Academia.eduAcademia.edu
Second Edition BALLISTICS Theory and Design of Guns and Ammunition © 2014 by Taylor & Francis Group, LLC © 2014 by Taylor & Francis Group, LLC Second Edition BALLISTICS Theory and Design of Guns and Ammunition Donald E. Carlucci and Sidney S. Jacobson Boca Raton London New York CRC Press is an imprint of the Taylor & Francis Group, an informa business © 2014 by Taylor & Francis Group, LLC MATLAB® is a trademark of The MathWorks, Inc. and is used with permission. The MathWorks does not warrant the accuracy of the text or exercises in this book. This book’s use or discussion of MATLAB® software or related products does not constitute endorsement or sponsorship by The MathWorks of a particular pedagogical approach or particular use of the MATLAB® software. CRC Press Taylor & Francis Group 6000 Broken Sound Parkway NW, Suite 300 Boca Raton, FL 33487-2742 © 2014 by Taylor & Francis Group, LLC CRC Press is an imprint of Taylor & Francis Group, an Informa business No claim to original U.S. Government works Version Date: 20130627 International Standard Book Number-13: 978-1-4665-6439-8 (eBook - PDF) This book contains information obtained from authentic and highly regarded sources. Reasonable efforts have been made to publish reliable data and information, but the author and publisher cannot assume responsibility for the validity of all materials or the consequences of their use. The authors and publishers have attempted to trace the copyright holders of all material reproduced in this publication and apologize to copyright holders if permission to publish in this form has not been obtained. If any copyright material has not been acknowledged please write and let us know so we may rectify in any future reprint. Except as permitted under U.S. Copyright Law, no part of this book may be reprinted, reproduced, transmitted, or utilized in any form by any electronic, mechanical, or other means, now known or hereafter invented, including photocopying, microfilming, and recording, or in any information storage or retrieval system, without written permission from the publishers. For permission to photocopy or use material electronically from this work, please access www.copyright.com (http:// www.copyright.com/) or contact the Copyright Clearance Center, Inc. (CCC), 222 Rosewood Drive, Danvers, MA 01923, 978-750-8400. CCC is a not-for-profit organization that provides licenses and registration for a variety of users. For organizations that have been granted a photocopy license by the CCC, a separate system of payment has been arranged. Trademark Notice: Product or corporate names may be trademarks or registered trademarks, and are used only for identification and explanation without intent to infringe. Visit the Taylor & Francis Web site at http://www.taylorandfrancis.com and the CRC Press Web site at http://www.crcpress.com © 2014 by Taylor & Francis Group, LLC MATLAB® is a trademark of The MathWorks, Inc. and is used with permission. The MathWorks does not warrant the accuracy of the text or exercises in this book. This book’s use or discussion of MATLAB® software or related products does not constitute endorsement or sponsorship by The MathWorks of a particular pedagogical approach or particular use of the MATLAB® software. CRC Press Taylor & Francis Group 6000 Broken Sound Parkway NW, Suite 300 Boca Raton, FL 33487-2742 © 2014 by Taylor & Francis Group, LLC CRC Press is an imprint of Taylor & Francis Group, an Informa business No claim to original U.S. Government works Printed on acid-free paper Version Date: 20130627 International Standard Book Number-13: 978-1-4665-6437-4 (Hardback) This book contains information obtained from authentic and highly regarded sources. Reasonable efforts have been made to publish reliable data and information, but the author and publisher cannot assume responsibility for the validity of all materials or the consequences of their use. The authors and publishers have attempted to trace the copyright holders of all material reproduced in this publication and apologize to copyright holders if permission to publish in this form has not been obtained. If any copyright material has not been acknowledged please write and let us know so we may rectify in any future reprint. Except as permitted under U.S. Copyright Law, no part of this book may be reprinted, reproduced, transmitted, or utilized in any form by any electronic, mechanical, or other means, now known or hereafter invented, including photocopying, microfilming, and recording, or in any information storage or retrieval system, without written permission from the publishers. For permission to photocopy or use material electronically from this work, please access www.copyright.com (http:// www.copyright.com/) or contact the Copyright Clearance Center, Inc. (CCC), 222 Rosewood Drive, Danvers, MA 01923, 978-750-8400. CCC is a not-for-profit organization that provides licenses and registration for a variety of users. For organizations that have been granted a photocopy license by the CCC, a separate system of payment has been arranged. Trademark Notice: Product or corporate names may be trademarks or registered trademarks, and are used only for identification and explanation without intent to infringe. Library of Congress Cataloging-in-Publication Data Carlucci, Donald E. Ballistics : theory and design of guns and ammunition / Authors, Donald E. Carlucci, Sidney S. Jacobson. -- 2nd edition. p. cm. Includes bibliographical references and index. ISBN 978-1-4665-6437-4 1. Ballistics. I. Jacobson, Sidney S. II. Title. UF820.C28 2014 623’.51--dc23 Visit the Taylor & Francis Web site at http://www.taylorandfrancis.com and the CRC Press Web site at http://www.crcpress.com © 2014 by Taylor & Francis Group, LLC 2013014363 To Peg C., Sandy J., and our families, without whose patience and support we could not have brought this work to completion © 2014 by Taylor & Francis Group, LLC © 2014 by Taylor & Francis Group, LLC Contents Preface to the Second Edition ...................................................................................................... xi Preface to the First Edition......................................................................................................... xiii Acknowledgments ........................................................................................................................xv Authors ........................................................................................................................................ xvii Part I Interior Ballistics 1. Introductory Concepts ...........................................................................................................3 1.1 Ballistic Disciplines.......................................................................................................4 1.2 Terminology ...................................................................................................................4 1.3 Units and Symbols ........................................................................................................5 2. Physical Foundation of Interior Ballistics .........................................................................7 2.1 Ideal Gas Law ................................................................................................................7 2.2 Other Gas Laws ........................................................................................................... 14 2.3 Thermophysics and Thermochemistry.................................................................... 15 2.4 Thermodynamics ........................................................................................................ 21 2.5 Combustion .................................................................................................................. 26 2.6 Solid Propellant Combustion .................................................................................... 38 2.7 Fluid Mechanics .......................................................................................................... 48 References ............................................................................................................................... 68 3. Analytic and Computational Ballistics ............................................................................ 69 3.1 Computational Goal ................................................................................................... 69 3.2 Lagrange Gradient ...................................................................................................... 70 3.3 Chambrage Gradient ................................................................................................ 100 3.4 Numerical Methods in Interior Ballistics .............................................................. 102 3.5 Sensitivities and Efficiencies ................................................................................... 106 References ............................................................................................................................. 109 4. Ammunition Design Practice ........................................................................................... 111 4.1 Stress and Strain........................................................................................................ 111 4.2 Failure Criteria .......................................................................................................... 115 4.3 Ammunition Types ................................................................................................... 119 4.4 Propellant Ignition .................................................................................................... 119 4.5 The Gun Chamber .................................................................................................... 120 4.6 Propellant Charge Construction ............................................................................. 121 4.7 Propellant Geometry ................................................................................................ 122 4.8 Cartridge Case Design ............................................................................................. 123 4.9 Projectile Design ....................................................................................................... 126 4.10 Shell Structural Analysis ......................................................................................... 126 4.11 Buttress Thread Design ............................................................................................ 148 vii © 2014 by Taylor & Francis Group, LLC Contents viii 4.12 Sabot Design .............................................................................................................. 156 References ............................................................................................................................. 165 Further Readings ................................................................................................................. 165 5. Weapon Design Practice .................................................................................................... 167 5.1 Fatigue and Endurance ............................................................................................ 167 5.2 Tube Design ............................................................................................................... 169 5.3 Gun Dynamics........................................................................................................... 175 5.4 Muzzle Devices and Associated Phenomena ....................................................... 181 Gun Dynamics Nomenclature ........................................................................................... 189 References ............................................................................................................................. 189 Further Readings ................................................................................................................. 