMENT 55<br />

that it was difficult to achieve these velocities exactly. <strong>The</strong>se nominal initial velocities were<br />

chosen in order to experimentally test the monotonicity properties described in Chapter 2.<br />

Figure 6.4 shows the net displacements for this experiment as well as the average<br />

displacement and standard deviation for each run. (Note that run C is included in both<br />

graphs and tables.) <strong>The</strong>re is a significant amount of variation most runs due to variations<br />

in both impact and sliding.<br />

6.2.1 Evaluation of the sliding model<br />

This subsection examines how well the sliding mechanics model the motion of the disk.<br />

First, the results of the monotonicity test are discussed, and then a general evaluation of<br />

the sliding model is done by a comparison of the actual and predicted displacements and<br />

by a comparison of the actual and predicted velocity profiles.<br />

Montonicity conclusions<br />

Figures 6.5 and 6.6 show graphs of the level curves of x f and f over the relevant space<br />

of initial velocities. Examining these graphs reveals that over the velocity ranges of this<br />

experiment, the expected monotonicity effect in x f would be less than 0.5 mm, in f it<br />

would be about 2 deg. However, the noise in the measured translational displacements is<br />

generally greater than the size of this effect. (See Figure 6.4.) For rotational displacements,<br />

the monotonicity effect can be observed for differences in translational velocity of more<br />

than 40–50 mm/sec.<br />

Comparison of displacements<br />

Figures 6.5 and 6.6 show graphs of the measured initial velocities in addition to plots of<br />

the level curves of x f and f . <strong>The</strong>se graphs show a large variation in the initial velocities<br />

of the object — variation due to the impact process. A comparison of the initial velocities<br />

graph with the graph of the level curves shows that even if the sliding model were perfect,<br />

we would expect a lot of variation in the object displacement because of these variations in<br />

initial velocities.<br />

In order to evaluate the sliding model independently of the impact process, the actual<br />

displacement of the object must be compared with the predicted displacement given the<br />

measured initial velocities. It is this difference that will show the accuracy of the sliding<br />

model.<br />

Table 6.2 shows the results of this comparison. In the worst case, the average translation<br />

difference is 5.5% of the average measured distance. It is not as meaningful to make<br />

a similar comparison for rotation since some of the rotations are rather small. For runs<br />

C–G, the average rotation difference ranges from 4.8% to 26.8% of the average measured<br />

rotation.<br />

Comparison of velocity profiles<br />

<strong>The</strong> solution to the inverse sliding problem described in Chapter 2 can be used to calculate<br />

what the initial velocities must have been to produce a given displacement. By comparing


56 CHAPTER 6. SINGLE TAP EXPERIMENTS<br />

rotation (deg)<br />

8<br />

7<br />

6<br />

5<br />

4<br />

3<br />

2<br />

1<br />

0<br />

−1<br />

Net Displacements, runs A−D<br />

C<br />

−2<br />

23 24 25 26 27 28 29 30<br />

translation (mm)<br />

A<br />

D<br />

B<br />

rotation (deg)<br />

11<br />

10<br />

9<br />

8<br />

7<br />

6<br />

5<br />

4<br />

3<br />

E<br />

Net Displacements, runs E, C, F, G<br />

C<br />

2<br />

15 20 25 30 35 40 45 50 55 60<br />

translation (mm)<br />

Run A B C D<br />

x x x x x x x x<br />

Angle of incidence (deg) 0 10 20 30<br />

Calculated striker velocity (mm/sec) 543 560 594 642<br />

Average distance (mm) 27.8 1.5 26.8 1.2 25.4 1.2 27.2 1.0<br />

Average rotation (deg) 0.9 0.3 -2.3 0.6 -4.1 0.9 -5.9 1.1<br />

Run E C F G<br />

x x x x x x x x<br />

Angle of incidence (deg) 30 20 10 10<br />

Striker velocity (mm/sec) 578 594 700 754<br />

Average distance (mm) 20.9 0.9 25.4 1.2 44.4 2.5 52.9 1.5<br />

Average rotation (deg) -4.1 0.6 -4.1 0.9 -6.2 1.6 -7.7 1.7<br />

Figure 6.4: Net displacements for the plexiglas disk experiment.<br />

F<br />

G


6.2. PLEXIGLAS DISK EXPERIMENT 57<br />

80<br />

Initial velocities, runs A−D<br />

80<br />

Level curves of X_f (mm)<br />

rotational velocity (deg/sec)<br />

70<br />

60<br />

50<br />

40<br />

30<br />

20<br />

C<br />

B<br />

10<br />

A<br />

0<br />

280 285 290 295 300 305 310 315 320<br />

translational velocity (mm/sec)<br />

D<br />

intial rot. velocity (deg/sec)<br />

70<br />

60<br />

50<br />

40<br />

30<br />

20<br />

10<br />

23 24 25 26 27 28<br />

0<br />

280 285 290 295 300 305 310 315 320<br />

initial trans. velocity (mm/sec)<br />

Figure 6.5: Initial velocities for Runs A–D of the plexiglas disk experiment. Also shown<br />

is a graph of level curves of x f over the same range of initial conditions; this gives some<br />

indication of how much variation in translation is expected due to the variation in initial<br />

velocities.<br />

60<br />

Initial velocities, runs E, C, F, G<br />

60<br />

Level curves of <strong>The</strong>ta_f (deg)<br />

rotational velocity (deg/sec)<br />

55<br />

50<br />

45<br />

40<br />

35<br />

30<br />

25<br />

E<br />

C<br />

F<br />

G<br />

intial rot. velocity (deg/sec)<br />

55<br />

50<br />

45<br />

40<br />

35<br />

30<br />

25<br />

3<br />

4<br />

5<br />

6<br />

7<br />

8<br />

20<br />

260 280 300 320 340 360 380 400 420 440<br />

translational velocity (mm/sec)<br />

20<br />

260 280 300 320 340 360 380 400 420 440<br />

initial trans. velocity (mm/sec)<br />

Figure 6.6: Initial velocities for Runs E, C, F, and G of the plexiglas disk experiment. Also<br />

shown is a graph of the level curves of f over the same range of initial velocities; this<br />

gives some indication of how much variation in f is expected due to variation in initial<br />

velocities.


58 CHAPTER 6. SINGLE TAP EXPERIMENTS<br />

A B C D E F G<br />

Average measured translation 27.8 26.8 25.4 27.2 20.9 44.4 52.9<br />

standard deviation 1.53 1.17 1.18 0.96 0.93 2.48 1.47<br />

Average translation difference 0.2 -1.1 -1.4 -0.4 -1.1 -0.2 1.3<br />

standard deviation 2.4 0.8 0.9 0.9 0.6 2.2 1.8<br />

Average measured rotation -0.8 2.3 4.1 5.9 4.1 6.2 7.7<br />

standard deviation 0.43 0.62 0.88 1.15 0.61 1.57 1.67<br />

Average rotation difference 0.9 0.6 1.1 1.0 0.7 -0.3 -0.6<br />

standard deviation 0.3 0.6 0.6 0.3 0.3 0.9 0.6<br />

Table 6.2: Comparison of displacements for the plexiglas disk experiment. <strong>The</strong> calculated<br />

displacements were the result of a forward simulation based on the measured initial velocities.<br />

This table shows the average difference between the computed displacement and the<br />

measured displacement for each run.<br />

the velocity profiles based upon these calculated velocities with the actual velocity profiles,<br />

the fidelity of the sliding model to the actual object behavior can be evaluated.<br />

This is perhaps the truest test of whether the sliding model accurately reflects the object<br />

behavior, and it avoids any error in measuring initial velocities that would affect the<br />

comparison of displacements.<br />

<strong>The</strong> translational velocity profiles all fit fairly well, but there is some variation in the<br />

rotational velocity profiles. Some profiles fit very well (Figure 6.7), most are acceptable<br />

(Figure 6.8), and some fit poorly (Figure 6.9). <strong>The</strong> angular velocity profiles are noisy enough<br />

that there is no meaningful statistical test of the goodness-of-fit.<br />

A quantitative measure is the difference between the calculated initial velocities and<br />

the measured initial velocities; however this does not provide as much information about<br />

how well the sliding mechanics model the object behavior. Like for the displacement comparison,<br />

this measure is computed for each trial in order to discount the variation due to<br />

the impact process; the results of this comparison are shown in Table 6.3. <strong>The</strong> average difference<br />

in initial translational velocity ranges from 2.4% to 7.5% of the average measured<br />

initial translational velocity; for initial rotational velocity, the range is 5.1% to 19.0%.<br />

6.2.2 Evaluation of the impact model<br />

Table 6.4 shows a comparison between the measured initial velocities and the calculated<br />

initial velocities, based on the calculated striker velocity.<br />

<strong>The</strong> translational velocities match very closely. Two of the predicted initial rotational<br />

velocities are very close, two are much too high, and two are much too low. <strong>The</strong> initial rotational<br />

velocity comes entirely from the tangential impulse, so in two cases the tangential<br />

impulse is about right, in two cases it is too high, and in two cases it is too low. However,<br />

the measured angle of the translational velocity (relative to the contact normal) is not as<br />

large as the model predicts, implying that there is more tangential impulse than what the<br />

model predicts for all but one trial.


6.2. PLEXIGLAS DISK EXPERIMENT 59<br />

400<br />

Run B, trial 6<br />

60<br />

Run B, trial 6<br />

trans. velocity (mm/sec)<br />

300<br />

200<br />

100<br />

0<br />

rot. velocity (deg/sec)<br />

40<br />

20<br />

0<br />

−100<br />

0 0.05 0.1 0.15 0.2 0.25<br />

time (seconds)<br />

500<br />

Run G, trial 2<br />

−20<br />

0 0.05 0.1 0.15 0.2 0.25<br />

time (seconds)<br />

80<br />

Run G, trial 2<br />

trans. velocity (mm/sec)<br />

400<br />

300<br />

200<br />

100<br />

0<br />

rot. velocity (deg/sec)<br />

60<br />

40<br />

20<br />

0<br />

−100<br />

0 0.1 0.2 0.3<br />

time (seconds)<br />

−20<br />

0 0.1 0.2 0.3<br />

time (seconds)<br />

Figure 6.7: Plexiglas disk velocity profiles, measured and calculated. <strong>The</strong>se examples are<br />

representative of trials that showed a good fit between the predicted and measured velocity<br />

profiles. <strong>The</strong> predicted velocity profiles were calculated based on measured displacements.


60 CHAPTER 6. SINGLE TAP EXPERIMENTS<br />

400<br />

Run C, trial 6<br />

80<br />

Run C, trial 6<br />

trans. velocity (mm/sec)<br />

300<br />

200<br />

100<br />

0<br />

rot. velocity (deg/sec)<br />

60<br />

40<br />

20<br />

0<br />

−100<br />

0 0.05 0.1 0.15 0.2 0.25<br />

time (seconds)<br />

400<br />

Run D, trial 1<br />

−20<br />

0 0.05 0.1 0.15 0.2 0.25<br />

time (seconds)<br />

100<br />

Run D, trial 1<br />

trans. velocity (mm/sec)<br />

300<br />

200<br />

100<br />

0<br />

rot. velocity (deg/sec)<br />

80<br />

60<br />

40<br />

20<br />

0<br />

−100<br />

0 0.05 0.1 0.15 0.2 0.25<br />

time (seconds)<br />

300<br />

Run E, trial 3<br />

−20<br />

0 0.05 0.1 0.15 0.2 0.25<br />

time (seconds)<br />

100<br />

Run E, trial 3<br />

trans. velocity (mm/sec)<br />

200<br />

100<br />

0<br />

rot. velocity (deg/sec)<br />

80<br />

60<br />

40<br />

20<br />

0<br />

−100<br />

0 0.05 0.1 0.15 0.2 0.25<br />

time (seconds)<br />

−20<br />

0 0.05 0.1 0.15 0.2 0.25<br />

time (seconds)<br />

Figure 6.8: Plexiglas disk velocity profiles, measured and predicted. <strong>The</strong>se examples are<br />

representative of the bulk of the trials, which showed an acceptable fit between the measured<br />

and predicted velocity profiles. <strong>The</strong> predicted velocity profiles were calculated based<br />

on measured displacements.


6.2. PLEXIGLAS DISK EXPERIMENT 61<br />

400<br />

Run C, trial 10<br />

80<br />

Run C, trial 10<br />

trans. velocity (mm/sec)<br />

300<br />

200<br />

100<br />

0<br />

rot. velocity (deg/sec)<br />

60<br />

40<br />

20<br />

0<br />

−100<br />

0 0.05 0.1 0.15 0.2 0.25<br />

time (seconds)<br />

−20<br />

0 0.05 0.1 0.15 0.2 0.25<br />

time (seconds)<br />

Figure 6.9: Plexiglas disk velocity profiles, measured and predicted. <strong>The</strong>se trials are representative<br />

of the few trials which did not show an acceptable fit between measured and<br />

predicted velocity profiles. <strong>The</strong> predicted velocity profiles were calculated based on measured<br />

displacements.<br />

A B C D E F G<br />

Average measured v 0 300 302 301 306 272 387 413<br />

standard deviation 6.2 3.8 11.6 6.3 5.7 6.7 7.6<br />

Average v 0 difference -21.9 -13.5 -6.3 -12.0 -6.5 -19.0 -30.8<br />

standard deviation 11.4 8.2 6.4 6.3 6.0 10.3 9.6<br />

difference percentage -7.3 -4.5 -2.1 -3.9 -2.4 -4.9 -7.5<br />

Average measured ! 0 -24.7 -46.2 -60.0 -47.2 -40.7 -44.9<br />

standard deviation 8.4 5.5 11.6 5.9 6.7 10.3<br />

Average ! 0 difference -7.5 -4.5 -8.8 -8.1 -6.2 2.1 3.5<br />

standard deviation 2.7 4.6 5.2 3.4 2.8 5.7 3.5<br />

difference percentage 18.2 19.0 13.5 13.2 -5.1 -7.8<br />

Table 6.3: Comparison of calculated (from sliding) and measured initial velocities for the<br />

plexiglas disk experiment. <strong>The</strong> calculated initial velocities were determined by applying<br />

the inverse sliding problem solution to the measured displacements. This table shows the<br />

average velocity difference for each run.


62 CHAPTER 6. SINGLE TAP EXPERIMENTS<br />

A B C D E F G<br />

Angle of incidence 0 10 20 30 30 10 10<br />

Calculated striker velocity 543 560 594 642 578 700 754<br />

Predicted v 0 (mm/sec) 299 304 309 309 278 380 410<br />

Average measured v 0 300 302 301 306 272 387 413<br />

standard deviation 6.2 3.8 11.6 6.3 5.7 6.7 7.6<br />

Predicted ! 0 (deg/sec) 0.0 24.5 51.1 80.8 72.7 30.6 33.0<br />

Average measured ! 0 24.7 46.2 60.0 47.2 40.7 44.9<br />

standard deviation 8.4 5.5 11.6 5.9 6.7 10.3<br />

Predicted direction (deg) 0 2.0 4.1 6.4 6.4 2.0 2.0<br />

Actual direction 0 3.2 5.2 6.7 6.2 3.4 3.2<br />

standard deviation 0.4 0.4 0.4 1.1 0.8 0.3 0.3<br />

Table 6.4: Comparison of initial velocities from impact for the plexiglas disk experiment.<br />

This table compares the predicted initial velocities and velocity direction (based on striker<br />

velocity, coefficient of restitution, and angle of incidence) with the measured values.<br />

6.3 Aluminum square experiment<br />

This experiment used the wrecking ball tapping device to tap an aluminum square sliding<br />

on an aluminum surface. This experiment consisted of five runs (lettered A–E). <strong>The</strong> rotational<br />

velocity profiles are considerably less noisy than in the plexiglas disk experiment;<br />

the primary reason for this is that the rotations are greater (in turn due to the fact that an<br />

impulse can have a greater lever arm due to the geometry of the square). <strong>The</strong> resulting<br />

displacements and the measured initial velocities are shown in Figure 6.10.<br />

I hypothesized that a “nearly” axisymmetric object such as a square would be modeled<br />

reasonably well by the axisymmetric mechanics. In this sense, this experiment was a<br />

combination of the axisymmetric mechanics with a more general impact problem.<br />

A key issue is what size disk to use to model the square. Without examining the force<br />

andtorquefunctions,Idecideduponusingthediskthathadthesametorqueasthesquare<br />

for pure rotation. Setting the expressions for this torque on a disk of radius R and a square<br />

of side s equal to one another, we get:<br />

2<br />

"2<br />

3 gRM = 1 12 gsM p<br />

2 , log<br />

p<br />

2s , s<br />

2<br />

!<br />

+ log<br />

p<br />

2s + s<br />

Given s we can then solve for R. For the square used in this experiment, s =76:2 mm and<br />

R =43:7 mm (87:4 mm in diameter).<br />

Figure 6.11 shows the force and torque functions for the square and for the disk approximation.<br />

<strong>The</strong> relevant range of ! in simulation is generally from 15 to 20, and the<br />

v<br />

difference between the real force and torque functions and those of the equivalent disk are<br />

fairly small in this range.<br />

However, the sliding model does not adequately reflect the motion of the square. In<br />

particular, the predicted rotational velocity profiles are qualitatively different than the ac-<br />

2<br />

!#<br />

(6.1)


6.3. ALUMINUM SQUARE EXPERIMENT 63<br />

rotation (deg)<br />

20<br />

15<br />

10<br />

5<br />

D<br />

C<br />

Displacements, runs A−E<br />

B<br />

E<br />

rotational velocity (deg/sec)<br />

A<br />

50<br />

A<br />

0<br />

11 12 13 14 15 16 17 18<br />

0<br />

180 200 220 240 260 280<br />

translation (mm)<br />

translational velocity (mm/sec)<br />

Run A B C D E<br />

x x x x x x x x x x<br />

Angle of incidence 0 0 -20 20 0<br />

Calculated striker velocity 1.05 1.05 1.05 1.05 1.19<br />

Average distance (mm) 14.7 0.9 13.8 1.1 12.3 0.7 12.8 0.4 16.6 0.5<br />

Average rotation (deg) -1.1 0.5 -13.0 1.2 -9.9 0.4 -14.8 0.4 -16.8 0.5<br />

350<br />

300<br />

250<br />

200<br />

150<br />

100<br />

Initial velocities, runs A−E<br />

Figure 6.10: Displacements and initial velocities for the aluminum square experiment.<br />

D<br />

C<br />

B<br />

E<br />

tual velocity profiles; the predicted rotation is significantly higher than the actual rotation.<br />

Figure 6.12 shows the measured and predicted velocity profiles for two typical trials.<br />

It turns out that scaling the torque by a factor of 2.17 enables the model to reasonably<br />

predict the object motion. <strong>The</strong> use of a scaling factor was suggested by examining a plot<br />

of the measured torque versus the measured ! ratio (all calculated from position data<br />

v<br />

and consequently quite noisy). <strong>The</strong> actual value of this scale factor was computed by an<br />

optimization, using the sum of the squares of displacement differences for each run and<br />

the sum of the squares of velocity differences for each run as objective functions. (<strong>The</strong><br />

displacement metric actually produced an optimal scaling factor of 2.14, the velocity metric<br />

2.20.) <strong>The</strong> remainder of this section uses scaled torque for both inverse sliding problem<br />

calculations and for trajectory simulation.<br />

6.3.1 Evaluation of the sliding model<br />

As in the plexiglas disk experiment, the differences between displacements and between<br />

initial velocities are used as a quantitative measure of how well the sliding model reflects<br />

object motion. However, since the objective function for optimizing the torque scaling is<br />

based on these differences, perhaps the most convincing argument of the suitability of the


64 CHAPTER 6. SINGLE TAP EXPERIMENTS<br />

0.44<br />

force on square<br />

0.015<br />

torque on square<br />

0.42<br />

Newtons<br />

0.4<br />

0.38<br />

0.36<br />

real<br />

disk<br />

Nm<br />

0.01<br />

0.005<br />

scaled<br />

real<br />

0.34<br />

disk<br />

0.32<br />

0 5 10 15 20 25<br />

w/v<br />

0<br />

0 5 10 15 20 25<br />

w/v<br />

Figure 6.11: Force and torque functions on a square object. <strong>The</strong> solid line is the real function;<br />

the dashed line is the function for the equivalent disk. In the torque plot, the dotted line is<br />

the scaled torque which is used later in this section.<br />

sliding model is a comparison of velocity profiles.<br />

Table 6.5 shows the comparison between displacements (measured translation versus<br />

calculated translation based on measured initial velocities). <strong>The</strong> difference in translation<br />

ranges from 1.6% to 11.6% of the average measured translation, the difference in rotation<br />

ranges from 2.9% to 6.6% of the average measured rotation.<br />

Figure 6.13 shows for three representative trials the measured velocity profiles and<br />

the predicted velocity profiles based on the initial velocities calculated from the measured<br />

displacements using the inverse sliding problem solution. <strong>The</strong> match between measured<br />

and calculated profiles for both translation and rotation seems good.<br />

Table 6.6 shows the comparison between initial velocities (measured velocities versus<br />

calculated). <strong>The</strong> difference in initial translational velocity ranges from 1.0% to 6.2% of the<br />

average measured initial velocity, the difference in rotation ranges from 1.2% to 5.3% of the<br />

average measured initial rotational velocity.<br />

6.3.2 Evaluation of the impact model<br />

<strong>The</strong>se runs in this experiment used two different contact points along the edge of the square<br />

as well as different angles of incidence of the striker. Table 6.7 shows the predicted initial<br />

velocities which are calculated from the coefficient of restitution, the striking geometry,<br />

and the estimated striker velocity. For comparison, the table also shows the measured<br />

initial velocities.


6.3. ALUMINUM SQUARE EXPERIMENT 65<br />

250<br />

Run B, trial 3<br />

50<br />

Run B, trial 3<br />

trans. velocity (mm/sec)<br />

200<br />

150<br />

100<br />

50<br />

0<br />

rot. velocity (deg/sec)<br />

0<br />

−50<br />

−100<br />

−150<br />

−200<br />

−50<br />

0 0.05 0.1 0.15 0.2 0.25<br />

time (seconds)<br />

300<br />

Run E, trial 10<br />

−250<br />

0 0.05 0.1 0.15 0.2 0.25<br />

time (seconds)<br />

100<br />

Run E, trial 10<br />

trans. velocity (mm/sec)<br />

200<br />

100<br />

0<br />

rot. velocity (deg/sec)<br />

0<br />

−100<br />

−200<br />

−100<br />

0 0.05 0.1 0.15 0.2 0.25<br />

time (seconds)<br />

−300<br />

0 0.05 0.1 0.15 0.2 0.25<br />

time (seconds)<br />

Figure 6.12: Measured and calculated velocity profiles for the aluminum square experiment<br />

using the unscaled torque. Although the translational velocity matches well, the rotational<br />

velocity is qualitatively different.


66 CHAPTER 6. SINGLE TAP EXPERIMENTS<br />

A B C D E<br />

Average measured translation 14.7 13.8 12.3 12.8 16.6<br />

standard deviation 0.9 1.1 0.7 0.4 0.5<br />

Calculated translation 13.0 12.8 12.1 11.8 17.0<br />

standard deviation 1.1 1.0 0.7 0.6 0.8<br />

Translation difference 1.7 1.0 0.2 0.9 -0.4<br />

standard deviation 0.5 0.3 0.5 0.4 0.4<br />

Translation diff. percentage 11.6 7.2 1.6 7.4 2.4<br />

Average measured rotation -1.1 -13.0 -9.9 -14.8 -16.8<br />

standard deviation 0.5 1.2 0.4 0.4 0.5<br />

Calculated rotation 0 -12.6 -10.4 -13.8 -17.5<br />

standard deviation 0 1.1 0.6 0.6 0.9<br />

Rotation difference -1.1 -0.4 0.4 -1.0 0.7<br />

standard deviation 0.5 0.5 0.6 0.5 0.5<br />

rotation diff. percentage 2.9 3.3 6.6 4.1<br />

Table 6.5: Comparison of displacements for the aluminum square experiment. <strong>The</strong> calculated<br />

displacements were determined from a forward simulation using the measured<br />

initial velocities (and the scaled torque). This table shows the average difference for each<br />

run and its percentage with respect to the average measured values.<br />

A B C D E<br />

Average measured v 0 231 220 217 207 253<br />

standard deviation 10.4 8.3 6.5 5.4 6.0<br />

Average v 0 difference -14.4 -9.3 -2.7 -8.5 2.5<br />

standard deviation 4.8 2.6 3.8 3.7 2.5<br />

difference percentage -6.2 -4.2 -1.2 -4.1 1.0<br />

Average measured ! 0 -241 -202 -284 -292<br />

standard deviation 12.8 9.8 7.5 11.4<br />

Average ! 0 difference 19.7 -2.9 -10.8 7.7 -9.0<br />

standard deviation 8.5 6.5 8.7 6.6 6.5<br />

difference percentage 1.2 5.3 -2.7 3.1<br />

Table 6.6: Comparison of calculated (from sliding) and measured initial velocities. <strong>The</strong> calculated<br />

initial velocities were determined by applying the inverse sliding problem solution<br />

(with scaled torque) to the measured displacements.


6.3. ALUMINUM SQUARE EXPERIMENT 67<br />

250<br />

Run B, trial 3<br />

50<br />

Run B, trial 3<br />

trans. velocity (mm/sec)<br />

200<br />

150<br />

100<br />

50<br />

0<br />

rot. velocity (deg/sec)<br />

0<br />

−50<br />

−100<br />

−150<br />

−200<br />

−50<br />

0 0.05 0.1 0.15 0.2 0.25<br />

time (seconds)<br />

250<br />

Run D, trial 5<br />

−250<br />

0 0.05 0.1 0.15 0.2 0.25<br />

time (seconds)<br />

100<br />

Run D, trial 5<br />

trans. velocity (mm/sec)<br />

200<br />

150<br />

100<br />

50<br />

0<br />

rot. velocity (deg/sec)<br />

0<br />

−100<br />

−200<br />

−50<br />

0 0.05 0.1 0.15 0.2 0.25<br />

time (seconds)<br />

300<br />

Run E, trial 10<br />

−300<br />

0 0.05 0.1 0.15 0.2 0.25<br />

time (seconds)<br />

100<br />

Run E, trial 10<br />

trans. velocity (mm/sec)<br />

200<br />

100<br />

0<br />

rot. velocity (deg/sec)<br />

0<br />

−100<br />

−200<br />

−100<br />

0 0.05 0.1 0.15 0.2 0.25<br />

time (seconds)<br />

−300<br />

0 0.05 0.1 0.15 0.2 0.25<br />

time (seconds)<br />

Figure 6.13: Measured and predicted velocity profiles for the aluminum square experiment.<br />

<strong>The</strong> predicted velocity profiles were computed using the inverse sliding problem solution<br />

(with the scaled torque) and the measured displacements.


68 CHAPTER 6. SINGLE TAP EXPERIMENTS<br />

A B C D E<br />

Striker velocity 1.05 1.05 1.05 1.05 1.19<br />

Strike point 0 -0.506 -0.506 -0.506 -0.506<br />

Angle of incidence 0 0 -20 20 0<br />

Predicted translational velocity (mm/sec) 231 221 206 211 251<br />

Measured translational velocity 231 220 217 207 253<br />

standard deviation 10.3 8.3 6.5 5.4 6.0<br />

Predicted rotational velocity (deg/sec) 0 -266 -218 -283 -302<br />

Measured rotational velocity -241 -202 -284 -292<br />

standard deviation 12.8 9.8 7.5 11.4<br />

Predicted velocity direction (deg) 0 1.6 -2.1 5.2 1.6<br />

Measured velocity direction 0 6.0 -0.1 14.5 6.8<br />

standard deviation 0.5 1.0 0.8 0.5 0.7<br />

Table 6.7: Comparison of initial velocities from impact for the aluminum square experiment.<br />

<strong>The</strong> predicted initial velocities and velocity direction were calculated from the impact<br />

geometry, the striker velocity, and angle of incidence.<br />

6.4 Aluminum disk experiment<br />

This experiment used the second spring-loaded tapping device to tap an aluminum disk<br />

sliding on a formica surface. This experiment consisted of four runs (lettered A–D), testing<br />

a few settings of the tapping device and angle of incidence. <strong>The</strong> displacements for this<br />

experiment and the measured initial velocitieis are shown in Figure 6.14.<br />

<strong>The</strong> primary purpose of this experiment was to determine material parameters of this<br />

disk and surface in order to perform positioning experiments, the subject of Chapter 7.<br />

<strong>The</strong>re were several pure translation ( = 0) trials at different tapper settings in order to<br />

determine a coefficient of restitution. <strong>The</strong>se are not discussed here but are mentioned in<br />

Section 6.5.6 in conjunction with observations on the dependence of the coefficient of restitution<br />

on velocity.<br />

<strong>The</strong> unmodified axisymmetric mechanics did not model the motion of this object very<br />

well, particularly with respect to rotational motions. Like the aluminum square, measured<br />

rotations were much less than predicted. Expanding upon the approach to the aluminum<br />

square experiment, I did an optimization on the coefficient of friction (between the object<br />

and support surface) first and then on the torque scaling factor.<br />

<strong>The</strong> objective function for the coefficient of friction optimization was computed by taking<br />

the difference between the measured initial velocity and the calculated initial velocity<br />

(based on measured displacement), averaging the differences over each run, and summing<br />

and squaring the average differences for each runs. <strong>The</strong> resulting coefficient of friction<br />

was 0.17, slightly lower than what I had originally estimated from the pure translational<br />

velocity profiles. <strong>The</strong> torque scaling factor was optimized next, using a similar objective<br />

function based on initial rotational velocity, and resulted in a scale factor of 3.64.


6.4. ALUMINUM DISK EXPERIMENT 69<br />

rotation (deg)<br />

8<br />

6<br />

4<br />

B<br />

Displacements, runs A−D<br />

C<br />

D<br />

rotational velocity (deg/sec)<br />

40<br />

2<br />

20<br />

A<br />

A<br />

0<br />

10 15 20 25 30 35 40<br />

0<br />

200 250 300 350 400<br />

translation (mm)<br />

translational velocity (mm/sec)<br />

Run A B C D<br />

x x x x x x x x<br />

Angle of incidence (deg) 0 14.5 7 7<br />

Striker velocity (mm/sec) 419 419 431 632<br />

Average distance (mm) 17.4 0.3 15.5 0.9 18.7 1.1 37.1 0.7<br />

Average rotation (deg) 0.5 0.1 4.7 0.7 4.2 0.4 7.6 0.3<br />

120<br />

100<br />

80<br />

60<br />

B<br />

Initial velocities, runs A−D<br />

Figure 6.14: Measured displacements and initial velocities for the aluminum disk experiment.<br />

C<br />

D


70 CHAPTER 6. SINGLE TAP EXPERIMENTS<br />

A B C D<br />

Average measured translation 17 16 19 37<br />

standard deviation 0.3 0.9 1.1 0.7<br />

Calculated translation 18 14 19 40<br />

standard deviation 0.4 0.9 1.6 1.1<br />

Translation difference -0.5 1.7 0.2 -2.7<br />

standard deviation 0.2 0.3 0.6 1.0<br />

Translation diff. percentage -2.6 10.8 1.2 -7.2<br />

Average measured rotation 0.5 4.7 4.2 7.6<br />

standard deviation 0.1 0.7 0.4 0.3<br />

Calculated rotation 0.0 4.1 4.4 8.2<br />

standard deviation 0.0 0.7 0.8 0.3<br />

Rotation difference 0.5 0.6 -0.1 -0.6<br />

standard deviation 0.1 0.4 0.6 0.2<br />

rotation diff. percentage 12.3 -3.2 -7.8<br />

Table 6.8: Comparison of displacements for the aluminum disk experiment. <strong>The</strong> calculated<br />

displacements were calculated from the measured initial velocities using a forward<br />

simulation.<br />

6.4.1 Evaluation of the sliding model<br />

As in the previous experiments, a comparison between displacements and between initial<br />

velocities is used for some quantitative measure of how well the model reflects the object<br />

behavior. Table 6.8 shows the comparison between the measured displacement and the<br />

displacement calculated from the measured initial velocities. <strong>The</strong> average difference in<br />

translation ranges from 1.2% to 10.8% of the average measured translation, and the average<br />

difference in rotation ranges from 3.2% to 12.3% of the average measured rotation.<br />

Figure 6.15 shows a number of representative measured and calculated translational<br />

and rotation trajectories.<br />

6.4.2 Evaluation of the impact model<br />

<strong>The</strong> impact model predicted translational velocities well, but the predicted rotational velocities<br />

were 25% to 50% of the measured initial rotational velocities. In order to have<br />

a usable model of impact for positioning experiments, I took the form of the model (i.e.<br />

the relationship between striker parameters and the resulting impulse, based on parameters<br />

computed from striker and object mass and geometry) and estimated the parameters<br />

based on experimental data.<br />

Specifically in terms of the impact model, I estimated the parameters B 1 and B 2 which<br />

are the ratios between relative object velocity and the resulting impulse for the tangential<br />

and normal directions respectively. Whereas B 2 is fairly consistent over the four runs, B 1<br />

varies considerably, particularly for Run B. <strong>The</strong> average B 2 results in underestimates of the<br />

initial rotational velocity for Runs C and D and an overestimate for Run B, however the


6.4. ALUMINUM DISK EXPERIMENT 71<br />

250<br />

Run B, trial 7<br />

100<br />

Run B, trial 7<br />

trans. velocity (mm/sec)<br />

200<br />

150<br />

100<br />

50<br />

0<br />

rot. velocity (deg/sec)<br />

80<br />

60<br />

40<br />

20<br />

0<br />

−50<br />

0 0.1 0.2 0.3<br />

time (seconds)<br />

250<br />

Run C, trial 7<br />

−20<br />

0 0.1 0.2 0.3<br />

time (seconds)<br />

100<br />

Run C, trial 7<br />

trans. velocity (mm/sec)<br />

200<br />

150<br />

100<br />

50<br />

0<br />

rot. velocity (deg/sec)<br />

80<br />

60<br />

40<br />

20<br />

0<br />

−50<br />

0 0.1 0.2 0.3<br />

time (seconds)<br />

400<br />

Run D, trial 7<br />

−20<br />

0 0.1 0.2 0.3<br />

time (seconds)<br />

150<br />

Run D, trial 7<br />

trans. velocity (mm/sec)<br />

300<br />

200<br />

100<br />

0<br />

rot. velocity (deg/sec)<br />

100<br />

50<br />

0<br />

−100<br />

0 0.1 0.2 0.3 0.4<br />

time (seconds)<br />

−50<br />

0 0.1 0.2 0.3 0.4<br />

time (seconds)<br />

Figure 6.15: Comparison of predicted and measured velocity profiles for the aluminum<br />

disk experiment. <strong>The</strong> predicted velocity profiles were computed based on initial velocities<br />

from the inverse sliding problem solution applied to the measured displacements.


72 CHAPTER 6. SINGLE TAP EXPERIMENTS<br />

A B C D<br />

Average measured v 0 244 214 248 363<br />

standard deviation 2.7 7.0 10.5 5.1<br />

Average calculated v 0 241 226 249 351<br />

standard deviation 2.1 6.2 7.2 3.1<br />

Average v 0 difference 3.5 -12.7 -1.5 12.4<br />

standard deviation 1.4 2.8 3.9 4.6<br />

difference percentage 1.4 -5.9 -0.6 3.4<br />

Average measured ! 0 90.9 83.0 106.4<br />

standard deviation 13.1 12.9 2.9<br />

Average calculated ! 0 9.1 97.7 80.1 102.2<br />

standard deviation 1.2 12.4 5.0 3.1<br />

Average ! 0 difference -9.1 -6.8 2.9 4.2<br />

standard deviation 1.2 9.2 11.9 1.7<br />

difference percentage -7.5 3.5 4.0<br />

Table 6.9: Comparison of initial velocities from sliding for the aluminum disk experiment.<br />

<strong>The</strong> calculated velocities were determined by applying the solution to the inverse sliding<br />

problem to the measured displacements.<br />

predicted values are closer to the measured values than they were before.<br />

Table 6.10 shows these predicted velocities (based on angle of incidence and measured<br />

striker velocities) and the measured object velocities.<br />

6.5 Experimental observations<br />

6.5.1 Striking height<br />

Since the striker may contact the object at a point above, below, or at the same height above<br />

the support surface as the COM, the impact may produce some moment about the x or y<br />

axis (i.e. a moment that would press one side of the object into the surface and lift the<br />

other side up). A few pure translational runs in the plexiglas disk experiment were taken<br />

in order to determine how much the striking height affects object motion.<br />

This experiment used the wrecking ball tapper which was adjusted in between runs so<br />

that for one run, it struck the disk at approximately 1 of the object height above the surface,<br />

4<br />

onerunattheleveloftheCOM, and one run near the top of the object.<br />

Since the wrecking ball delivered a different impulse for each of these three runs, it<br />

is not informative to examine the resulting displacements or initial velocities. Instead, I<br />

present a comparison of the coefficients of friction of these trials. This serves to lump all<br />

effects of the three dimensional behavior into the coefficient of friction.<br />

As Table 6.11 shows, the coefficient of friction was slightly higher for the run in which<br />

the striker struck the object above the COM and slightly lower for the run in which the<br />

contact point was below the COM. However, the differences between the coefficients of


6.5. EXPERIMENTAL OBSERVATIONS 73<br />

A B C D<br />

Striker velocity 419 419 431 632<br />

Angle of incidence 0 -14.5 -7 -7<br />

Predicted translational velocity (mm/sec) 247 240 252 370<br />

Measured translational velocity 244 214 248 363<br />

standard deviation 2.7 7.0 10.5 5.1<br />

Predicted rotational velocity (deg/sec) 0 125 63 92<br />

Measured rotational velocity 90.9 83.0 106.4<br />

standard deviation 13.1 12.9 2.9<br />

Predicted velocity direction (deg) 0 4.9 2.3 2.3<br />

Measured velocity direction 0 12.6 7.0 7.4<br />

standard deviation 0.2 1.2 0.1 0.1<br />

Table 6.10: Comparison of initial velocities from impact for the aluminum disk experiment.<br />

<strong>The</strong> predicted initial velocities were calculated based upon the striker velocity and angle<br />

of incidence and the coefficient of restitution.<br />

<br />

Below COM 0.1778 0.0058<br />

At COM 0.1818 0.0110<br />

Above COM 0.1888 0.0088<br />

Table 6.11: Effect of striker height. This table shows the measured coefficient of friction<br />

from runs where the contact point of the striker was varied with respect to the level of the<br />

COM.<br />

friction are less than the standard deviations of the data. <strong>The</strong> effect of striker height appears<br />

to be small but noticeable.<br />

6.5.2 Object overhang<br />

In some of the first experiments with the wrecking ball tapper, the initial object position was<br />

such that a small portion of the object extended beyond the edge of the support surface.<br />

This was done in order to prevent the possibility of the wrecking ball colliding with the<br />

edge of the support surface and thereby affecting the impact.<br />

I noticed a slight spike followed by a slight dip in many of the velocity profiles, which<br />

I attributed to measurement error and treated as a slight annoyance in measuring initial<br />

velocities. This effect can be observed in the measured velocity profiles for the plexiglas<br />

disk experiment presented in Section 6.2.<br />

An experiment with the aluminum square produced much more dramatic spikes in<br />

the velocity profiles, as shown in Figure 6.16. In fact, a change of resistance could be felt as<br />

one pushed the the object over the edge of the support surface. A later experiment with the<br />

object starting entirely on the support surface (with the wrecking ball carefully adjusted so<br />

that it would not strike the edge of the support surface) showed no such spike.<br />

In the original aluminum square experiment, I expect that this effect was due to the


74 CHAPTER 6. SINGLE TAP EXPERIMENTS<br />

250<br />

Run B, trial 2<br />

200<br />

Run B, trial 2<br />

trans. velocity (mm/sec)<br />

200<br />

150<br />

100<br />

50<br />

0<br />

rot. velocity (deg/sec)<br />

150<br />

100<br />

50<br />

0<br />

−50<br />

0 0.05 0.1 0.15 0.2 0.25<br />

time (seconds)<br />

−50<br />

0 0.05 0.1 0.15 0.2 0.25<br />

time (seconds)<br />

Figure 6.16: Spikes in velocity profiles due to object overhang.<br />

topography of the surface of the object. Perhaps some “lip” of the object was catching on<br />

the edge of the support surface. To some lesser extent, there are probably other effects,<br />

such as some aerodynamic effect, in this and in other experiments.<br />

6.5.3 Object path<br />

<strong>The</strong> path of sliding objects tends to curve, as shown in Figure 6.17. <strong>The</strong> likely explanation<br />

for this is that the COM of the object is above the plane in which friction acts on the<br />

object. This produces a moment about the COM that pushes the front end of the disk downward,<br />

thus increasing the effect of the friction that opposes rotation. A object with positive<br />

rotation then tends to curve towards the right.<br />

6.5.4 Breaking static friction<br />

For the aluminum disk experiment, a number of pure translational trials in which the object<br />

motion and the striker motion were measured allow calculation of striker velocity before<br />

and after impact. Taken with the initial object velocity, the momentum and energy before<br />

and after the impact can be compared.<br />

Table 6.12 shows the momentum comparison. Some momentum is lost during the impact,<br />

presumably due to breaking static friction. This does not seem to be a fixed amount or<br />

fixed percentage of momentum, but one might expect that breaking static friction involves<br />

exerting some force over a distance, resulting in a fixed energy loss instead.<br />

Table 6.13 shows the energy lost during collision. Of course, most of this energy loss is<br />

due to plastic deformation during impact. It is difficult to say how much energy is used in<br />

breaking static friction of whether this energy is constant because it is not easy to determine<br />

both the energy loss and the coefficient of restitution from measured velocities.