190 Part II Exterior Ballistics 6. Introductory Concepts ....................................................................................................... 193 References ............................................................................................................................. 203 Further Reading ................................................................................................................... 203 7. Dynamics Review ............................................................................................................... 205 Reference ............................................................................................................................... 221 Further Readings ................................................................................................................. 221 8. Trajectories ...........................................................................................................................223 8.1 Vacuum Trajectory ....................................................................................................223 8.2 Simple Air Trajectory (Flat Fire).............................................................................. 231 8.3 Wind Effects on a Simple Air Trajectory ............................................................... 248 8.4 Generalized Point Mass Trajectory......................................................................... 261 8.5 Six Degree-of-Freedom (6-DOF) Trajectory........................................................... 270 8.6 Modified Point Mass Trajectory .............................................................................. 290 8.7 Probability of First Round Hit................................................................................. 301 References .............................................................................................................................305 Further Readings .................................................................................................................305 9. Linearized Aeroballistics .................................................................................................. 307 9.1 Linearized Pitching and Yawing Motions ............................................................309 9.2 Gyroscopic and Dynamic Stabilities ...................................................................... 320 9.3 Yaw of Repose............................................................................................................ 328 9.4 Roll Resonance .......................................................................................................... 329 References ............................................................................................................................. 331 10. Mass Asymmetries ............................................................................................................. 333 References ............................................................................................................................. 335 11. Lateral Throwoff ................................................................................................................. 337 11.1 Static Imbalance ........................................................................................................340 11.2 Dynamic Imbalance..................................................................................................342 References .............................................................................................................................346 © 2014 by Taylor & Francis Group, LLC Contents ix 12. Swerve Motion .................................................................................................................... 347 12.1 Aerodynamic Jump................................................................................................... 347 12.2 Epicyclic Swerve ........................................................................................................ 351 12.3 Drift ............................................................................................................................. 352 Reference ............................................................................................................................... 353 13. Nonlinear Aeroballistics ................................................................................................... 355 13.1 Nonlinear Forces and Moments.............................................................................. 355 13.2 Bilinear and Trilinear Moments.............................................................................. 358 References ............................................................................................................................. 361 Part III Terminal Ballistics 14. Introductory Concepts ....................................................................................................... 365 15. Penetration Theories .......................................................................................................... 369 15.1 Penetration and Perforation of Metals ................................................................... 369 15.2 Penetration and Perforation of Concrete ............................................................... 391 15.3 Penetration and Perforation of Soils ....................................................................... 399 15.4 Penetration and Perforation of Ceramics ............................................................... 406 15.5 Penetration and Perforation of Composites........................................................... 414 References ............................................................................................................................. 416 16. Shock Physics ...................................................................................................................... 