6.5. EXPERIMENTAL OBSERVATIONS 75<br />

1<br />

Object paths, aluminum disk experiment, Run A<br />

(mm)<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

0<br />

0 2 4 6 8 10 12 14 16 18 20<br />

(mm)<br />

Figure 6.17: Curvature in object paths. This graph shows object paths for Run A of the<br />

aluminum disk experiment. Ideally, these paths should be straight lines. Instead, the path<br />

curves slightly, likely due to the slight positive rotation observed in this run.<br />

Run Pbefore Pafter P percentage<br />

hi 0.0737 0.0684 0.0054 7.3<br />

0.0006 0.0012<br />

md 0.0476 0.0457 0.0020 4.2<br />

0.0009 0.0012<br />

lo 0.0229 0.0212 0.0017 7.4<br />

0.0011 0.0014<br />

Table 6.12: Momentum comparison. <strong>The</strong> momentum before and after impact for the aluminum<br />

disk experiment is compared. All momenta are in kg m/sec, and the percentages<br />

are in relation to the momentum before the impact.<br />

Run Ebefore Eafter E percentage<br />

hi 0.0239 0.0192 0.0045 18.3<br />

0.0004 0.0004<br />

md 0.0100 0.0088 0.0011 11.0<br />

0.0004 0.0003<br />

lo 0.0023 0.0021 0.0002 8.7<br />

0.0002 0.0002<br />

Table 6.13: Energy comparison. This table compares the kinetic energy of the object and<br />

striker before and after collision.


76 CHAPTER 6. SINGLE TAP EXPERIMENTS<br />

Run v s0 (mm/sec) e <br />

hi 648 0.814 0.025<br />

md 419 0.893 0.039<br />

lo 202 0.887 0.047<br />

Table 6.14: Variation in the coefficient of restitution. This table shows the average computed<br />

coefficient of restitution for three trials using different striker velocities. In accordance<br />

with typical variation, the coefficient of restitution is the lowest for the trial with the<br />

highest striker velocity.<br />

6.5.5 Three dimensional effects<br />

One way to know the exact support distribution is to use three points of support. I did<br />

an experiment using a disk which had been fitted with three rounded “feet” spaced at<br />

120 degree intervals at a common radius. However, this seemed to only exacerbate three<br />

dimensional effects, perhaps because the COM of the object was higher. <strong>The</strong> measured<br />

trajectories, as shown by representative trials in Figure 6.18, did not match the predicted<br />

trajectories. In particular, the predicted trajectories did not reverse direction or stop and<br />

start. My hypothesis is that the disk is actually skittering — there is some combination of<br />

translation and rotation about one of the feet.<br />

6.5.6 Variance in e with velocity<br />

<strong>The</strong> coefficient of restitution does have some dependence upon the velocity of impact because<br />

of varying amounts of plastic deformation with the energy of the impact. Table 6.14<br />

shows the calculated coefficient of restitution for each of three pure translational runs conducted<br />

during the aluminum disk experiment. Despite variance in the average, it is clear<br />

that the run with highest velocity impact has the lowest coefficient of restitution.<br />

6.6 Conclusions<br />

<strong>The</strong>se experiments have shown that the sliding model (modified, if necessary) can predict<br />

object motion within 5–10% for translations and 10–25% for for rotations. <strong>The</strong> impact<br />

model generally predicts initial translational velocities to within a few percent but may be<br />

anywhere from a few percent to 33% off in predicted rotational velocity.<br />

<strong>The</strong> addition of a torque scaling factor is somewhat distressing because there is no<br />

clear explanation why it is required. In addition, this factor has varied widely — no scaling<br />

factor was required for the plexiglas disk experiment; a factor of 2.17 was used for the<br />

aluminum square experiment, a factor of 3.64 for the aluminum disk experiment. <strong>The</strong>se<br />

factors are too large to be accounted for solely by a nonuniform support distribution; the<br />

maximum amount of torque is produced by having all points of support as far from the<br />

COM as possible, and this would only increase the torque by a factor of 2 or less. <strong>The</strong> three<br />

dimensional extent of objects is another possible cause, but the fact that the COM of the<br />

object is not in the plane likely has a small effect on the pressure distribution. It seems


6.6. CONCLUSIONS 77<br />

2<br />

orientation<br />

50<br />

angular velocity<br />

(degrees)<br />

1.5<br />

1<br />

0.5<br />

(deg/sec)<br />

40<br />

30<br />

20<br />

10<br />

0<br />

0<br />

0 0.05 0.1 0.15 0.2 0.25<br />

time (sec)<br />

0<br />

−0.5<br />

orientation<br />

−10<br />

0 0.05 0.1 0.15 0.2 0.25<br />

time (seconds)<br />

20<br />

0<br />

angular velocity<br />

(degrees)<br />

−1<br />

−1.5<br />

−2<br />

−2.5<br />

(deg/sec)<br />

−20<br />

−40<br />

−60<br />

−3<br />

0 0.05 0.1 0.15 0.2 0.25<br />

time (sec)<br />

−80<br />

0 0.05 0.1 0.15 0.2 0.25<br />

time (seconds)<br />

Figure 6.18: Representative trials from the tripod disk experiment. Note how the rotation<br />

slows down and resumes in the first example and how it reverses direction in the second.


78 CHAPTER 6. SINGLE TAP EXPERIMENTS<br />

plausible that frictional torque is caused by some mechanism that differs from classical<br />

mechanics since Coulomb friction is an approximation to a complex nonlinear phenomena.<br />

Additional experiments with more precise objects and surfaces and with faster and more<br />

accurate measurements would be required to investigate this matter further.<br />

<strong>The</strong> impact model seems much better at predicting rotational velocities when they are<br />

produced by normal impulse acting at some lever arm, as was the case for the aluminum<br />

square experiment, instead of by the tangential impulse which is due to friction, as in the<br />

disk experiments. <strong>The</strong> impact model does not predict the direction of the impulse well; this<br />

also bears some relationship to the amount of tangential impulse. In fact, the measured<br />

initial velocities are usually not consistent with an impulse that can be produced at the<br />

contact point, indicating that there may be other processes at work.<br />

Despite these shortcomings, it still appears that the sliding and impact models are capable<br />

of predicting object motion well enough to be used for positioning experiments. Even<br />

if the models perfectly predicted the averages for each trial, there is still considerable variation<br />

due to the physical phenomena. This error, along with any error due to the models,<br />

will have to be addressed by planning and feedback strategies in order to do positioning<br />

experiments.<br />

<strong>The</strong> experiments in this chapter were done with reasonable care. While it is possible<br />

to more carefully fabricate objects, use smoother and more homogeneous support surfaces,<br />

and build more precise tapping devices, i.e. make the world conform better to the assumptions<br />

of the models used, it is not clear that we want to. In order for manipulation by<br />

tapping to be generally and easily applicable, it must address the factors that deviate from<br />

model assumptions but are common in real world applications — three dimensional parts,<br />

variation in the coefficient of restitution, variations in striking height, etc.


Chapter 7<br />

Positioning Experiments<br />

In order to apply the mechanics and the planning approach described in this thesis towards<br />

positioning an object, a number of practical issues must be addressed. <strong>The</strong> planning<br />

methods described in this chapter explicitly consider the range of displacements that the<br />

tapping device can produce as well as the displacement errors due to variations in impact<br />

and friction. Chapter 4 demonstrated that axisymmetric objects can reach any configuration<br />

within two taps but suggested no criteria for choosing among the infinite number of<br />

possible subgoals. Furthermore, it may be desirable to use more than two taps to reach a<br />

goal.<br />

<strong>The</strong> planning methods used in this chapter can be viewed as feedback control strategies<br />

— given the discrepancy between the object’s configuration and the goal configuration,<br />

they determine the first of a sequence of taps that will take the object to the goal.<br />

<strong>The</strong>se planning methods show a progression in how the error model is incorporated, and<br />

each has a different underlying assumption about the relative importance of minimizing<br />

the number of taps, minimizing the length of each tap, and producing a robust plan.<br />

<strong>The</strong> first half of this chapter describes three experiments in positioning a circular object.<br />

<strong>The</strong> first experiment uses two tap plans under the assumption that there are no errors,<br />

the second uses a simple error model to plan two tap plans, and the third uses a slightly<br />

more sophisticated error model to plan multi-tap plans. <strong>The</strong> second half of this chapter<br />

describes an experiment and analysis in support of the conjecture that tapping can be used<br />

to position an object more precisely than the manipulator can position the tapping device.<br />

7.1 Preliminaries<br />

7.1.1 Experimental Setup<br />

As illustrated in Figure 7.1, the positioning experiments were done using an Adept 550<br />

(SCARA) robot to position the second spring-loaded tapping device relative to the object.<br />

An overhead camera was used to determine the object position and orientation by locating<br />

a black circular target on the object which has a white stripe across its diameter; the<br />

accuracy of the vision system is between 0.5 mm and 1.0 mm in position.<br />

79


80 CHAPTER 7. POSITIONING EXPERIMENTS<br />

Figure 7.1: Experimental setup for positioning experiments.<br />

<strong>The</strong> object was the same aluminum disk used in the single tap experiments of Chapter<br />

6. For this object, the second spring-loaded tapping device can produce a minimum<br />

translation of 1.7 mm, and I have set the maximum translation to 35 mm in order to allow<br />

a full rotation range over the entire translation range. This disk slides on a white formica<br />

surface (the surface is a formica-covered particle board bookshelf). By default, the Adept<br />

taps the object from its current location directly towards the goal; in the third experiment,<br />

however, there was the option to tap the object away from the goal (in a direction 180<br />

degrees from the goal).<br />

<strong>The</strong> task for all the experiments in this chapter was to move the object 10 mm with<br />

a rotation of 10 degrees; this is approximately three times as much rotation as can be accomplished<br />

in a 10 mm translation. <strong>The</strong> desired positioning accuracy was 1 degree in<br />

orientation and 1 mm in position.<br />

7.1.2 Error models<br />

I have assumed that the error model is linear although this is difficult to show conclusively<br />

from the single tap experiments of the previous chapter. <strong>The</strong> error model relates the<br />

displacement of a tap to the radii of an error ellipse in configuration space:<br />

<br />

rx<br />

r <br />

<br />

=<br />

<br />

exx<br />

e x<br />

e x<br />

e <br />

<br />

xf<br />

j f j<br />

In general, this error matrix has full rank and would represent a projection like that illustrated<br />

in the leftmost graph of Figure 7.2, reflecting the fact that a translational displacement<br />

primarily results in translational error but will also produce some rotational error,<br />

and likewise with a rotational displacement.<br />

<strong>The</strong> consequence of this projection is that the set of states which have an error ellipsoid<br />

smaller than some given dimensions in rotation and translation forms a nonconvex set. <strong>The</strong><br />

process of intersecting the reachable displacement cone with the acceptable error ellipsoid<br />

dimensions is also shown in Figure 7.2. Since this shape is unwieldy to use in planning, I<br />

first considered a conservative kite-shaped approximation to this set. However even this<br />

approximated shape was awkward to use for planning, so I turned to simpler error models.<br />

<br />

(7.1)


7.1. PRELIMINARIES 81<br />

r<br />

θ<br />

θ<br />

f<br />

r<br />

θ<br />

θ<br />

f<br />

x<br />

f<br />

max r<br />

θ<br />

r<br />

x<br />

max r<br />

x<br />

r<br />

x<br />

x<br />

f<br />

Figure 7.2: <strong>The</strong> error projection and the effect of maximum ellipse radii on the set of reachable<br />

states.<br />

One simplification is to assume that the error matrix is diagonal — r x and r are independent<br />

and proportional to x f and f respectively. However, the single tap experiments<br />

showed that variation in rotation is much greater than that in translation, so in the interest<br />

of simplicity, I have considered only rotational error, which collapses the error ellipsoid to<br />

a line. For the second experiment (conservative two-tap plans), I have assumed an error<br />

matrix of the form: <br />

0 0<br />

<br />

(7.2)<br />

0 e <br />

<strong>The</strong> drawback of this error matrix is that a pure translation, no matter how far, will have<br />

zero rotational error. For the third experiment (multi-tap plans), I have made the more<br />

conservative assumption that the radius of the error ellipsoid is constant for a given<br />

translational displacement and have set the error matrix:<br />

<br />

0 0<br />

<br />

(7.3)<br />

e x 0<br />

to reflect the maximum radius for all possible rotations.<br />

<strong>The</strong> planning methods in this chapter do not explicitly consider error in the direction<br />

of the tap. Since the object is circular and can thus be tapped in any direction, errors in<br />

direction simply contribute to errors in distance to the goal. (Recall that these experiments<br />

are done so that the object is always tapped directly towards (or away from) the goal.)<br />

7.1.3 Positioning figures<br />

<strong>The</strong> figures in this chapter used to graphically illustrate positioning experiments are of two<br />

types: cartesian space and configuration space. In both cases, the start position is marked<br />

with a black dot labeled with the letter S. For each tap, the intended tap is shown by a<br />

dotted line with a gray dot at the end. <strong>The</strong> actual tap is shown with a black line with a<br />

black dot at the end, labeled with the tap number.<br />

In cartesian space plots, the x and y axes are scaled differently in order to clearly show<br />

the path of the object; displacements in the y direction are typically much smaller than that<br />

in the x direction. <strong>The</strong>se plots are always oriented so that the goal lies directly to the right<br />

of the start configuration.


82 CHAPTER 7. POSITIONING EXPERIMENTS<br />

θ<br />

goal<br />

subgoal<br />

start<br />

x<br />

Figure 7.3: <strong>The</strong> exact two-tap planning method takes the closest feasible subgoal.<br />

In configuration space plots, the error ellipsoid is shown about the intended goal configuration<br />

in a dotted line. Although the error ellipsoid has been collapsed to a line for the<br />

second and third planning strategies, the error ellipse is drawn with some radius in the x<br />

dimension. <strong>The</strong> configuration space plots show the object position projected onto the line<br />

connecting the start to the goal configuration, so the actual translational displacement of<br />

taps may be slightly larger than these plots would indicate.<br />

7.2 Planning methods<br />

7.2.1 Exact two-tap plans<br />

<strong>The</strong> first set of experiments assumed that there was no error in actuation, so the closest<br />

subgoal was used in formulate a two-tap plan, as illustrated in Figure 7.3. For a given<br />

object state, if the goal can be reached with one tap, then that tap is executed; if not, the<br />

first tap of a two-tap plan is executed. This rule is repeated until the goal is reached.<br />

<strong>The</strong> drawback of this strategy is that if there is any error in the first tap, the object may<br />

not be within the backprojection from the goal configuration; replanning is then required.<br />

On the other hand, this strategy has the benefit of being very simple.<br />

Figure 7.4 is the best example of this planning method. Although the first tap was<br />

slightly off in direction, the object was still within the backprojection of the goal, and the<br />

second tap brought it to the goal configuration. Not all runs were quite so good; Figure 7.5<br />

shows a more typical run. Here, replanning was done for Taps 2, 3, and 5 (although after<br />

Tap 5, the object was close enough to the goal to terminate the run).<br />

Of the three methods presented in this chapter, this seemed to be the most robust,<br />

consistently taking 5 or fewer taps to position the object.<br />

7.2.2 Conservative two-tap plans<br />

In order to improve upon the exact two-tap planning strategy, the conservative two-tap<br />

planning strategy takes into account the error of the first tap. As shown in Figure 7.6, this


7.2. PLANNING METHODS 83<br />

th (deg)<br />

12.5<br />

2<br />

10.0<br />

1 7.5<br />

5.0<br />

S<br />

2.5<br />

2<br />

S<br />

x (mm) 0.0<br />

0 5 10 15<br />

0 5 10 15<br />

Figure 7.4: Example of an exact two-tap plan.<br />

1<br />

x (mm)<br />

y (mm)<br />

0.0<br />

-0.5<br />

5<br />

4<br />

th (deg)<br />

3<br />

3<br />

10.0<br />

2<br />

7.5<br />

1<br />

5.0 S<br />

5<br />

2.5<br />

2<br />

S<br />

x (mm) 0.0<br />

4<br />

0 5 10 15 20<br />

0 5 10 15 20<br />

Figure 7.5: Example of an exact two-tap plan with replanning.<br />

y (mm)<br />

1 1.5<br />

1.0<br />

0.5<br />

0.0<br />

-0.5<br />

x (mm)<br />

θ<br />

goal<br />

subgoal<br />

start<br />

x<br />

Figure 7.6: <strong>The</strong> conservative two-tap planning method makes sure the error ellipsoid for<br />

the first tap completely fits within the space of feasible subgoals.


84 CHAPTER 7. POSITIONING EXPERIMENTS<br />

4<br />

3<br />

2<br />

S<br />

0 5 10 15 20 25<br />

1<br />

th (deg)<br />

10.0<br />

7.5<br />

5.0<br />

2.5<br />

0.0<br />

x (mm)<br />

2<br />

S<br />

4<br />

3<br />

0 5 10 15 20 25<br />

y (mm)<br />

1.0<br />

1<br />

0.5<br />

0.0<br />

-0.5<br />

-1.0<br />

x (mm)<br />

Figure 7.7: Example of a conservative two-tap plan with replanning.<br />

strategy aims for a point in the set of possible subgoals that completely contains an error<br />

ellipsoid. As discussed in Section 7.1.2, the error model for this experiment results in error<br />

ellipses that have zero radius in the x direction. This strategy guarantees (subject to the<br />

assumptions inherent in this error model) that after the first tap, the object will be in a state<br />

that can directly reach the goal configuration. Like the exact two-tap planning strategy, if<br />

the goal can be reached with a single tap, that tap is executed; otherwise, the first tap of the<br />

two-tap plan is executed, and this process is repeated until the goal is reached.<br />

Surprisingly, this strategy did not perform quite as well as the exact two-tap plans.<br />

Figure 7.7 shows one of the better runs, which only required replanning for Tap 3. Most<br />

runs took about 5 taps.<br />

<strong>The</strong> worst case run for this strategy is shown in Figure 7.8. This points to the problem<br />

with my implementation of this strategy — the tapping accuracy is not as good as the error<br />

model; all of the taps in this trial over-rotated. Generally, the first tap of a two-tap plan<br />

did bring the object within the backprojection from the goal, but after the next tap, another<br />

two-tap plan had to be created. Although the object was repeatedly tapped back and forth,<br />

itdidgetcloserandclosertothegoalconfiguration.<br />

7.2.3 Multi-tap plans<br />

One drawback of the first two strategies is that the final tap to reach the goal is often a large<br />

tap. Since the error scales with the displacement, a better solution is to have a sequence of<br />

taps that decrease in size as the object approaches the goal. In addition, every tap should<br />

be guaranteed (again, within the assumptions of the error model) to leave the object in a<br />

state from which it can reach the next subgoal.<br />

As discussed in Section 7.1.2, the error model for this experiment also results in error<br />

ellipses with zero radius in the x direction.<br />

Planning method<br />

Consider the set of states which can reach the goal in one tap subject to a maximum bound<br />

on the error (in ) and to the minimum displacement that can be produced by the tapping


7.2. PLANNING METHODS 85<br />

5 3<br />

9 7<br />

10<br />

4<br />

8<br />

11<br />

6<br />

2<br />

S<br />

0 5 10 15 20 25<br />

1<br />

x (mm)<br />

th (deg)<br />

12.5<br />

10.0<br />

7.5<br />

5.0<br />

2.5<br />

0.0<br />

S<br />

2<br />

6<br />

48<br />

10 11<br />

9<br />

7 5<br />

0 5 10 15 20 25<br />

3<br />

1<br />

x (mm)<br />

y (mm)<br />

0.5<br />

0.0<br />

-0.5<br />

-1.0<br />

-1.5<br />

Figure 7.8: Example of a conservative two-tap plan with replanning. (Worst case)<br />

device. This set, which I will refer to as Cone 1, forms a symmetric trapezoidal region, as<br />

illustrated in Figure 7.9.<br />

<strong>The</strong> set of states that can reach Cone 1 in a single tap, subject to the same constraints,<br />

is the union of two symmetric trapezoidal regions. This second cone may overlap the first,<br />

depending upon the choice of the acceptable error in and the minimum displacement.<br />

Since any state in Cone 2 that is also in Cone 1 should aim directly for the goal, I consider<br />

Cone 2 to be the set of states outside of Cone 1. This second cone can in turn be backprojected<br />

to form Cone 3. Figure 7.9 illustrates the process of backprojecting these cones, and<br />

Figure 7.10 shows the first four cones together in relationship to each other and the goal.<br />

This sequence of cones is fixed relative to the goal, and by definition, once the object<br />

is within one cone, it can always reach the next cone towards the goal. <strong>The</strong> question recone<br />

1<br />

cone 2<br />

cone 3<br />

goal<br />

cone 1<br />

cone 2<br />

Figure 7.9: Construction of a sequence of cones leading to the goal for multi-tap planning.<br />

<strong>The</strong> backprojection of a cone can be found by examining the backprojection using the<br />

smallest tap and that using the largest tap. <strong>The</strong> backprojected cone will include all states<br />

in between. <strong>The</strong> lightly shaded region is the part of the backprojection that overlaps with<br />

the cone.


86 CHAPTER 7. POSITIONING EXPERIMENTS<br />

goal<br />

cone 1<br />

cone 2<br />

cone 3<br />

cone 4<br />

Figure 7.10: <strong>The</strong> first four cones from the goal to the right. Cones 3 and 4 and the second<br />

half of Cone 2 are vertical slices of the same cone.<br />

10<br />

th<br />

0<br />

-10<br />

0 51015<br />

-50<br />

0<br />

x<br />

50<br />

r_th<br />

Figure 7.11: <strong>The</strong> sequence of cones leading to the goal for multi-tap planning. <strong>The</strong> cones<br />

shrink with increasing error ellipsoid radius r .


7.2. PLANNING METHODS 87<br />

15<br />

10<br />

5<br />

0<br />

-5<br />

-10 th<br />

-15<br />

-20<br />

-25<br />

-30<br />

-80 -60<br />

-40<br />

x<br />

-20<br />

0<br />

151010<br />

5<br />

0<br />

r_th<br />

Figure 7.12: <strong>The</strong> intersection of the extended reachable displacement cone and the sequence<br />

of cones leading to the goal is the set of possible subgoals for the first tap.<br />

maining is how to get from an arbitrary start configuration to one of these cones; again, the<br />

error ellipsoid from that tap should fit entirely within the entry cone.<br />

<strong>The</strong> size of the error ellipsoid, by assumption, varies with the translational displacement<br />

of the object. In order to determine where the object can properly enter a cone, the<br />

radius of the error ellipsoid r must be considered as another dimension to this problem.<br />

Drawing upon principles used in the construction of configuration spaces, we reduce<br />

the size of the sequence of cones to the goal in the direction as r increases. This extension<br />

along the r dimension is shown in Figure 7.11. <strong>The</strong> reachable displacement cone from the<br />

start configuration, which normally exists in the x- plane will now extend linearly into the<br />

r dimension. <strong>The</strong>se constructs are shown together in Figure 7.12.<br />

<strong>The</strong>re is often a set of points in the intersection of these constructs to choose as a subgoal.<br />

I have made particular choices for this experiment: when entering a cone from a point<br />

not on any cone, I chose the point closest to the goal orientation and then the point closest<br />

to the start configuration; when transitioning from one cone to the next, I chose the point<br />

closest to the goal orientation and then the point closest to the middle of the cone.<br />

Experiment<br />

<strong>The</strong> multi-tap planning experiments were the most consistent of all the experiments in this<br />

chapter. With one exception, the plans all took 5 taps and required no replanning. A typical<br />

run is shown in Figure 7.13. From the start configuration, the object reached Cone 4 in the<br />

first tap, and progressed to the next cone closer to the goal for each subsequent tap.<br />

One run which hinted at the drawbacks of this planning method was the first run,<br />

which used a lower limit (0.5 degrees instead of 1.0 degrees) on the allowable orientation<br />

error at the goal; this made all the cones narrower in this run than in the others. This run<br />

is shown in Figure 7.14. Replanning occurred after Tap 3, taking the object away from the<br />

goal in order to get back into the sequence of cones. In Tap 8 (from Cone 1), the object<br />

rotated too much, leaving it outside the sequence of cones. More taps were required to get<br />

it back into a cone. This is, of course, a failure of the error model to accurately reflect the


88 CHAPTER 7. POSITIONING EXPERIMENTS<br />

1<br />

4<br />

2 3<br />

S<br />

-25 -20 -15 -10 -5 0 5 10<br />

S 4<br />

3<br />

1 2<br />

-25 -20 -15 -10 -5 0 5 10<br />

5<br />

x (mm)<br />

5<br />

x (mm)<br />

th (deg)<br />

10.0<br />

7.5<br />

5.0<br />

2.5<br />

0.0<br />

y (mm)<br />

0.0<br />

-0.5<br />

-1.0<br />

Figure 7.13: Example of a multi-tap plan.<br />

behavior of the object, but it underscores the tendency of this model to quickly correct any<br />

slight deviations from the planned sequence of taps.<br />

7.3 High precision positioning<br />

Early in the course of this work, I conjectured that tapping could be used to position an<br />

object more precisely than a manipulator can position a tapping device relative to the object.<br />

<strong>The</strong> intuition was that the amount of impulse delivered to an object is not too sensitive<br />

to the distance of the tapping device from the object. However, variations in the tapping<br />

device position and orientation also affect the contact normal.<br />

This section describes an experiment to test this conjecture and presents some basic<br />

analysis to support it.<br />

7.3.1 Experiment<br />

In this experiment, the exact two-tap planning method was used, and the requested position<br />

and orientation of the tapping device was then rounded with respect to the global<br />

frame. <strong>The</strong> first experiment rounded positions to the nearest 10 mm and orientations to<br />

the nearest 5 degrees. <strong>The</strong> second experiment rounded positions to the nearest 15 mm and<br />

orientations to the nearest 10 degrees. In all trials, the goal was reached (within 1 mm and<br />

1 degree); a typical trial took 5–6 taps to reach the goal. Figure 7.15 shows a typical trial<br />

from the second experiment.


7.3. HIGH PRECISION POSITIONING 89<br />

1<br />

3 8<br />

4<br />

7<br />

5 6<br />

2<br />

9<br />

S<br />

-25 -20 -15 -10 -5 0 5 10 15<br />

th (deg)<br />

10.0<br />

7.5<br />

5.0<br />

2.5<br />

0.0<br />

x (mm)<br />

1 2<br />

3<br />

S<br />

4 5<br />

6<br />

7<br />

9<br />

8<br />

10 11 x (mm)<br />

11<br />

10<br />

y (mm)<br />

1.0<br />

0.5<br />

0.0<br />

-0.5<br />

-1.0<br />

-25 -20 -15 -10 -5 0 5 10 15<br />

Figure 7.14: Example of a multi-tap plan. This run was ended prematurely.<br />

5<br />

4<br />

2<br />

S 1<br />

0 5 10 15 20<br />

6<br />

3<br />

3<br />

th (deg)<br />

10.0<br />

5<br />

1<br />

7.5<br />

5.0S<br />

6<br />

2.5<br />

0.0<br />

x (mm)<br />

2 4<br />

0 5 10 15 20<br />

y (mm)<br />

4.0<br />

3.0<br />

2.0<br />

1.0<br />

0.0<br />

-1.0<br />

x (mm)<br />

Figure 7.15: A high precision positioning trial. <strong>The</strong> desired tapper position and orientation<br />

were rounded to the nearest 15 mm in position and nearest 10 degrees in orientation<br />

(relative to the global frame).


90 CHAPTER 7. POSITIONING EXPERIMENTS<br />

x<br />

β − υ<br />

υ<br />

y<br />

(i,j) striker offset<br />

φ<br />

P<br />

x<br />

j<br />

nominal contact frame<br />

φ<br />

0<br />

y<br />

P<br />

i<br />

β<br />

Figure 7.16: Notation for high precision positioning analysis. <strong>The</strong> striker nominally contacts<br />

the object at an angle of incidence of 0 relative to the contact normal and produces<br />

an impulse at an angle of 0 . If the striker is offset by a displacement (x; y) relative to<br />

the striker COM at the nominal contact position, the contact frame is rotates, the angle of<br />

incidence changes, and the angle of the resultant impulse changes.<br />

7.3.2 Sensitivity Analysis<br />

This subsection examines the sensitivity of a single tap to variations in the striker parameters<br />

by evaluating the jacobian for a nominal tap.<br />

Figure 7.16 shows the effect of displacing the striker with respect to the nominal contact<br />

point and describes the notation used in this subsection. <strong>The</strong> contact frame is rotated<br />

by an angle:<br />

j , i tan <br />

(7.4)<br />

R + r<br />

where R is the radius of the object and r the radius of the striker tip. One important quantity<br />

is the change of the direction of the impulse relative to its nominal direction, given<br />

by:<br />

a = + , 0 (7.5)<br />

A tap will produce some distance x f in the direction a relative to the nominal direction<br />

and some rotation f . As shown in Figure 7.17, these parameters are related to the final<br />

configuration of the object (x; y; ) by:<br />

x = x f cos a (7.6)<br />

y = x f sin a (7.7)<br />

= f (7.8)<br />

<strong>The</strong> sensitivity of this system to striker variations is characterized (to first order) by


7.3. HIGH PRECISION POSITIONING 91<br />

y<br />

(x,y)<br />

x<br />

a<br />

x f<br />

Figure 7.17: Final position in terms of distance x f and angle a relative to the nominal<br />

translation direction.<br />

the jacobian:<br />

2<br />

4 @x<br />

@y<br />

@<br />

3<br />

5 =<br />

2 6 4<br />

@x<br />

@i<br />

@y<br />

@i<br />

@<br />

@i<br />

@x<br />

@j<br />

@y<br />

@j<br />

@<br />

@j<br />

@x<br />

@<br />

@y<br />

@<br />

@<br />

@<br />

@x<br />

@v s 0<br />

@y<br />

@v s 0<br />

@<br />

@v s 0<br />

3<br />

75 2 64 @i<br />

@j<br />

@<br />

@v s0<br />

3<br />

75 (7.9)<br />

where the striker displacement is (i; j), itsangleofincidence, and its velocity v s0 .<br />

Derivation of the jacobian<br />

<strong>The</strong> jacobian can be decomposed into several parts, the first of which relates the final position<br />

(x; y; ) to the displacement (x f ; f ) and its direction a, based on the derivatives of<br />

Equations 7.6–7.8 evaluated at a =0:<br />

2 3 2 3<br />

4 @x<br />

@y 5 =<br />

@<br />

4<br />

1 0 0<br />

<br />

0 0 5 2 3<br />

4 @x f<br />

x 180 f @ f<br />

5 (7.10)<br />

0 1 0 @a<br />

where @a is assumed to be in degrees.<br />

<strong>The</strong> change in displacement and its direction are related to changes in the striker parameters<br />

by:<br />

2<br />

4 @x f<br />

@ f<br />

@a<br />

3<br />

5 =<br />

2 6 4<br />

@x f<br />

@i<br />

@ f<br />

@i<br />

@a<br />

@i<br />

@x f<br />

@j<br />

@ f<br />

@j<br />

@a<br />

@j<br />

@x f<br />

@<br />

@ f<br />

@<br />

@a<br />

@<br />

@x f<br />

@v s 0<br />

@ f<br />

@v s 0<br />

@a<br />

@v s 0<br />

3<br />

75 2 64 @i<br />

@j<br />

@<br />

@v s0<br />

<strong>The</strong> first two rows of this jacobian can be broken down into four components:<br />

<br />

@xf<br />

"<br />

@x f @x f<br />

@v<br />

= 0 @! 0<br />

#"<br />

@v 0 @v 0<br />

@P x @P y<br />

#"<br />

@P x @P x<br />

@S 0 @C 0<br />

#"<br />

@S 0 @S 0 @S 0<br />

@i @j @<br />

@ f<br />

@ f<br />

@v 0<br />

@ f<br />

@! 0<br />

@! 0<br />

@P x<br />

@! 0<br />

@P y<br />

@P y<br />

@S 0<br />

@P y<br />

@C 0<br />

@C 0<br />

@i<br />

@C 0<br />

@j<br />

3<br />

75 (7.11)<br />

@C 0<br />

@<br />

@S 0<br />

@v s 0<br />

@C 0<br />

@v s 0<br />

# 2 64 @i<br />

@j<br />

@<br />

(7.12)<br />

<strong>The</strong> first component relates the change in initial velocities to the change in displacement;<br />

these are the derivatives of the displacement functions of Chapter 2. <strong>The</strong>re appears to be<br />

no analytic expression for these derivatives, so this jacobian is numerically estimated. <strong>The</strong><br />

@v s0<br />

3<br />

75


92 CHAPTER 7. POSITIONING EXPERIMENTS<br />

other components can be calculated from the following formulas:<br />

qP 2 x + P 2 y<br />

v 0 =<br />

M<br />

! 0 = P xR<br />

I<br />

(7.13)<br />

(7.14)<br />

P x = S 0<br />

B 1<br />

(7.15)<br />

P y = (1 + e)C 0<br />

B 2<br />

(7.16)<br />

S 0 = (v s0 , k x ) sin( , )<br />

v s0<br />

(7.17)<br />

C 0 = (v s0 , k x ) cos( , )<br />

v s0<br />

(7.18)<br />

Since the second spring loaded tapping device is used for tapping an aluminum disk, Equations<br />

7.18 and 7.18 assume a generalized central impact and a striker of infinite angular<br />

inertia. <strong>The</strong> constant k in these equations models the deceleration of the striker due to<br />

friction as it travels towards the object.<br />

<strong>The</strong> last row of the jacobian in Equation 7.11 is obtained by differentiating Equation 7.5.<br />

For the assumed mode of impact, the angle of the momentum is related to the angle of<br />

incidence , (see Figure 7.16) by:<br />

<strong>The</strong> resulting derivatives are:<br />

@a<br />

@ =<br />

@a<br />

@x = , tan <br />

R + r<br />

@a<br />

@y = 1<br />

R + r<br />

<br />

x<br />

(R + r) cos 2 +<br />

2<br />

tan =<br />

B 2<br />

(1 + e)B 1<br />

tan( , ) (7.19)<br />

5 (7.20)<br />

cos<br />

B 2 2<br />

( , )<br />

3<br />

5 (7.21)<br />

cos 2 B 2<br />

( , )<br />

<br />

1<br />

1<br />

4 1 ,<br />

B 2<br />

sin 2 (1+e)B 1<br />

( , ) +<br />

2<br />

(1+e)B1<br />

1<br />

4 1 ,<br />

B 2<br />

sin 2 (1+e)B 1<br />

( , ) + (1+e)B1<br />

1+ x<br />

R + r<br />

1<br />

cos 2 <br />

B 2<br />

(1+e)B 1<br />

sin 2 ( , ) + (1+e)B1<br />

B 2<br />

cos 2 ( , )<br />

(7.22)<br />

3<br />

Evaluation of the jacobian<br />

2<br />

4 @x<br />

@y<br />

@<br />

3 2<br />

5 = 4<br />

,0:0426 0:26 0:243 0:0764<br />

,9:39 10 ,6 ,0:000052 0:352 0<br />

,0:00434 0:0507 0:0473 0:0115<br />

3<br />

5 2 64 @i<br />

@j<br />

@<br />

@v s0<br />

<strong>The</strong> jacobian of Equation 7.9 can now be assembled. For a 20 mm, 6 degree tap, we have:<br />

75<br />

(7.23)<br />

3


7.3. HIGH PRECISION POSITIONING 93<br />

10.<br />

6.7<br />

thf<br />

3.4<br />

0<br />

5<br />

10<br />

20<br />

15<br />

xf<br />

25<br />

30<br />

1<br />

0.8<br />

0.6 ||J||<br />

0.4<br />

0.2<br />

0<br />

35<br />

Figure 7.18: Norm of the final configuration jacobian as a function of displacement.<br />

2<br />

andfora10mm,3degreetap:<br />

4 @x<br />

@y<br />

@<br />

3 2<br />

5 = 4<br />

,0:0689 0:135 0:126 0:0562<br />

,4:7 10 ,6 ,0:000026 0:176 0<br />

,0:0088 0:0251 0:0234 0:00803<br />

3<br />

5 2 64 @i<br />

@j<br />

@<br />

@v s0<br />

3<br />

75 (7.24)<br />

Multiplying these jacobians by the striker error vector [7.5 7.5 5 0] T (which corresponds<br />

to the 15 mm, 10 degree rounding done in the experiment) yields [2.85 1.76 0.584] T for<br />

the 20 mm, 6 degree tap and [1.13 0.879 0.239] T for the 10 mm, 3 degree tap. For these taps<br />

(ignoring error in the striker velocity), the maximum error in [@x @y @] T that results from<br />

the maximum error in [@i @j @] T is smaller in each dimension. This offers some analytic<br />

support for the conjecture that tapping can position more precisely than the manipulator.<br />

One way to characterize the jacobian matrices is to find its norm. <strong>The</strong> norm of the 20<br />

mm, 6 degree jacobian is 0.468, that of the 10 mm, 3 degree jacobian is 0.246. Figure 7.18<br />

shows a plot of the jacobian norms for a range of displacements. From this graph, it appears<br />

that smaller taps have smaller norms (and are thus less sensitive to striker parameter<br />

errors) and that sufficiently large taps will have norms greater than 1. Intuitively, the latter<br />

observation is due to the fact that small angular errors will result in large errors in final<br />

position for large taps.<br />

7.3.3 Discussion<br />

For positioning tasks, we can generally express the linearized relationship between manipulator<br />

positioning error @u and change in object configuration @x by a matrix A:<br />

@x = A@u (7.25)<br />

For pick and place manipulation, the A matrix is the identity, so errors in the object position<br />

are the same magnitude as error in the manipulator. With tapping, we have seen that the


94 CHAPTER 7. POSITIONING EXPERIMENTS<br />

A matrix can have a norm less than one. Under this condition, @x will always be smaller<br />

than @u.<br />

In the preceding analysis, there is an implicit weighting of the different dimensions of<br />

the vectors @u and @x. In particular, one degree, one millimeter, and one millimeter per<br />

second are taken to be equivalent. <strong>The</strong> previous subsection showed that for the apparatus<br />

of this chapter’s experiments, there is a broad range of taps which have norm less than<br />

one. <strong>The</strong>se taps result in an error in object configuration that is smaller than the error in<br />

the manipulator configuration.<br />

Note that this analysis pertains to a single tap, which is limited in the set of configurations<br />

it can reach. This analysis is applicable to a proof of the stability of tapping which<br />

must address multiple-tap planning and control strategies. It is not difficult to define a<br />

metric that measures distance to the goal and to show that a planning strategy monotonically<br />

reduces this distance. <strong>The</strong> sensitivity analysis for a single tap will determine limits<br />

on acceptable manipulator error so that the system is still guaranteed to make progress<br />

towards the goal and thus maintain stability. <strong>The</strong> difficulties in such a proof come from<br />

careful definition of the distance metric and consideration of the acceptable positioning<br />

accuracy and the minimum and maximum possible taps.