419 16.1 Shock Hugoniots ....................................................................................................... 419 16.2 Rarefaction Waves ..................................................................................................... 437 16.3 Stress Waves in Solids .............................................................................................. 460 16.4 Detonation Physics....................................................................................................480 16.5 Explosives Equations of State .................................................................................. 499 16.5.1 JWL Equation of State .................................................................................500 16.5.2 JWLB Equation of State ............................................................................... 501 16.5.3 Analytic Cylinder Model ............................................................................ 501 References .............................................................................................................................505 Further Readings ................................................................................................................. 506 17. Introduction to Explosive Effects .................................................................................... 507 17.1 Gurney Method ......................................................................................................... 507 17.2 Taylor Angles ............................................................................................................. 511 17.3 Mott Formula ............................................................................................................. 516 References ............................................................................................................................. 522 Further Reading ................................................................................................................... 522 18. Shaped Charges .................................................................................................................. 523 18.1 Shaped Charge Jet Formation ................................................................................. 525 18.2 Shaped Charge Jet Penetration................................................................................534 References .............................................................................................................................546 Further Reading ...................................................................................................................546 © 2014 by Taylor & Francis Group, LLC x Contents 19. Wound Ballistics ................................................................................................................. 547 References ............................................................................................................................. 557 Further Reading ................................................................................................................... 557 Appendix A ................................................................................................................................. 559 Appendix B.................................................................................................................................. 569 Index ............................................................................................................................................. 575 © 2014 by Taylor & Francis Group, LLC Preface to the Second Edition When we were asked to create a second edition to Ballistics: Theory and Design of Guns and Ammunition, we were a bit concerned about what could or should be added or changed. One thing that was certain was that the numerous errors that we became aware of through teaching and student observations had to be corrected. For that, we are indebted to the students and readers who have taken the time and effort to point these out. It was certainly a humbling experience. Hopefully we got them all this time. With the errors corrected, the task of deciding what we could put in to add more value to the book while maintaining an introductory treatment rose to the forefront. A great deal of work went into this edition that might not be apparent on the surface. We added a large number of new problems and updated the solutions manual accordingly. In fact, the solutions manual increased in size from 247 pages to 543 pages! One might notice that the new problems do not have answers in the text. This was intentionally done to allow instructors to assign these problems as tests or graded homework. We also added some Mathcad® codes to the CRC website to assist our readers with solution of the problems. These are available at http://www.crcpress.com/product/isbn/9781466564374. The more obvious changes include an increase in topical areas. We decided to add a section to Chapter 8 that would, in a general sense, discuss the topic of probability of first round hit for direct fire weapons. This treatment is general and certainly does not even attempt to reflect the state of the art since that data are not appropriate for a general audience. Another area of major improvement is the additional sections added to Chapter 16, which discuss explosive equations of state. Dr. Ernie Baker, our professional colleague and recognized expert in the field of explosives effects and insensitive munition design, added this section. We are extremely grateful to him for this. His schedule was incredibly busy, yet he found the time to assist us. Finally, there was a significant update to Chapter 19 on wound ballistics. During any course on terminal ballistics, it seems that this topic area is the one people seem most interested in. Granted, the treatment is still general, and, as is common in our experience, any statement made about the wounding process will be hotly debated, we decided to put