Chapter 8<br />

Vibratory <strong>Manipulation</strong><br />

Vibratory manipulation is a variant of impulsive manipulation. I define vibratory manipulation<br />

to be any manipulation process involving repeated impacts due to a striker which<br />

follows some regular or servoed periodic motion, generally at a high frequency and low<br />

amplitude. Whereas impulsive manipulation primarily considers the mechanics of a single<br />

tap or a sequence of single taps, vibratory manipulation is concerned with object behavior<br />

due to repeated impacts — a more systems-oriented view. My primary interest is in stable<br />

periodic behaviors of an object that result from such excitation.<br />

This chapter considers vibratory manipulation in the context of a part that slides on<br />

a support surface in between impacts. <strong>The</strong>re are many different scenarios for this sort of<br />

manipulation that we can consider; these include:<br />

Open-loop control of a pusher with superimposed vibration. A pusher with a vibrating tip<br />

follows a predetermined trajectory while a striker tip executes some periodic motion.<br />

<strong>The</strong> actual object excitation is a high frequency, low amplitude sequence of collisions<br />

between the striker and the object.<br />

Feedback control of a pusher with superimposed vibration. Same as above except that the<br />

gross motion of the pusher is determined by some feedback control law.<br />

Feedback control of impulses. <strong>The</strong> object is sensed before every impact, and a control<br />

law determines an appropriate impulse to drive the object towards some goal configuration.<br />

Open-loop parts transfer. A sequence of impacts is planned to move an object from start<br />

to goal assuming ideal behavior of the slider and perfect knowledge of all relevant<br />

parameters. <strong>The</strong> shape of the striker and the details of the striking motion can be<br />

designed so that small errors give rise to restoring impulses.<br />

<strong>The</strong>re are two types of tapping that result under this vibratory manipulation: intermittent<br />

tapping, where the object comes to rest in between taps, and continuous tapping, where<br />

the object is always tapped again before it comes to rest.<br />

This chapter first develops the idea of vibratory manipulation in one dimension. Vibratory<br />

manipulation in one dimension that results in continuous tapping is closely related<br />

95


96 CHAPTER 8. VIBRATORY MANIPULATION<br />

to bouncing a ball on a sinusoidally vibrating table and to robotic juggling. Results from<br />

work in these areas is reviewed and adapted to vibratory manipulation in order to determine<br />

the conditions for stable periodic motion.<br />

<strong>The</strong> two dimensional case of vibratory manipulation is more complicated; it is related<br />

to the control strategies used for positioning experiments in Chapter 7. <strong>The</strong>se strategies<br />

specify a striker behavior for repeated intermittent tapping; however they involve replanning<br />

at each step instead of a regular or servoed behavior.<br />

<strong>The</strong> basis of vibratory manipulation in two dimensions is in the limiting cases of impulsive<br />

manipulation. <strong>The</strong>se limiting cases examine the object behavior in the ideal case<br />

when the number of taps approaches infinity and the size of each tap approaches zero.<br />

This chapter presents a detailed analysis of the limiting cases for axisymmetric objects and<br />

examines the relationship between the limiting cases and pushing, concluding with a few<br />

examples.<br />

8.1 Vibratory manipulation in one dimension<br />

Suppose a given task is to move an object a distance d. A single tap could be used to give<br />

the object enough velocity so that it will slide and come to rest after traveling a distance<br />

d. Alternatively, several taps could be used to cover the distance. We could let the object<br />

come to rest after each tap, or we could deliver the next tap before the object stops.<br />

This gives rise to a number of different ways in which the relationship between the<br />

impulse of a tap and the period between taps may be defined. Regardless of the definition<br />

used, I take the limiting case to be the limiting behavior as the impulse delivered by a<br />

single tap approaches zero (and the number of taps approaches infinity). <strong>The</strong> first two<br />

subsections describe the limiting cases of intermittent tapping and continuous tapping.<br />

<strong>The</strong>se limiting cases are related to the ideal case of vibratory manipulation — they<br />

presume a sequence of identical taps without error and describe how the mechanics change<br />

as the size of a tap decreases. <strong>The</strong> last two subsections address the more applied problem<br />

of how errors and striker motion interact with the mechanics to affect the resulting object<br />

behavior.<br />

8.1.1 Intermittent tapping<br />

Suppose we are able to move the object a distance d with a single tap. Consider the tap<br />

with 1 of the impulse, producing 1 of the initial velocity. <strong>The</strong> deceleration due to friction<br />

n n<br />

will be the same, so the distance covered by this tap will be d . It will require n such taps<br />

n<br />

(letting the object come to rest after each one) to move the object a distance d. See Figure 8.1<br />

for an illustration.<br />

<strong>The</strong> initial velocity required to move the object a distance d is n<br />

v 0 =r2gd<br />

n<br />

(8.1)


8.1. VIBRATORY MANIPULATION IN ONE DIMENSION 97<br />

1 tap<br />

velocity<br />

2 taps<br />

time<br />

n taps<br />

Figure 8.1: One dimensional intermittent tapping. We want to move the object a given<br />

distance using one or more taps, letting the object come to rest after each tap. As the number<br />

of taps increases, the average velocity approaches zero and the total time approaches<br />

infinity.<br />

and the time required for the object to come to rest is<br />

t tap =s<br />

2d<br />

gn<br />

(8.2)<br />

Note that the total amount of time for n taps is<br />

t =s2dn<br />

g<br />

(8.3)<br />

and that the average velocity is<br />

v =r<br />

gd<br />

(8.4)<br />

2n<br />

Under this definition of the limiting case, as the number of taps n approaches infinity and<br />

the initial velocity of each tap approaches zero, the total time required to move the object a<br />

distance d approaches infinity, and the average velocity approaches zero.<br />

8.1.2 Continuous tapping<br />

Although the limiting case for intermittent tapping has useful properties, it would also be<br />

useful to have a limiting case which converges to a nonzero average velocity. This can be<br />

done by tapping the object again before it has come to rest. This limiting case appears to<br />

approach a quasistatic equilibrium in the sense that a constant velocity implies that the<br />

frictional forces are balanced by the “driving forces”.<br />

<strong>The</strong>re are a number of ways in which the relationship between the impulse of a tap and<br />

the period between taps can be defined. <strong>The</strong> definition used will affect how the velocity<br />

converges. Figure 8.2 shows a few examples.


98 CHAPTER 8. VIBRATORY MANIPULATION<br />

v<br />

(a) (b) (c)<br />

2 taps<br />

t<br />

3 taps<br />

n taps<br />

Figure 8.2: Velocity profiles for possible definitions of the limiting case of continuous tapping<br />

for one dimension. In all cases, the object is tapped before it has come to rest. In Figure<br />

(a), the average velocity remains constant in the limit, whereas in Figures (b) and (c),<br />

the average velocity approaches a limiting value from above and from below respectively.<br />

This approach to the limiting case for continuous tapping departs somewhat from the<br />

stated task of moving an object a distance d, which implicitly assumes that the object starts<br />

and finishes at rest. Although it is possible to address the starting and stopping issues<br />

in this definition, I am primarily interested in characterizing the steady state (periodic)<br />

behaviors that can be achieve in the limit.<br />

8.1.3 Vibratory manipulation and ball bouncing<br />

Suppose that the striker moves with some periodic vibration superimposed on a constant<br />

velocity and that this constant velocity is sufficiently high that the object does not come<br />

to rest in between impacts. When viewed in a frame moving at that constant velocity (see<br />

Figure 8.3), this system is equivalent to a ball bouncing on a vibrating table — between<br />

impacts, the object is subject to a constant acceleration produced by friction, whereas the<br />

ball is subject to constant acceleration produced by gravity.<br />

<strong>The</strong> problem of a ball bouncing on a sinusoidally vibrating table has been studied<br />

both in theory and experimentally. By applying results from the bouncing ball problem<br />

to one dimensional continuous tapping, the stability of object behaviors can be related to<br />

parameters of a sinusoidal striker vibration.<br />

Holmes [31] studied the bouncing ball problem under the assumption that the vibration<br />

of the table is small compared to the displacement of the ball. Under this assumption,<br />

the velocity of the ball before an impact is the same as its velocity after the previous impact.<br />

Bapat et al. [7] examined and compared the approximated mechanics and the exact<br />

mechanics. In this chapter, I adopt most conventions and notation of Bapat et al.<br />

<strong>The</strong> motion of the table is given by:<br />

A sin !t (8.5)<br />

and its velocity by:<br />

A! cos !t (8.6)


8.1. VIBRATORY MANIPULATION IN ONE DIMENSION 99<br />

Figure 8.3: An illustration of regular periodic bouncing. <strong>The</strong> table moves with a sinusoidal<br />

vibration, and the object follows a parabolic trajectory while in flight. This same illustration<br />

applies to tapping an object using a striker with a sinusoidal vibration superimposed on a<br />

constant velocity (when viewed in the frame of that constant velocity).<br />

where ! here is 2 times the frequency of the table oscillation. <strong>The</strong> velocity of the object<br />

just before or after a collision is denoted by V .<br />

By reasoning about the return map generated by the approximated mechanics ( mapping<br />

the velocity of the ball (just before an impact) and the phase of the impact (relative to<br />

the table oscillation) from one impact to the next ), Holmes showed that for a coefficient of<br />

restitution less than 1, the trajectory of the ball will be bounded. Given the frequency and<br />

amplitude of the oscillation, the velocity of the ball just before an impact (or just after the<br />

previous impact) will eventually satisfy<br />

(1 + e)<br />

jV j < A! (8.7)<br />

(1 , e)<br />

Fixed points of the return map correspond to different modes of stable periodic bouncing.<br />

<strong>The</strong> simplest mode of object behavior is what Holmes called period one motions and Bapat<br />

et al. called (n,1) motions (a collision every n cycles of the table with a collision pattern<br />

of length 1). For a given n and !, stability analysis of the fixed points showed that in order<br />

to maintain stable periodic bouncing the amplitude of the table must satisfy:<br />

ng (1 , e)<br />

! 2 (1 + e)


100 CHAPTER 8. VIBRATORY MANIPULATION<br />

10<br />

8<br />

Amplitude (mm)<br />

6<br />

4<br />

2<br />

0<br />

2 4 6 8 10 12 14<br />

frequency (Hz)<br />

Figure 8.4: Relationship between frequency and amplitude of a sinusoidally vibrating<br />

striker for stable periodic bouncing (period one, n =1bouncing). Here we have assumed<br />

e =0:8 and =0:25. Increasing the amplitude beyond the range shown here will result in<br />

period doubling and will eventually lead to chaotic behavior. Using an amplitude below<br />

this range will likely result in the object coming to rest against the striker and being pushed<br />

along.<br />

An example<br />

Suppose we take n =1, the coefficient of restitution e =0:8, and the coefficient of friction<br />

=0:25. From equation 8.8, we find that the striker amplitude must satisfy<br />

0:856<br />

! 2 v O ) in order to ensure that the object does not come to rest before it<br />

is struck again. In fact, some tolerance will be required since the mechanics used here are<br />

approximate.<br />

Under continuous motion, an object being “pushed” by a sinusoidally vibrating striker<br />

moving at a constant average velocity v might appear as shown in Figure 8.3 (in the frame<br />

of reference moving at velocity v).


8.2. VIBRATORY MANIPULATION IN TWO DIMENSIONS 101<br />

Figure 8.5: Illustration of an object bouncing off a sinusoidally vibrating striker under intermittent<br />

tapping. This sinusoidal vibration is superimposed on a constant velocity v,<br />

and we view the system from a frame moving at that velocity. Here, the object follows a<br />

parabolic trajectory until it comes to rest relative to the surface; in the moving frame, it<br />

appears to be “falling” with a velocity ,v.<br />

Intermittent bouncing<br />

If the object does come to rest in between impacts, then in the frame moving at the average<br />

velocity of the striker, the object will appear to stop “falling” along a parabolic path<br />

and move at a constant velocity ,v, as illustrated in Figure 8.5. Although not nearly as<br />

amenable to analysis as continuous bouncing, we should still be able to achieve stable periodic<br />

impacts with intermittent motion of the object.<br />

8.1.4 Robotic juggling<br />

Schaal and Atkeson [56] have implemented paddle jugging with a “trampoline-like” racket<br />

and a ping-pong ball. <strong>The</strong>y observed that negative acceleration of the paddle trajectory<br />

at impact seems to provide stability. <strong>The</strong>y calculated the basin of attraction for periodic<br />

juggling to be 0.257 the area of the viable space of initial conditions. However, for a number<br />

of reasons (including mechanical variability and air resistance) the basin of attraction was<br />

significantly larger. In fact, they were unable to avoid getting a periodic (mostly period<br />

one) juggling pattern.<br />

Bühler and Koditschek [16] demonstrated a system that juggles a puck on an inclined<br />

plane using a bat. In order to achieve stable juggling, they developed a “mirror” control<br />

law which servos the bat to nonlinearly reflect the position of the puck. Using this sort of<br />

simple control law for vibratory manipulation is appealing but sensing the object would<br />

be difficult for high frequency, low amplitude motions.<br />

8.2 Vibratory manipulation in two dimensions<br />

<strong>The</strong> two dimensional case is much more complicated than the one dimensional case because<br />

of the complexity of the mechanics (which is in turn due to the strong coupling between<br />

translational and rotational motion). This section addresses vibratory manipulation<br />

for axisymmetric objects under the idealized assumption that the object is manipulated by<br />

a sequence of identical taps with no error. This section does not address how errors and


102 CHAPTER 8. VIBRATORY MANIPULATION<br />

x f /2<br />

Net COR<br />

θ f<br />

x f<br />

2 tan (θ<br />

f/2)<br />

x f<br />

θ f<br />

Figure 8.6: Construction for finding the net COR for a tap where the object moves a distance<br />

x f and turns an angle f .<br />

striker motion interact with the mechanics to affect stable object behavior.<br />

Since all impacts are assumed to be identical, this section examines the constraints on<br />

object motion in the limit, i.e. as the impulse of a single tap approaches zero. This section<br />

addresses the limiting case of the axisymmetric mechanics for intermittent tapping and for<br />

continuous tapping.<br />

<strong>The</strong> two limiting cases can be characterized by the locus of centers of rotation (CORs)<br />

that they can produce. As illustrated in Figure 8.6, any planar displacement can be represented<br />

as a pure rotation about some point in the plane. <strong>The</strong> locus of CORscanbeused<br />

to compare two different modes of manipulation; it is also sufficient information to create<br />

plans for that mode of manipulation<br />

This section assumes that the object boundary is circular and develops the limiting<br />

cases for positive rotations; the constraints on object motion for negative rotations will be<br />

symmetric.<br />

8.2.1 Limiting case of intermittent tapping<br />

For intermittent tapping, the object undergoes a sequence of identical impacts, always coming<br />

to rest before the next impact. In state space, the object repeatedly follows a path starting<br />

at the initial velocities and ending at the origin (zero velocities). I take the limiting<br />

case to be the resultant behavior as the impulse approaches zero but the ratio of the initial<br />

velocities remains the same.<br />

Net COR in the limit<br />

Given the distance traveled and the angle turned, the net COR for the motion can be determined<br />

with a geometric construction as illustrated in Figure 8.6. <strong>The</strong> location of the net<br />

COR is:<br />

~x f<br />

2 + ^z ~x f<br />

2 tan f 2<br />

(8.12)


8.2. VIBRATORY MANIPULATION IN TWO DIMENSIONS 103<br />

Since the impulse of a tap approaches zero by scaling the initial velocities, the trajectory<br />

scaling of Section 2.1.6 (scaling the velocities by k) is used to take the limit:<br />

k 2 ~x f<br />

lim<br />

k!0 2<br />

+ ^z k2 ~x f<br />

2 tan k2 f<br />

2<br />

= lim<br />

k!0<br />

^z k 2 ~x f<br />

2 tan k2 f<br />

2<br />

(8.13)<br />

Applying l’Hôpital’s rule (differentiating both numerator and denominator with respect to<br />

k), yields:<br />

^z ~x f<br />

lim<br />

= ^z ~x f<br />

(8.14)<br />

k!0<br />

f sec 2 k2 f f<br />

2<br />

Thus, as the impulse delivered by a tap approaches zero, the distance from the center of<br />

mass (COM) to the net COR approaches xf<br />

f<br />

.<br />

Locus of CORs<br />

This limiting case can be characterized by the locus of possible CORs. This represents all<br />

possible motions of the object for a given initial velocity direction.<br />

<strong>The</strong> trajectory scaling of Section 2.1.6 shows that each value of !0<br />

v 0<br />

corresponds to a<br />

single value of xf<br />

f<br />

, regardless of the magnitude of the initial velocities or the resulting<br />

displacements. It was also shown in Section 2.1.6 that xf<br />

f<br />

is a continuous monotonic decreasing<br />

function of !0<br />

v 0<br />

. <strong>The</strong>refore the minimum and maximum values of !0<br />

v 0<br />

that can be<br />

achieved via an impact correspond to the maximum and minimum values of xf<br />

f<br />

for the<br />

possible resulting displacements. <strong>The</strong>se values of xf<br />

f<br />

in turn determine the closest and<br />

furthest CORs possible for this limiting case.<br />

<strong>The</strong> range of possible !0<br />

v 0<br />

that can be produced by impact was introduced in Section<br />

3.2.2 as impact cones (for the case where the set of possible velocity ratios does not<br />

change with velocity direction). <strong>The</strong> impact cone for a circular object can be expressed as:<br />

p ! 0<br />

rRM<br />

v 0 I 1+ 2 r<br />

(8.15)<br />

<strong>The</strong> minimum value of !0<br />

v 0<br />

is 0, corresponding to pure translational velocity (which results<br />

in a pure translation of the object: xf<br />

f<br />

= 1); the maximum value is given by the equality<br />

condition of Equation 8.15. Hence, the range of CORs for the limiting case of intermittent<br />

tapping will be from infinity to the closest COR. Letc min = xf<br />

for the maximum value of<br />

! 0<br />

v 0<br />

that satisfies Equation 8.15. Applying this range of xf<br />

f<br />

to Equation 8.14 determines the<br />

locus of all possible CORs:<br />

fc^z ^v 0 j c 2 [c min ; 1]g (8.16)<br />

This locus consists of the entire plane except for an open disk of radius c min centered at the<br />

object COM.<br />

8.2.2 Limiting case of continuous tapping<br />

For continuous tapping, the object undergoes a sequence of regularly spaced impacts which<br />

always strike the object before it comes to rest. In state space, the object repeatedly follows<br />

f


104 CHAPTER 8. VIBRATORY MANIPULATION<br />

w<br />

(a) (b) (c)<br />

ω<br />

stable<br />

Negative Impact Cone<br />

ω stable<br />

eigendirection<br />

eigendirection<br />

Negative<br />

Impact Cone<br />

v<br />

v<br />

stable<br />

eigendirection<br />

v<br />

Figure 8.7: Illustrations of the state space for a sliding rotating axisymmetric object; the<br />

stable eigendirection is always the dashed-dotted line. Figure (a) shows a typical vector<br />

field, which is a function of only ! . Figure (b) illustrates the case where the initial slope of<br />

v<br />

the path in state space aligns with the negative impact cone above the stable eigendirection.<br />

In this case, the trajectory curves in to the cone. Figure (c) illustrates the case where this<br />

alignment occurs below the eigendirection and the trajectory curves away from the cone.<br />

a path starting at its initial velocities and continues until the object is tapped again; this<br />

“kicks” the state back to the original initial velocities. I take the limiting case to be the<br />

resulting behavior of the object as the amount of time before the next impact approaches<br />

zero.<br />

<strong>The</strong> same expression (Equation 8.12) for the location of the net COR is used for this<br />

limiting case; however, the limit is now taken with respect to time between impacts t f .As<br />

t f decreases, the displacements ~x f and f both approach zero:<br />

lim<br />

t f !0<br />

~x f<br />

2 + ^z ~x f<br />

2 tan f 2<br />

= lim<br />

tf!0<br />

^z ~x f<br />

2 tan f 2<br />

(8.17)<br />

Assume for the moment that this limit exists. Applying l’Hôpital’s rule (differentiating<br />

numerator and denominator with respect to t f ), yields:<br />

lim<br />

t f!0<br />

^z d<br />

dt f<br />

sec 2 f<br />

~x f<br />

d<br />

dt f<br />

f<br />

= lim<br />

t f!0<br />

^z d<br />

dt f<br />

d<br />

dt f<br />

R<br />

t f<br />

R t f<br />

~vdt 0<br />

t f!0<br />

0 !dt = lim<br />

^z ~v(t f )<br />

!(t f )<br />

= ^z ~v 0<br />

! 0<br />

(8.18)<br />

Thus, the COR in the limit is simply the initial (instantaneous) COR.<br />

<strong>The</strong> remainder of this subsection examines the conditions under which this limit exists,<br />

and this will determine the locus of CORs possible under this limiting case.<br />

Impact constraint for moving objects<br />

Appendix A demonstrates that it is possible to generate any impact within a friction cone<br />

in impulse space by striking an object at rest. For a given impulse in the friction cone,<br />

we can compute the striker velocity and contact point required to achieve that change in<br />

velocity. Not surprisingly, so long as the same relative velocities between the striker and<br />

the object are recreated, the same change in velocity can be achieved. Thus the impact cone


8.2. VIBRATORY MANIPULATION IN TWO DIMENSIONS 105<br />

!<br />

v<br />

Disk<br />

COR<br />

!<br />

v<br />

Ring<br />

COR<br />

Intermittent Tapping LC 0.080 0.208<br />

Continuous Tapping LC/Pushing 16.395 0.061 8.951 0.112<br />

Impact Cone 9.704 0.103 4.851 0.206<br />

Eigendirection 30.798 0.032 20.000 0.050<br />

Radius 0.050 0.050<br />

Table 8.1: Motion constraints for a disk and a ring for both limiting cases (LCs). <strong>The</strong> distance<br />

to the COR given for the impact cone, limiting cases, and pushing is the distance to<br />

the closest COR. <strong>The</strong> distance to the COR corresponding to the stable eigendirection and<br />

the radius of the object are given for comparison. Quasistatic pushing has the same motion<br />

constraints as the limiting case of continuous tapping. <strong>The</strong>se results were numerically<br />

generated, and assume that = r =0:25.<br />

for a circular object (Equation 8.15) can be restated in terms of change in velocities (for<br />

positive and negative changes in rotational velocity):<br />

j!j<br />

v rRM<br />

I p 1+ 2 r<br />

(8.19)<br />

<strong>The</strong> result is that we can move the impact cone to any point in state space.<br />

This section will primarily make use of the negative impact cone, whichisthesetof<br />

states that can reach the current state with a single impact. Since the impact cone is invariant<br />

with respect to its position in state space, the negative impact cone is simply the<br />

reflection of the impact cone through the current state.<br />

Existence of the limit & locus of CORs<br />

<strong>The</strong> limit taken in Equations 8.17 and 8.18 exists if and only if the path of the object in state<br />

space passes through the negative impact cone (at the initial velocities) for a nonzero period<br />

of time. In order to determine which initial conditions satisfy this property, the properties<br />

of paths in state space must be examined. <strong>The</strong> eigendirections of Goyal et al. [27] serve to<br />

guide this analysis.<br />

<strong>The</strong>re are a number of properties of paths in state space for a sliding axisymmetric<br />

object. <strong>The</strong>y are stated here in terms of the vector field (_v; _!) =(,F=M;,T=I) over the<br />

state space.<br />

as ! v increases, the orientation of (_v; _!) monotonically increases from to 3 2 .<br />

(_v; _!) points towards the origin only at values of ! corresponding to eigendirections;<br />

v<br />

most objects have a single stable eigendirection that is neither pure translation nor<br />

rotation.<br />

at all points below the stable eigendirection, (_v; _!) points above the origin, drawing<br />

the system towards the stable eigendirection.


106 CHAPTER 8. VIBRATORY MANIPULATION<br />

at all points above the stable eigendirection, (_v; _!) points below the origin, drawing<br />

the system towards the stable eigendirection.<br />

<strong>The</strong>se properties are illustrated in Figure 8.7a.<br />

If the negative impact cone is placed on the line corresponding to ! = 0(a COR at<br />

v<br />

infinity), the path in state space will lie entirely inside the cone. As the value of ! is increased,<br />

the vector (_v; _!) will rotate counterclockwise. <strong>The</strong> path of the object in state space<br />

v<br />

will continue to pass through the cone until this vector aligns with the bottom edge of the<br />

cone. At that point:<br />

F ( ! v )<br />

T ( ! v ) = R<br />

(8.20)<br />

q1 + 1 2 r<br />

Let 1=c min be the value of !0<br />

v 0<br />

which satisfies this equation.<br />

If this alignment occurs above the stable eigendirection then the path will curve into<br />

thecone(seeFigure8.7b),sothelocusofpossibleCORsis<br />

fc^z ^v 0 j c 2 [c min ; 1]g (8.21)<br />

If this alignment occurs below the stable eigendirection then the path will curve away from<br />

thecone(seeFigure8.7c),sothelocusofpossibleCORsis<br />

fc^z ^v 0 j c 2 (c min ; 1]g (8.22)<br />

<strong>The</strong>se loci cover the plane except for an open or closed disk (depending upon when alignmentoccurs)ofradiusc<br />

min centered at the COM of the object.<br />

8.3 Comparison with pushing<br />

For comparison to the limiting cases of impulsive manipulation, this section develops the<br />

constraints on the motion of a pushed object.<br />

<strong>The</strong> pushing force exerted on the object must lie within the friction cone at the contact<br />

point. This force and the resulting torque are resisted by frictional forces which are<br />

determined by the velocities of the object. Thus, the range of forces and torques that can be<br />

exerted on the object indirectly determine the range of velocities and therefore the range of<br />

CORs.<br />

Consider a circular axisymmetric object and presume, akin to the limiting cases of<br />

impulsive manipulation, that this object can be pushed at any point on the boundary in<br />

order to push the object in a single given direction. A force that produces no torque can<br />

always be exerted on the object; this produces pure translation, corresponding to a COR at<br />

infinity. As illustrated in Figure 8.8, the maximum torque that can be exerted for a given<br />

force is determined by the coefficient of friction between the pusher and the object, r .<br />

<strong>The</strong>netforceandtorque,F ( ! v ) and T ( ! ), are monotonic decreasing and increasing<br />

v<br />

functions of ! respectively. <strong>The</strong> maximum torque to force ratio then determines the maximum<br />

value of ! (the minimum value of v ) and therefore the closest possible COR.<br />

v<br />

v !


8.4. EXAMPLES FOR THE LIMITING CASES 107<br />

-1<br />

tan µ<br />

Figure 8.8: <strong>The</strong> maximum value of T for pushing a disk is achieved by producing the<br />

F<br />

greatest amount of torque for a given force. Friction cones are drawn here at the contact<br />

points which have the maximum lever arm for a frictional force in the ^x direction.<br />

<strong>The</strong> maximum torque that can be generated for a given force is due to a force on the<br />

edge of the friction cone. This torque is:<br />

<br />

T = j~r jF sin ( tan ,1 r<br />

r )=RFp (8.23)<br />

1+<br />

2<br />

r<br />

let 1=c min be the value of ! v<br />

that satisfies:<br />

F ( ! v )<br />

T ( ! v ) =<br />

R<br />

q1 + 1 2 r<br />

(8.24)<br />

<strong>The</strong> set of rotation centers given the velocity direction ^v is<br />

fc^z ^v j c 2 [c min ; 1]g (8.25)<br />

Note that this is the same condition for the minimum COR as in the limiting case of<br />

continuous tapping. (Compare Equations 8.24 and 8.20). With the possible exception of<br />

the rotation centers at c min^z ^v , pushing and the limiting case of continuous tapping can<br />

maintain the same range of CORs.<br />

8.4 Examples for the limiting cases<br />

Table 8.1 shows the results of analyzing the possible motions of a uniform disk and a ring<br />

under the two limiting cases. <strong>The</strong>se results are graphically illustrated in Figure 8.9. Note<br />

that the “alignment” for the limiting case of continuous tapping occurs below the stable<br />

eigendirection (i.e. the value of ! (slope in state space) for the limiting case is less than that<br />

v<br />

for the eigendirection).<br />

This table also includes the size of the impact cone (i.e the maximum value of !0<br />

v 0<br />

achievable by impact) which is given by<br />

!<br />

v <br />

R<br />

2q1 + 1<br />

2 r<br />

(8.26)


108 CHAPTER 8. VIBRATORY MANIPULATION<br />

20<br />

int.<br />

int.<br />

cont.<br />

continuous LC<br />

15<br />

10<br />

5<br />

cont.<br />

continuous LC<br />

intermittent LC<br />

cm<br />

intermittent LC<br />

Figure 8.9: Graphic illustration of the motion constraints on a disk and a ring due to tapping.<br />

For each limiting case, the object can move about CORs from infinity (pure translation)<br />

up to the closest COR, indicated by a dot; a radius is drawn to the path of maximum<br />

curvature that the object can follow. <strong>The</strong> shaded regions represent the locations the objects<br />

can reach; the set of locations for intermittent tapping is a subset of the locations for<br />

continuous tapping for these examples.<br />

and will is a function of the radius of the object. <strong>The</strong> smaller the radius, the smaller the<br />

radius of gyration, and the wider the impact cone will be.<br />

Note that for both examples, the range of CORs possible for the limiting cases is greater<br />

than that of the impact cone, i.e. there are states at which the limiting case can be sustained<br />

but which cannot be reached directly when starting the object from rest. Also note that<br />

the range of possible CORs for both limiting cases in both examples lies below the eigendirection;<br />

therefore the constraints on object motion for pushing and for the limiting case of<br />

continuous tapping differ with respect to motion about the CORsatc min (^z ^v).<br />

8.5 Discussion<br />

8.5.1 Limiting case of intermittent tapping<br />

<strong>The</strong> limiting case of intermittent tapping is not in itself useful because the average velocity<br />

approaches zero in the limit, but approximations to the limiting case have several useful<br />

properties.<br />

First of all, an approximation to this limiting case would be the simplest and perhaps<br />

most practical to implement since the object comes to rest in between taps. This makes it<br />

easier to incorporate sensory feedback while executing a series of taps and allows for more<br />

exact execution of planned sequence of taps. In contrast, the limiting case of continuous<br />

tapping is more dynamic, so it would be difficult to repeatedly strike the object at exactly<br />

the same state and “kick” it back precisely to the initial state.<br />

Intermittent tapping also provides some independence between planning a path and<br />

executing it because the exact same path that can be executed with a single tap can be<br />

executed by arbitrarily many taps. In fact, since error tends to grow with increased energy,


8.5. DISCUSSION 109<br />

ω<br />

stable eigendirection<br />

alignment<br />

value<br />

edge of impact cone<br />

Figure 8.10: When the initial impact cone lies below the eigendirection, we can reach points<br />

in the state space outside the this cone by repeated impacts. Since all points below the<br />

eigendirection asymptotically approach the eigendirection, ! increases beyond the impact<br />

v<br />

cone. By repeatedly striking the object, we can maintain the energy of the object and thus<br />

reach any point in the state space for which the limiting case of continuous tapping exists.<br />

<strong>The</strong> solid lines show the object sliding; the dotted lines correspond to impacts.<br />

v<br />

smaller taps yield smaller error, so it may be advantageous to use a sequence of smaller<br />

taps, possibly with feedback.<br />

With the range of possible CORs a given object, a nonholonomic motion planner could<br />

be adapted to plan complex motions which avoid obstacles, as has been done for pushing<br />

by Lynch and Mason [41]. One caveat is that motion about a COR may need to be followed<br />

arbitrarily closely in order to avoid obstacles.<br />

8.5.2 Continuous tapping limiting case<br />

In the examples presented in Section 8.4, the closest possible COR for the limiting case of<br />

continuous tapping was closer than the COR corresponding to the edge of the impact cone,<br />

i.e. the limiting case of continuous tapping can be sustained at states which cannot be<br />

directly reached from rest.<br />

When !0<br />

v 0<br />

is below the eigendirection, the trajectory will eventually leave the negative<br />

impact cone. If the object is tapped again before it leaves the negative impact cone, the<br />

value of !0<br />

v 0<br />

can be increased. Figure 8.10 illustrates this process repeated in order to reach<br />

states outside of the impact cone at the origin. If the object is tapped again after it leaves<br />

the negative impact cone, then the value of !0<br />

v 0<br />

will decrease, regardless of the tap.<br />

8.5.3 Tracking versus following the object<br />

One issue in the two dimensional limiting cases is whether to track or follow the object, i.e.<br />

whether to always strike the object at the same contact point relative to the world frame<br />

or relative to the object frame. <strong>The</strong> question is essentially whether to ignore rotation or to<br />

track rotation. <strong>The</strong> distinction is clearest for objects with circular boundaries because the<br />

boundary of a circle is same regardless of its orientation, so either choice may be taken. For


110 CHAPTER 8. VIBRATORY MANIPULATION<br />

noncircular object boundaries, only tracking the object is possible. (Following an object<br />

with a noncircular boundary would not generally lead to a stable periodic motion.) <strong>The</strong><br />

decision between following or tracking affects whether the COR of the motion remains fixed<br />

within the world frame or relative to the body.


Chapter 9<br />

Conclusions<br />

9.1 Contributions<br />

This thesis has demonstrated, both analytically and experimentally, how tapping can be<br />

used for manipulation. <strong>The</strong> main results are:<br />

Analysis of the mechanics of impact and friction to solve the planning problem for<br />

tapping.<br />

Experimental validation that the models can predict motion of real objects with reasonable<br />

accuracy.<br />

Development of planning methods (or feedback control strategies) to compensate for<br />

experimental errors.<br />

Experimental demonstration that tapping can be used to position an object to greater<br />

precision than the manipulator can position a tapping device.<br />

Development of devices to generate a single controlled repeatable impact and identification<br />

of design goals and issues for these devices.<br />

In the analysis portion, this thesis presents a solution to planning a single tap by solving<br />

the inverse sliding problem and the impact problem. For the inverse sliding problem,<br />

I have described methods to determine initial velocities for an object to slide a given displacement<br />

(translation and rotation) and have proven the existence and uniqueness of the<br />

solution for the axisymmetric case. For the impact problem, I have applied the mechanics<br />

of impact to determine how to strike an object to produce desired initial velocities and have<br />

also described a method to characterize the velocities that can be produced by an impact.<br />

Not all displacements can be reached in a single tap, so a sequence of taps may be<br />

required. For axisymmetric and nearly axisymmetric objects, I have described constructs<br />

and methods for planning multiple-tap plans and have shown that any goal configuration<br />

can be reached within two taps. <strong>The</strong>se constructs are also used to show that tapping is<br />

small time locally controllable for all axisymmetric and at least some nearly axisymmetric<br />

objects. Thus, tapping can be used to make these object follow any path arbitrarily closely.<br />

111


112 CHAPTER 9. CONCLUSIONS<br />

Tapping can also be used in a manner which I refer to as vibratory manipulation;<br />

here, an object interacts with a striker which follows some regular or servoed periodic<br />

motion (generally a high frequency low amplitude motion). I have studied this problem<br />

in one dimension, applying results from related work to determine conditions for stable<br />

periodic object motion, and I have laid the foundation for two dimensions by formulating<br />

the limiting cases of the mechanics for a single tap.<br />

<strong>The</strong> experiments are a major part of the contributions of this thesis. My experimental<br />

efforts began with an exploration of ways to produce a controlled impact. I have designed<br />

and used five different tapping devices in the course of this work; this thesis describes<br />

those devices and their performance and discusses the design issues I have found to be<br />

most important.<br />

<strong>The</strong> experimental setup did not fully conform to the idealized assumptions used in<br />

modelling and analysis, primarily with respect to the uniformity of support distributions.<br />

<strong>The</strong> key question in the first experiments was whether the impact and sliding models<br />

would still be adequate to predict object motion. <strong>The</strong>se experiments showed that the models<br />

can be modified to predict object motion with acceptable accuracy; this modification is<br />

the use of a scaling factor to increase the amount of torque acting upon the object. For the<br />

three experiments described in this thesis, one required no scaling factor; the other two required<br />

scaling factors of 2.17 and 3.64. <strong>The</strong> reason for this deviation is not completely clear.<br />

It is likely due in part to the nonuniformity of the support distribution and to the three dimensional<br />

extent of the object. However, it also seems plausible that frictional torque may<br />

be caused by a mechanism that differs from classical mechanics since Coulomb friction is<br />

an approximation to a complex nonlinear phenomena.<br />

In order to use tapping for positioning tasks, planning methods must take into account<br />

both errors due to modelling and errors due to variation in the impact and friction<br />

phenomena. I incorporated different error models into three different planning methods<br />

that essentially serve as feedback control laws. <strong>The</strong> modified mechanics models and these<br />

planning methods were used to demonstrate a positioning task via tapping. Finally, it was<br />

experimentally shown that tapping can be used to position an object to greater precision<br />

than the tapping device itself is positioned by rounding positions and orientations of the<br />

tapping device with respect to the global frame.<br />

9.2 Future work<br />

This thesis is a first step in the formal study and application of impulsive manipulation.<br />

I have shown that tapping, one form of impulsive manipulation, is a viable and useful<br />

method of manipulation; it possesses a unique combination of attributes, namely its speed,<br />

nonprehensile nature, and limited interaction between the manipulator and the object. This<br />

thesis covers a wide range of topics directed towards the analysis and experimental use of<br />

tapping, but there are many areas that deserve more attention.<br />

Friction modeling<br />

It is unclear whether the necessity of a torque scaling factor is due to some fundamental<br />

mechanism or whether it is simply due to deviations from the idealized models. This will


9.2. FUTURE WORK 113<br />

require many careful experiments with precise measurements.<br />

Stability analysis<br />

A formal proof of the stability of tapping under some planning method should be possible,<br />

building upon the sensitivity analysis of Chapter 7. This would result in a characterization<br />

of the bounded input bounded output stability of the system in the presence of errors.<br />

<strong>The</strong> planning methods described in this thesis have different distributions of the probability<br />

of reaching the goal as a function of the number of taps. It would be interesting to<br />

characterize the tradeoff between the average plan length and its variation.<br />

Adaptive control and active perception<br />

Although it is not possible in general to determine both material properties and tapping<br />

device calibration at the same time, with knowledge of one we can determine the other. It<br />

should be possible to formulate an adaptive controller to learn the properties of an object,<br />

essentially implementing a form of active perception.<br />

Vibratory manipulation<br />

Experimental demonstration of one dimensional vibratory manipulation would validate<br />

the analysis presented in Chapter 8 and would pave the way for a two dimensional implementation.<br />

Further analysis is required to combine the limiting cases of impulsive manipulation<br />

with a given striker behavior in order to characterize the motion of an object subject to such<br />

excitation.<br />

Kinematic constraint on objects<br />

Tapping an object subject to some kinematic constraint (such as an object in contact with<br />

other fixed objects) likely provides some additional stability or reduction in error for a<br />

given tap. This sort of tapping would be useful for parts assembly.<br />

Other forms of impulsive manipulation<br />

Exploration of impulsive manipulation where the object becomes airborne in between taps<br />

would involve some different mechanics but would be applicable to different types of vibratory<br />

parts feeding systems.