more flesh on the bones of this section. Hopefully it is of help, keeping in mind the introductory nature of the treatment. Michael S. L. Hollis is the creator of the cover art, which depicts a typical kinetic energy penetrator in the midst of discarding its sabot. The sabot consists of the three “petals,” each with bow shocks forming ahead of them, peeling away from the subprojectile. The flames emanating from the rear of the subprojectile represent the tracer that allows the firer to visualize its trajectory downrange. These projectiles are generally fired at high Mach numbers and are used in weapons with bore diameters that range from 20–120 mm. We are indebted to Jonathan Plant of CRC Press, Taylor & Francis Group, for encouraging (i.e., gently prodding) us to undertake the work of revision and facilitating its publication. xi © 2014 by Taylor & Francis Group, LLC xii Preface to the Second Edition Disclaimer The design, fabrication, and use of guns, ammunition, and explosives are, by their very nature, dangerous. The techniques, theories, and procedures developed in this book should not be utilized by anyone without the proper training and certifications. In the checking and editing of these techniques, theories, and procedures, every effort has been made to identify potential hazardous steps, and safety precautions have been inserted where appropriate. However, these techniques, theories, and procedures must be exercised at one’s own risk. The authors and the publisher, its subsidiaries and distributors, assume no liability and make no guarantees or warranties, express or implied, for the accuracy of the contents of this book or the use of information, methods, or products described within. In no event shall the authors, the publisher, its subsidiaries or distributors be liable for any damages and expense resulting from the use of information, methods, or products described in this book. MATLAB® is a registered trademark of The MathWorks, Inc. For product information, please contact: The MathWorks, Inc. 3 Apple Hill Drive Natick, MA 01760-2098 USA Tel: 508-647-7000 Fax: 508-647-7001 E-mail: info@mathworks.com Web: www.mathworks.com © 2014 by Taylor & Francis Group, LLC Preface to the First Edition This book is an outgrowth of a graduate course taught by the authors for the Stevens Institute of Technology at the Picatinny Arsenal in New Jersey. Engineers and scientists at the arsenal have long felt the need for an armature of the basic physics, chemistry, electronics, and practice on which to flesh out their design tasks as they go about fulfilling the needs and requirements of the military services for armaments. The Stevens Institute has had a close association with the arsenal for several decades, providing graduate programs and advanced degrees to many of the engineers and scientists employed there. It is intended that this book be used as a text for future courses and as a reference work in the day-to-day business of weapons development. Ballistics as a human endeavor has a very long history. From the earliest developments of gunpowder in China more than a millennium ago, there has been an intense need felt by weapon developers to know how and why a gun works, how to predict its output in terms of the velocity and range of the projectiles it launched, how best to design these projectiles to survive the launch, fly to the target and perform the functions of lethality, and the destructions intended. The discipline over the centuries has divided itself into three natural regimes: Interior ballistics or what happens when the propellant is ignited behind the projectile until the surprisingly short time later when the projectile emerges from the gun; exterior ballistics or what happens to the projectile after it emerges and flies to the target and how to get it to fly there reproducibly shot after shot; and terminal ballistics or once it is in the vicinity of the target, how to extract the performance from the projectile for which the entire process was intended, usually lethality or destruction. Ballisticians, those deeply involved in the science of ballistics, tend to specialize in only one of the regimes. Gun and projectile designers, however, must become proficient in all the regimes if they are to successfully field weapons that satisfy the military needs and requirements. The plan of this book is bilateral: first, an unfolding of the theory of each regime in a graduated ascent of complexity, so that a novice engineer gets an early feeling for the subject and its nomenclature and is then brought into a deeper understanding of the material; second, an explanation of the design practice in each regime. Most knowledge of weapon design has been transmitted by a type of apprenticeship with experienced designers sharing their learning with newer engineers. It is for these engineers that this work is intended, with the hope that it will make their jobs easier and their designs superior. xiii © 2014 by Taylor & Francis Group, LLC © 2014 by Taylor & Francis Group, LLC Acknowledgments In the four years of development of this book, many people have contributed in a variety of ways. This has probably been the most difficult section that we had to write for fear of neglecting key people. We are a generation apart in our careers, mentors, and experience, but not in our enthusiasm for the subject. As the elder, Sidney S. Jacobson must acknowledge several people who inspired his enthusiasm beginning in the 1950s: Robert Schwartz, designer of the 280 mm atomic shell who also wrote the first set of ammunition design notes; Alfred A. Loeb, aeroballistician and friend; and Ralph F. Campoli, a peerless ammunition designer and longtime friend. Donald E. Carlucci would particularly like to thank for encouragement and mentorship Michael P. Devine, William DeMassi, James Pritchard, Howard Brunvoll, Vincent Marchese, Robert Reisman, Dr. Daniel Pillasch, Donald Rybarczyk, Dale Kompelien, Anthony Fabiano, Carmine Spinelli, Stephen Pearcy, Ami Frydman, Walter Koenig, Dr. Peter