114 CHAPTER 9. CONCLUSIONS


Appendix A<br />

Impact Analysis<br />

This appendix contains the details of the mechanics of impact which are used throughout<br />

this thesis. <strong>The</strong> primary result of this appendix is that any impulse within a friction cone in<br />

impulse space at the contact point can be generated by an impact. <strong>The</strong> details of why this<br />

is so and how to determine the striker parameters for a given impulse are covered in this<br />

appendix.<br />

Since I will be making detailed references to the mechanics of two dimensional impact<br />

with friction, as described in Wang and Mason [63], I first summarize their results. With<br />

a few simplifying assumptions, I then show that any impulse within a friction cone in impulse<br />

space can be generated and describe the procedure to calculate the required striker<br />

parameters for a particular impulse. Finally, the impact parameters for some specific strikers<br />

and striker/object geometries are presented.<br />

A.1 Two dimensional rigid-body collisions with friction<br />

I have used slightly different notation and terminology than Wang and Mason [63] to summarizing<br />

their results here. What Wang and Mason call Body 1 and Body 2, I refer to as<br />

the “object” and “striker” respectively. Quantities related to the striker have a subscript s,<br />

e.g. M s , x s ,andy s ; quantities related to the object do not have a subscript, e.g. M, x, and<br />

y. Whereas Wang and Mason express moments of inertia in terms of the mass and radius<br />

and gyration, here I simply use the letter I. Finally, instead of using for the coefficient of<br />

friction between the striker and the object, I have substituted r to differentiate it from the<br />

coefficient of friction between the object and the support surface.<br />

A.1.1<br />

Equations of motion<br />

We assume that two rigid bodies, the striker and the object, collide at a single contact point.<br />

(See Figure A.1.) We attach a frame at the contact point with the y axis normal to the<br />

boundary and pointing into the object. <strong>The</strong> COM of the object relative to this frame is (x; y),<br />

the COM of the striker, (x s ;y s ). <strong>The</strong>ir respective linear velocities at the time of impact are<br />

(_x 0 ; _y 0 ) and (_x s0 ; _y s0 ). <strong>The</strong>re may also be rotational velocities, given by _ 0 and _ s0 . <strong>The</strong><br />

115


116 APPENDIX A. IMPACT ANALYSIS<br />

initial normal (compression) velocity is:<br />

C 0 =(_y 0 , _ 0 x) , (_y s0 , _ s0 x s )<br />

(A.1)<br />

<strong>The</strong> initial tangential (sliding) velocity is:<br />

S 0 =(_x 0 + _ 0 y) , (_x s0 + _ s0 y s )<br />

(A.2)<br />

<strong>The</strong> impact process undergoes a compression and a restitution phase during which we<br />

assume that there is no motion of the object. We will, however, assume that the velocity of<br />

the object and striker changes continuously in accordance with accumulated impulse. <strong>The</strong><br />

effect of this impulse on the object velocities _x, _y, and _ is:<br />

M (_x , _x 0 )=P x<br />

M (_y , _y 0 )=P y<br />

(A.3)<br />

(A.4)<br />

I( _ , _ 0 )=P x y , P y x (A.5)<br />

<strong>The</strong>re is an analogous set of equations for the effect of the impulse on the striker velocities.<br />

As the collision progresses, the accumulated impulse will change the sliding velocity<br />

(tangential velocity between the object and the striker) and the compression velocity (relative<br />

velocity along the contact normal). By combining the dynamics and the kinematics,<br />

we can express these velocities in terms of the accumulated impulse:<br />

where<br />

S = S 0 + B 1 P x , B 3 P y<br />

C = C 0 , B 3 P x + B 2 P y<br />

B 1 = 1 M s<br />

+ 1 M + y2<br />

I + y2 s<br />

I s<br />

(A.6)<br />

(A.7)<br />

(A.8)<br />

B 2 = 1 + 1 M s M + x2<br />

I + x2 s<br />

(A.9)<br />

I s<br />

B 3 = xy<br />

I + x sy s<br />

(A.10)<br />

I s<br />

where M s is the mass of the striker and I s its moment of inertia. (Note that the above expression<br />

for B 3 is correct; it appears incorrectly in Equation 21 in (Wang and Mason [63]).)<br />

A.1.2<br />

Impulse space constructs<br />

Routh’s method [54] involves constructing a number of lines in impulse space along which<br />

the state of the collision travels or at which the state of the collision changes.<br />

<strong>The</strong> first two lines are related to the relative normal and tangential velocity of the two<br />

objects. By setting the equation A.7 to zero, we get the line of maximum compression (C):<br />

C 0 , B 3 P x + B 2 P y =0<br />

(A.11)


A.1. TWO DIMENSIONAL RIGID-BODY COLLISIONS WITH FRICTION 117<br />

object<br />

y<br />

(x, y)<br />

x<br />

(x s,y s )<br />

striker<br />

Figure A.1: Impact geometry and notation.<br />

By setting the equation A.6 to zero, we get the line of sticking (S):<br />

S 0 + B 1 P x , B 3 P y =0<br />

(A.12)<br />

We are also interested in the line of termination, obtained by setting equation A.7 to ,eC 0<br />

(1 + e)C 0 , B 3 P x + B 2 P y =0 (A.13)<br />

This is the line at which the collision will end under Newton’s law of restitution (to be<br />

discussed further in section A.1.4. <strong>The</strong> fourth line of interest is the line of limiting friction<br />

(L), given by<br />

P x = ,s r P y<br />

(A.14)<br />

where r is the coefficient of friction between the striker and the object and s is the sign of<br />

the initial sliding velocity S 0 :<br />

s = S 0<br />

jS 0 j<br />

if S 0 6=0 (A.15)<br />

<strong>The</strong>re is one more line of interest which will be discussed in the next subsection.<br />

A.1.3<br />

Impact process<br />

As impact progresses through the compression and restitution phases, impulse along the<br />

P y axis will accumulate, changing the normal velocity until the collision is over. Impulse<br />

along the P x axis accumulates due to friction between the striker and the object in accordance<br />

with Coulomb friction.<br />

<strong>The</strong> collision follows a path in impulse space starting at the origin, and initially following<br />

the line of limiting friction. <strong>The</strong> collision continues (in a sliding mode) along this<br />

line until the collision is over or until it crosses the line of sticking. At that point, the object<br />

either sticks (then following the line of sticking) or slides in the opposite direction. If the<br />

line of sticking lies within the friction cone in impulse space at this intersection point, then


118 APPENDIX A. IMPACT ANALYSIS<br />

the object will stick, otherwise, the object will slide, entering the regime of reversed sliding<br />

in which its progress through impulse space along the line of reversed limiting friction (RF):<br />

dP x = s r dP y<br />

(A.16)<br />

To summarize, the collision accumulates impulse along the line of limiting friction until it<br />

intersects with the line of sticking, at which point, it will follow either the line of sticking<br />

or the line of reversed friction, whichever is steeper, until the collision ends.<br />

A.1.4<br />

End of the collision<br />

<strong>The</strong> end of the collision is determined by the restitution law. Wang and Mason [63] give results<br />

for both Newton’s law and for Poisson’s hypothesis. Since I have adopted Poisson’s<br />

law in this thesis, I only summarize those results here. Note that Newton’s law, Poisson’s<br />

hypothesis, and Stronge’s internal dissipation hypothesis all give the same answer<br />

for direct central impacts (when B 3 =0and S 0 =0) and when the impact remains in a<br />

sliding mode. Only when sliding stops or reverses direction do they differ. See Section 3.1,<br />

Stronge [57], or Wang and Mason [63] for further discussion.<br />

Poisson’s law terminates the impact based on the normal impulse accumulated during<br />

compression and restitution (P r and P c respectively). It defines the coefficient of restitution<br />

as:<br />

e = P r<br />

(A.17)<br />

P c<br />

A.1.5<br />

Impulse solutions<br />

<strong>The</strong> path of the accumulated impulse through impulse space depends upon the relative<br />

slopes and intersection points of the line of sticking, the line of maximum compression,<br />

and the line of limiting friction.<br />

We classify a collision by three criteria: whether the collision stopped sliding or not; if<br />

so, whether it entered a sticking mode or a reversed sliding mode; and whether it changed<br />

modes during the compression or restitution phase of the impact. We can classify the collision<br />

using the coefficient of friction between the striker and the object r and the quantities:<br />

P d = sS 0 (B 2 + sB 3 )<br />

P q = ,C 0 (B 1 + sB 3 )<br />

s = , B 3<br />

B 1<br />

Table A.1 shows the different collision modes in terms of these four quantities.<br />

<strong>The</strong> resultant impulses are:<br />

(A.18)<br />

(A.19)<br />

(A.20)<br />

sliding:<br />

P x = ,s r P y<br />

C 0<br />

P y = ,(1 + e)<br />

B 2 + s r B 3<br />

(A.21)<br />

(A.22)


A.2. GENERATING IMPULSE WITHIN A FRICTION CONE 119<br />

r > j s j r < j s j<br />

P d > (1 + e)P q<br />

sliding<br />

(1 + e)P q >P d >P q R-sticking R-reversed sliding<br />

P q >P d C-sticking C-reversed sliding<br />

Table A.1: Sliding modes classified by the parameters P d , P q , s ,and r . In addition, the<br />

reversed sliding modes require that S 0 B 3 > 0.<br />

R-sticking:<br />

C-sticking:<br />

R-reversed sliding:<br />

C-reversed sliding:<br />

where<br />

P x = B 3P y , S 0<br />

B 1<br />

C 0<br />

P y = ,(1 + e)<br />

B 2 + s r B 3<br />

P x = B 3P y , S 0<br />

B 1<br />

P y = ,(1 + e) B 1C 0 + B 3 S 0<br />

B 1 B 2 , B 2 3<br />

<br />

2S 0<br />

P x = s r P y ,<br />

B 3 + s r B 1<br />

C 0<br />

P y = ,(1 + e)<br />

B 2 + s r B 3<br />

<br />

2S 0<br />

<br />

P x = s r P y ,<br />

B 3 + s r B 1<br />

1+e<br />

P y = ,<br />

C 0 + 2s rB 3 S 0<br />

B 2 , s r B 3 B 3 + s r B 1<br />

s =<br />

S<br />

0<br />

jS 0j<br />

if S 0 6=0<br />

1 if S 0 =0<br />

<br />

<br />

(A.23)<br />

(A.24)<br />

(A.25)<br />

(A.26)<br />

(A.27)<br />

(A.28)<br />

(A.29)<br />

(A.30)<br />

(A.31)<br />

A.2 Generating impulse within a friction cone<br />

<strong>The</strong> ability to generate any impulse within a friction cone in impulse space likely holds in<br />

general, but I will only demonstrate it here under two assumptions appropriate for the sort<br />

of manipulation in this thesis.<br />

<strong>The</strong> first assumption is that the striker has no angular velocity when it collides with<br />

the object. Thus the striker is controlled by two parameters: v s0 , the linear velocity of the<br />

striker, and , the angle of incidence of the striker (measured CCW from the normal). <strong>The</strong><br />

initial compression and sliding velocities are then:<br />

S 0 = v s sin <br />

(A.32)


120 APPENDIX A. IMPACT ANALYSIS<br />

C 0 = ,v s cos <br />

(A.33)<br />

This allows the striker velocity to be factored out from all the equations for impulse delivered<br />

by an impact (equations A.21 through A.24). Consequently, each mode of impact can<br />

be characterized by the range of values of Px<br />

P y<br />

that it can produce.<br />

<strong>The</strong> second assumptions is:<br />

the striker has infinite mass or infinite angular inertia, or<br />

the COM of the striker always lies on the contact normal.<br />

This serve the purpose of making the B i constant with respect to the angle of incidence<br />

. <strong>The</strong> slopes of the line of sticking and the line of maximum compression then remain<br />

constant, only moving up and down with changes in .<br />

By varying the angle of incidence of the striker, a continuous range of impulse ratios<br />

from , r to r can be generated, thus covering all the rays in the friction cone in impulse<br />

space. Any particular point along a ray can be achieved by using the appropriate striker<br />

velocity.<br />

Although being able to generate any impulse in the friction cone in impulse space is<br />

not too difficult to see by graphical reasoning (by taking advantage of Routh’s method),<br />

I present the results of a more analytical approach here; these results are useful for determining<br />

the actual striker velocities for a desired impulse.<br />

A.2.1<br />

General case<br />

Table A.2 gives a summary of the impulse ratio ranges and impact modes for ranges of <br />

under the assumption that reversed sliding does not occur, i.e. r > j s j. <strong>The</strong> collision<br />

mode will be either sliding, R-sticking, or C-sticking. <strong>The</strong>se expressions are the result of<br />

applying the conditions and solutions in Section A.1.5 and in Table A.1.<br />

Table A.3 gives the same information under the assumption that reversed sliding can<br />

occur, i.e. r < j, s j. <strong>The</strong> collision mode will be either sliding, R-reversed sliding, or<br />

C-reversed sliding. <strong>The</strong> behavior now depends upon whether B 3 is positive or negative<br />

because S 0 B 3 must be positive in order for reversed sliding to occur. Although the impulse<br />

ratio ranges were calculated from the impulse solutions in Section A.1.5 (originally from<br />

Wang and Mason [63]), the expressions for the ranges for this case are somewhat different<br />

than what Table A.1 (also from [63]) suggests.<br />

In each case, the impulse ratio ranges, taken together, cover the entire friction cone,<br />

from Px<br />

=<br />

P y<br />

, r to Px<br />

=<br />

P y<br />

r . In addition, there is always some angle of incidence beyond<br />

which the magnitude of the impulse ratio does not increase. Note that the direction of the<br />

impulse will in general be different from the angle of incidence of the striker.<br />

In order to determine the striker velocities to generate a given impulse at a given contact<br />

point, first determine the impulse ratio of the desired impulse. If j Px<br />

P y<br />

j > r , then this<br />

impulse cannot be generated by an impact at this point. Next, determine whether reversed<br />

sliding can occur, and use Table A.2 or A.3 as appropriate to determine which impact mode<br />

is required to generate the desired impulse ratio. <strong>The</strong>n solve the corresponding equations<br />

in Section A.1.5 for the initial compression and sliding velocities C 0 and S 0 . <strong>The</strong> striker<br />

velocity v s0 and the angle of incidence can be solved using Equations A.32 and A.33.


mode<br />

A.3. SPECIAL CASES 121<br />

range<br />

sliding tan ,1h,(1 r B1,B3<br />

+ e)<br />

B 2, rB 3i<br />

R-sticking<br />

tan<br />

tan ,1h,(1 r B1,B3<br />

+ e)<br />

B 2, rB 3i<br />

r B1,B3<br />

tan ,1h,<br />

B 2, rB 3i<br />

0<br />

C-sticking<br />

,1h <br />

C-sticking 0 tan<br />

R-sticking<br />

sliding<br />

,1 rB1+B3<br />

tan<br />

B 2+ rB 3<br />

tan ,1h(1 + e) rB1+B3<br />

B 2+ r B 3i<br />

rB1,B3<br />

,1h,<br />

B 2, rB 3i<br />

rB 1+B 3<br />

B 2+ rB 3i<br />

B<br />

tan ,1h(1 r B1+B3<br />

+ e)<br />

B 2+ rB 3i<br />

<br />

P<br />

P x<br />

P y<br />

x<br />

P y<br />

r Px<br />

P y<br />

B<br />

B 3<br />

+ rB1,B3<br />

B 1 (1+e)B 1<br />

3<br />

B 1<br />

3<br />

B 1<br />

, rB1+B3<br />

(1+e)B 1<br />

range<br />

Px<br />

P y<br />

Px<br />

P y<br />

Px<br />

P y<br />

P x<br />

P y<br />

= r<br />

= , r<br />

B3<br />

+ rB1,B3<br />

B 1 (1+e)B 1<br />

B3<br />

B 1<br />

B3<br />

B 1<br />

, rB1+B3<br />

(1+e)B 1<br />

, r<br />

Table A.2: Generating impact in a friction cone in impulse space, sticking or sliding case.<br />

This table shows the impact modes, ranges, and resulting Px<br />

P y<br />

ranges required to generate<br />

an impulse in any part of the friction cone in impulse space.<br />

A.2.2<br />

Generalized central impacts<br />

When B 3 =0, the mechanics of impact become much simpler. This occurs when (in addition<br />

to the two assumptions of this section), the COM of the object lies on the contact<br />

normal. <strong>The</strong> line of maximum compression becomes:<br />

P y = ,C 0<br />

(A.34)<br />

B 2<br />

and the line of sticking becomes:<br />

P x = ,S 0<br />

(A.35)<br />

B 1<br />

Like Tables A.2 and A.3 for the more general case, Table A.4 shows the relationship between<br />

ranges and impulse ratio ranges for the different sliding modes. However, this situation<br />

is far simpler than might appear from the table. So long as the desired impulse is within<br />

the friction cone in impulse space, that impulse can be achieved by the initial compression<br />

and sliding velocities:<br />

C 0 = , B 2<br />

1+e P y (A.36)<br />

S 0 = ,B 1 P x<br />

(A.37)<br />

<strong>The</strong> striker velocity v s0 and the angle of incidence can again be solved using Equations<br />

A.32 and A.33.<br />

A.3 Special Cases<br />

This section briefly looks at a number of specific instances of striker and striker/object<br />

situations relevant to the experiments performed in this thesis.


122 APPENDIX A. IMPACT ANALYSIS<br />

mode<br />

B 3 > 0<br />

range<br />

P x<br />

P y<br />

range<br />

P<br />

sliding < 0<br />

,1h x<br />

=<br />

P y<br />

r<br />

<br />

C-reversed sliding 0 < tan rB 1+B 3<br />

B 2+ r B 3i<br />

r > Px<br />

e,1<br />

P y<br />

r e+1<br />

R-reversed sliding tan ,1h(1 + e) B3+rB1<br />

B 2+ rB 3i<br />

tan ,1h(1 B3+r B1<br />

+ e)<br />

B 2+ r B 3i<br />

e,1<br />

r Px<br />

e+1 P y<br />

, r<br />

sliding tan ,1h(1 + e) B3+rB1<br />

B 2+ rB 3i<br />

P<br />

<br />

x<br />

P y<br />

= , r<br />

mode<br />

B 3 < 0<br />

range<br />

sliding tan ,1h(1 + e) B3,rB1<br />

B 2+ rB 3i<br />

tan ,1h(1 + e) B3,rB1<br />

B 2+ rB 3i ,1h B<br />

tan 3, rB 1<br />

B 2+ rB 3i<br />

R-reversed sliding<br />

,1h B<br />

C-reversed sliding tan<br />

3, rB 1<br />

B 2+ rB 3i<br />

sliding 0 <<br />

P<br />

P x<br />

P y<br />

x<br />

P y<br />

range<br />

= r<br />

r Px<br />

1,e<br />

P y<br />

r 1+e<br />

1,e<br />

< 0 r Px<br />

1+e P y<br />

> , r<br />

P x<br />

P y<br />

= , r<br />

Table A.3: Generating impact in a friction cone in impulse space, reversed sliding or sliding<br />

case. This table shows the impact modes, ranges, and resulting Px<br />

P y<br />

ranges required to<br />

generate an impulse in any part of the friction cone in impulse space, both when B 3 > 0<br />

and when B 3 < 0.


A.3. SPECIAL CASES 123<br />

mode<br />

range<br />

sliding tan ,1h, r (1 + e) B1<br />

B 2i<br />

R-sticking tan ,1h, r (1 + e) B1<br />

B 2i<br />

tan ,1h, B 1 r B 2i<br />

P<br />

P x<br />

P y<br />

range<br />

x<br />

P y<br />

= r<br />

r Px<br />

P y<br />

<br />

r<br />

1+e<br />

C-sticking tan ,1h, r<br />

B 1<br />

B 2i<br />

0<br />

r<br />

1+e<br />

Px<br />

P y<br />

0<br />

C-sticking 0 tan ,1h r<br />

B 1<br />

B 2i<br />

R-sticking tan ,1h r<br />

B 1<br />

B 2i<br />

sliding<br />

tan ,1h r (1 + e) B1<br />

B 2i<br />

0 Px<br />

P y<br />

, r<br />

1+e<br />

tan ,1h r (1 + e) B1<br />

B 2i<br />

, r<br />

Px<br />

1+e P y<br />

, r<br />

P<br />

<br />

x<br />

=<br />

P y<br />

, r<br />

Table A.4: Generalized central impacts. This table shows the relationship between impulse<br />

ratios and ranges for various impact modes under the assumption that B 3 =0. Reversed<br />

sliding never occurs under this condition.<br />

A.3.1<br />

Spherical striker<br />

With a spherical striker (I s = 2 5 M sR 2 s), the B i are:<br />

B 1 = 7<br />

2M s<br />

+ 1 M + y2<br />

I +<br />

(A.38)<br />

B 2 =<br />

1 + 1 M s M + x2<br />

I<br />

B 3 = xy<br />

I<br />

(A.39)<br />

(A.40)<br />

A.3.2<br />

Spherical striker and circular object<br />

With a circular object (I = 1 2 MR2 ), the B i are:<br />

B 1 = 7<br />

2M s<br />

+ 3 M<br />

B 2 = 1 M s<br />

+ 1 M<br />

B 3 =0<br />

(A.41)<br />

(A.42)<br />

(A.43)<br />

A.3.3<br />

Long thin rod striker<br />

With a long thin rod striker with infinite angular inertia, the B i are:<br />

B 1 =<br />

1 M s<br />

+ 1 M + y2<br />

I<br />

(A.44)


124 APPENDIX A. IMPACT ANALYSIS<br />

B 2 =<br />

1 M s<br />

+ 1 M + x2<br />

I +<br />

B 3 = xy<br />

I<br />

(A.45)<br />

(A.46)<br />

A.3.4<br />

Long thin rod striker with circular object<br />

With a circular object (I = 1 2 MR2 ), the B i are:<br />

B 1 = 1 M s<br />

+ 3 M<br />

(A.47)<br />

B 2 = 1 M s<br />

+ 1 M<br />

B 3 =0<br />

(A.48)<br />

(A.49)


Bibliography<br />

[1] S. Akella. Robotic <strong>Manipulation</strong> for Parts Transfer and Orienting: Mechanics, Planning, and<br />

Shape Uncertainty. PhD thesis, <strong>Carnegie</strong> <strong>Mellon</strong> University, 1996.<br />

[2] S. Akella, W. H. Huang, K. M. Lynch, and M. T. Mason. Planar manipulation on a<br />

conveyor with a one joint robot. In G. Giralt and G. Hirzinger, editors, International<br />

Symposium on <strong>Robotics</strong> Research: <strong>The</strong> Seventh International Symposium. Springer-Verlag,<br />

1996.<br />

[3] S. Akella, W. H. Huang, K. M. Lynch, and M. T. Mason. Sensorless parts feeding with<br />

a one joint robot. In J.-P. Laumond and M. Overmars, editors, Algorithms for Robotic<br />

Motion and <strong>Manipulation</strong>: 1996 Workshop on the Algorithmmic Foundations of <strong>Robotics</strong>,<br />

pages 229–38. A K Peters, 1997.<br />

[4] S. Akella and M. T. Mason. Posing polygonal objects in the plane by pushing. In IEEE<br />

International Conference on <strong>Robotics</strong> and Automation, pages 2255–2262, Nice, France,<br />

1992.<br />

[5] J. C. Alexander and J. H. Maddocks. Bounds on the friction-dominated motion of a<br />

pushed object. International Journal of <strong>Robotics</strong> Research, 12(3):231–248, 1993.<br />

[6] B. Armstrong-Hélouvry. ControlofMachineswithFriction. Kluwer Academic Publishers,<br />

1991.<br />

[7] C. N. Bapat, S. Sankar, and N. Popplewell. Repeated impacts on a sinusoidally vibrating<br />

table reappraised. Journal of Sound and Vibration, 108(1):99–115, 1986.<br />

[8] D. R. Berkowitz and J. Canny. A comparison of real and simulated designs for vibratory<br />

parts feeding. In IEEE International Conference on <strong>Robotics</strong> and Automation, 1997.<br />

[9] K. Böhringer, V. Bhatt, and K. Goldberg. Sensorless manipulation using transverse<br />

vibrations of a plate. In IEEE International Conference on <strong>Robotics</strong> and Automation, volume<br />

2, pages 1989–96, 1995.<br />

[10] G. Boothroyd. Assembly Automation and Product Design. Marcel Dekker, 1992.<br />

[11] G. Boothroyd, C. Poli, and L. E. Murch. Automatic Assembly. Marcel Dekker, 1982.<br />

[12] F. P. Bowden and D. Tabor. <strong>The</strong> Friction and Lubrication of Solids: Part I. Clarendon<br />

Press, 1950.<br />

125


126 BIBLIOGRAPHY<br />

[13] F. P. Bowden and D. Tabor. <strong>The</strong> Friction and Lubrication of Solids: Part II. OxfordUniversity<br />

Press, 1964.<br />

[14] F. P. Bowden and D. Tabor. Friction: An Introduction to Tribology. Anchor<br />

Press/Doubleday, 1973. Reprinted by Robert E. Krieger Publishing Company, 1982.<br />

[15] R. M. Brach. Mechanical Impact Dynamics: Rigid Body Collisions. John Wiley & Sons,<br />

1991.<br />

[16] M. Bühler and D. E. Koditschek. From stable to chaotic juggling: <strong>The</strong>ory, simulation,<br />

and experiments. In IEEE International Conference on <strong>Robotics</strong> and Automation, pages<br />

1976–1981, Cincinnati, OH, 1990.<br />

[17] R. R. Burridge, A. A. Rizzi, and D. E. Koditschek. Toward a dynamical pick and place.<br />

In IEEE/RSJ International Conference on Intelligent Robots and Systems, pages 2: 292–297,<br />

1995.<br />

[18] M. Caine. <strong>The</strong> design of shape interactions using motion constraints. In IEEE International<br />

Conference on <strong>Robotics</strong> and Automation, 1994.<br />

[19] J. Canny and K. Goldberg. A risc approach to robotics. IEEE <strong>Robotics</strong> & Automation<br />

Magazine, 1(1):26–8, March 1994.<br />

[20] J. F. Canny and K. Y. Goldberg. “risc” industrial robotics: Recent results and open<br />

problems. In IEEE International Conference on <strong>Robotics</strong> and Automation, pages 1951–8,<br />

1994.<br />

[21] A. Christiansen, A. Edwards, and C. Coello. Automated design of part feeders using<br />

a genetic algorithm. In IEEE International Conference on <strong>Robotics</strong> and Automation,<br />

volume 1, pages 846–51, 1996.<br />

[22] M. A. Erdmann. An exploration of nonprehensile two-palm manipulation: Planning<br />

and execution. In G. Giralt and G. Hirzinger, editors, International Symposium<br />

on <strong>Robotics</strong> Research: <strong>The</strong> Seventh International Symposium. Springer-Verlag, 1996.<br />

[23] M. A. Erdmann and M. T. Mason. An exploration of sensorless manipulation. IEEE<br />

Transactions on <strong>Robotics</strong> and Automation, 4(4):369–379, Aug. 1988.<br />

[24] R. Feynman. <strong>The</strong> Feynman Lectures on Physics, volume 1. Addison-Wesley, 1963.<br />

[25] K. Y. Goldberg. Orienting polygonal parts without sensors. Algorithmica, 10:201–225,<br />

1993.<br />

[26] W. Goldsmith. Impact: <strong>The</strong> <strong>The</strong>ory and Physical Behaviour of Colliding Solids. Edward<br />

Arnold (Publishers), 1960.<br />

[27] S. Goyal, A. Ruina, and J. Papadopoulos. Planar sliding with dry friction. Part 1. Limit<br />

surface and moment function. Wear, 143:307–330, 1991.<br />

[28] S. Goyal, A. Ruina, and J. Papadopoulos. Planar sliding with dry friction. Part 2.<br />

Dynamics of motion. Wear, 143:331–352, 1991.


BIBLIOGRAPHY 127<br />

[29] T. Higuchi. Application of electromagnetic impulsive force to precise positioning tools<br />

in robot systems. In O. Faugeras and G. Giralt, editors, International Symposium on<br />

<strong>Robotics</strong> Research: <strong>The</strong> Third International Symposium. MIT Press, 1986.<br />

[30] H. Hitakawa. Advanced parts orientation system has wide application. Assembly<br />

Automation, 8(3):147–150, 1988.<br />

[31] P. J. Holmes. <strong>The</strong> dynamics of repeated impacts with a sinusoidally vibrating table.<br />

Journal of Sound and Vibration, 84(2):173–189, 1982.<br />

[32] R. D. Howe and M. R. Cutkosky. Practical force-motion models for sliding manipulation.<br />

International Journal of <strong>Robotics</strong> Research, 15(6):557–72, Dec. 1996.<br />

[33] W. H. Huang, E. P. Krotkov, and M. T. Mason. <strong>Impulsive</strong> manipulation. In IEEE<br />

International Conference on <strong>Robotics</strong> and Automation, volume 1, pages 120–5, 1995.<br />

[34] W. H. Huang and M. T. Mason. Limiting cases of impulsive manipulation. Technical<br />

Report CMU–RI–TR–96–24, <strong>The</strong> <strong>Robotics</strong> <strong>Institute</strong>, <strong>Carnegie</strong> <strong>Mellon</strong> University,<br />

September 1996.<br />

[35] W. H. Huang and M. T. Mason. Mechanics for vibratory manipulation. In IEEE International<br />

Conference on <strong>Robotics</strong> and Automation, 1997.<br />

[36] J. Krishnasamy. Mechanics of Entrapment with Application to Design of Industrial Parts<br />

Feeders. PhD thesis, Massachusetts <strong>Institute</strong> of Technology, 1996.<br />

[37] M. Kurisu and T. Yoshikawa. Trajectory planning for an object in pushing operation.<br />

In 1994 Japan–USA Symposium on Flexible Automation, pages 1009–1016, 1994.<br />

[38] K. M. Lynch. <strong>The</strong> mechanics of fine manipulation by pushing. In IEEE International<br />

Conference on <strong>Robotics</strong> and Automation, pages 2269–2276, Nice, France, 1992.<br />

[39] K. M. Lynch. Estimating the friction parameters of pushed objects. In IEEE/RSJ International<br />

Conference on Intelligent Robots and Systems, pages 186–193, Yokohama, Japan,<br />

1993.<br />

[40] K. M. Lynch. Nonprehensile Robotic <strong>Manipulation</strong>. PhD thesis, <strong>Carnegie</strong> <strong>Mellon</strong> University,<br />

March 1996. Available as <strong>Carnegie</strong> <strong>Mellon</strong> University <strong>Robotics</strong> <strong>Institute</strong> Technical<br />

Report CMU-RI-TR–96–05.<br />

[41] K. M. Lynch and M. T. Mason. Controllability of pushing. In IEEE International Conference<br />

on <strong>Robotics</strong> and Automation, Nagoya, Japan, 1995.<br />

[42] M. Mani and W. Wilson. A programmable orienting system for flat parts. In North<br />

American Manufacturing Research <strong>Institute</strong> Conference XIII, 1985.<br />

[43] M. T. Mason. Mechanics and planning of manipulator pushing operations. International<br />

Journal of <strong>Robotics</strong> Research, 5(3):53–71, Fall 1986.<br />

[44] M. T. Mason and J. K. Salisbury, Jr. Robot Hands and the Mechanics of <strong>Manipulation</strong>. <strong>The</strong><br />

MIT Press, 1985.


128 BIBLIOGRAPHY<br />

[45] T. McGeer. Passive dynamic walking. International Journal of <strong>Robotics</strong> Research, 9(2):62–<br />

82, 1990.<br />

[46] T. McGeer. Passive walking with knees. In IEEE International Conference on <strong>Robotics</strong><br />

and Automation, pages 1640–1645, 1990.<br />

[47] B. Mirtich and J. Canny. Impulse-based dynamic simulation. In K. Goldberg,<br />

D. Halperin, J.-C. Latombe, and R. Wilson, editors, Algorithmic Foundations of <strong>Robotics</strong>,<br />

pages 407–418. A. K. Peters, Boston, MA, 1995.<br />

[48] E. Paljug and X. Yun. Experimental results of two robot arms manipulating large<br />

objects. In IEEE International Conference on <strong>Robotics</strong> and Automation, 1993.<br />

[49] M. A. Peshkin and A. C. Sanderson. <strong>The</strong> motion of a pushed, sliding workpiece. IEEE<br />

Journal of <strong>Robotics</strong> and Automation, 4(6):569–598, Dec. 1988.<br />

[50] M. A. Peshkin and A. C. Sanderson. Planning robotic manipulation strategies for<br />

workpieces that slide. IEEE Journal of <strong>Robotics</strong> and Automation, 4(5):524–531, Oct. 1988.<br />

[51] D. T. Pham, K. C. Cheung, and S. H. Yeo. Initial motion of a rectangular object being<br />

pushed or pulled. In IEEE International Conference on <strong>Robotics</strong> and Automation, pages<br />

1046–1050, Cincinnati, OH, 1990.<br />

[52] W. H. Press, S. A. Teukolsky, W. T. Vetterling, and B. P. Flannery. Numerical Recipes in<br />

C. Cambridge University Press, second edition, 1992.<br />

[53] A. A. Rizzi and D. E. Koditschek. Progress in spatial robot juggling. In IEEE International<br />

Conference on <strong>Robotics</strong> and Automation, pages 775–780, Nice, France, 1992.<br />

[54] E. Routh. <strong>The</strong> Elementary Part of a Treatise on the Dynamics of a System of Rigid Bodies.<br />

Dover Publications, eighth edition, 1960. Originally published by Macmillan and<br />

Company, 1905.<br />

[55] K. Salisbury, B. Eberman, M. Levin, and W. Townsend. <strong>The</strong> design and control of<br />

an experimental whole-arm manipulator. In <strong>Robotics</strong> Research: <strong>The</strong> Fifth International<br />

Symposium. MIT Press, 1990.<br />

[56] S. Schaal and C. G. Atkeson. Open loop stable control strategies for robot juggling.<br />

In IEEE International Conference on <strong>Robotics</strong> and Automation, pages 3:913–918, Atlanta,<br />

GA, 1993.<br />

[57] W. J. Stronge. Rigid body collisions with friction. Proceedings of the Royal Society,<br />

London, A431:169–181, 1990.<br />

[58] P. J. Swanson, R. R. Burridge, and D. E. Koditschek. Global asymptotic stability of a<br />

passive juggler: A parts feeding strategy. In IEEE International Conference on <strong>Robotics</strong><br />

and Automation, pages 1983–8, 1995.<br />

[59] J. C. Trinkle, A. O. Farahat, and P. F. Stiller. First-order stability cells of active multirigid-body<br />

systems. IEEE Transactions on <strong>Robotics</strong> and Automation, 11(4):545–557, Aug.<br />

1995.


BIBLIOGRAPHY 129<br />

[60] J. C. Trinkle, R. C. Ram, A. O. Farahat, and P. F. Stiller. Dexterous manipulation planning<br />

and execution of an enveloped slippery workpiece. In IEEE International Conference<br />

on <strong>Robotics</strong> and Automation, pages 2: 442–448, 1993.<br />

[61] N. B. Tufillaro and A. M. Albano. Chaotic dynamics of a bouncing ball. American<br />

Journal of Physics, 54(10):939–944, October 1986.<br />

[62] K. Voyenli and E. Eriksen. On the motion of an ice hockey puck. American Journal of<br />

Physics, 53(12):1149–53, December 1985.<br />

[63] Y. Wang and M. T. Mason. Two-dimensional rigid-body collisions with friction. ASME<br />

Journal of Applied Mechanics, 59:635–641, Sept. 1992.<br />

[64] K. Wiesenfeld and N. B. Tufillaro. Suppression of period doubling in the dynamics of<br />

a bouncing ball. Physica D, 26(1–3):321–335, May-June 1987.<br />

[65] Y. Yamagata and T. Higuchi. A micropositioning device for precision automatic assembly<br />

using impact force of piezoelectric elements. In IEEE International Conference<br />

on <strong>Robotics</strong> and Automation, 1995.<br />

[66] Y. Yamagata, T. Higuchi, N. Nakamura, and S. Hamamura. A micro mobile mechanism<br />

using thermal expansion and its theoretical analysis. In IEEE Micro Electro Mechanical<br />

Systems: An Investigation of Micro Structures, Sensors, Actuators, Machines, and<br />

Robotic Systems, pages 142–7, Oiso, Japan, Jan. 1994.<br />

[67] Y. Yamagata, T. Higuchi, H. Saeki, and H. Ishimaru. Ultrahigh vacuum precise positioning<br />

device utilizing rapid deformations of piezoelectric elements. Journal of Vacuum<br />

Science & Technology A, 8(6):4098–100, November/December 1990.<br />

[68] T. Yoshikawa and M. Kurisu. Identification of the center of friction from pushing an<br />

object by a mobile robot. In IEEE/RSJ International Conference on Intelligent Robots and<br />

Systems, pages 449–54, Osaka, Japan, 1991.<br />

[69] X. Yun. Object handling using two arms without grasping. International Journal of<br />

<strong>Robotics</strong> Research, 12(1), 1993.<br />

[70] C. Zhu, Y. Aiyama, T. Chawanya, and T. Arai. Releasing manipulation. In IEEE/RSJ International<br />

Conference on Intelligent Robots and Systems, volume 2, page 1, Osaka, Japan,<br />

Nov. 1996.<br />

[71] N. B. Zumel. A Nonprehensile Method for Reliable Parts Orienting. PhD thesis, <strong>Carnegie</strong><br />

<strong>Mellon</strong> University, Jan. 1997.

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!