Plostins, and William R. Smith, all of whom have contributed to his professional development and thus to the ultimate publication of this book. Both of us must acknowledge as mentor, role model, and friend, the late Victor Lindner, an acknowledged leader in the world of weaponry, whose career spanned both of ours. On the inspirational as well as technical side of the ledger, our heartfelt thanks go to Paul Cooper and Dr. John Zukas, whose gentle prodding to get the project moving and encouragement throughout has been unflagging. From a technical standpoint, we would like to thank Dan Pangburn for allowing us to incorporate his method of buttress thread calculation, Dr. Bryan Cheeseman for his comments and help on the composites and ceramics sections, all of the reviewers of the book, Dr. Costas Chassapis, and Dr. Siva Thangam for their support while the material was being developed and taught as courses for Stevens Institute of Technology, Mark Minisi, Stanley DeFisher, Shawn Spickert-Fulton, Miroslav Tesla, Patricia Van Dyke, Dr. Wei-Jen Su, Yin Chen, John Thomas, Dr. Bill Drysdale, Dr. Bill Walters, Igbal Mehmedagic, and Julio Vega who, as teachers, students, friends, and co-workers, have either contributed analyses, checked problems, or suggested corrections to the manuscript. We would like to thank the following persons who very kindly assisted in pointing out errors in the text: Dr. Ernie Baker, James Boatright, Sean Brandt, Liam Buckley, Dominic Cimorelli, Paul Cooper, Michael Colonnello, Charles T. Freund, John Geaney, Nicholas Grossman, Jonathan Jablonski, Dennis Mackin, Mark Minisi, N. Rus Payne, Stephen Recchia, Lindsay Roberts, Oscar Ruiz, Shawn Spickert-Fulton, Robert Terhune and Caitlin Weaver. Additionally, we would like to thank Jonathan Plant and Richard Tressider for their help and encouragement, patience, and commitment to the publication of both the first and second editions of the book. With regard to the first edition, we wish to thank Sathyanarayanamoorthy Sridharan for actually seeing the edition through publication. Similarly, we wish to thank Arun Kumar Aranganathan for seeing this second edition through publication. Only we are responsible for any errors of commission or omission. xv © 2014 by Taylor & Francis Group, LLC © 2014 by Taylor & Francis Group, LLC Authors Donald E. Carlucci has been an engineer at the U.S. Army Armament, Research, Development and Engineering Center, Picatinny Arsenal, since May 1989. He is currently the U.S. Army Senior Scientist for Computational Structural Modeling based at Picatinny. He was formerly chief of the Analysis and Evaluation Technology Division, Fuze and Precision Munitions Technology Directorate responsible for the modeling and evaluation of cannon-launched munitions programs at Picatinny, and chief scientist for the XM982 Excalibur guided projectile. Dr. Carlucci also has formerly held the position of development program officer (chief engineer) for sense and destroy armor (SADARM). Before his employment at Picatinny, he was a design engineer for Titanium Industries, located in Fairfield, New Jersey. Dr. Carlucci has held positions as chief engineer, quality assurance manager, and purchasing manager for Hoyt Corporation, located in Englewood, New Jersey. He is a licensed professional engineer in the states of New Jersey and New York and holds a doctor of philosophy in mechanical engineering (2002) and a master of engineering (mechanical) (1995) degree from the Stevens Institute of Technology, Hoboken, New Jersey. In 1987, he received his bachelor of science degree in mechanical engineering from the New Jersey Institute of Technology, Newark, New Jersey. Dr. Carlucci is an adjunct professor of mechanical engineering at the Stevens Institute of Technology where he teaches graduate classes on interior, exterior, and terminal ballistics as well as undergraduate classes on engineering design. Sidney S. Jacobson was a researcher, designer, and developer of ammunition and weapons at the U.S. Army’s Picatinny Arsenal in New Jersey for 35 years. He rose from junior engineer through eight professional levels in research and development laboratories to become associate director for R&D at the arsenal. His specialty for most of his career was in the development of large caliber tank munitions and cannons. Many of these weapons, such as the long rod, kinetic energy penetrators (APFSDS rounds), and the shaped charge, cannon-fired munitions (HEAT rounds), have become standard equipment in the U.S. Army. For these efforts and successes he earned several awards from the army including, in 1983, the Department of the Army Meritorious Civilian Service Medal. In 1972, he was awarded an Arsenal Educational Fellowship to study continuum mechanics at Princeton University where he received his second MS degree (1974). He earned a master of science in applied mechanics from Stevens Institute of Technology (1958) and a bachelor of arts in mathematics from Brooklyn College (1951). He retired in 1986 but maintains his interest in the field through teaching, consulting, and lecturing. He holds two patents and was a licensed professional engineer in New Jersey. xvii © 2014 by Taylor & Francis Group, LLC © 2014 by Taylor & Francis Group, LLC Part I Interior Ballistics © 2014 by Taylor & Francis Group, LLC © 2014 by Taylor & Francis Group, LLC 1 Introductory Concepts The subject of ballistics has been studied for centuries by people at every level of academic achievement. Some of the world’s greatest mathematicians and physicists such as Newton, Lagrange, Bernoulli, and others solved problems in mathematics and mechanics that either directly or indirectly were applied to the various ballistic disciplines. At the other end of the academic scale, there are individuals such as James Paris Lee (inventor of the Lee-Enfield rifle) who developed his first weapon (not the famous Lee-Enfield) at age 12 with no formal education. The dominant characteristic of any of the ballistic disciplines is the “push–pull” relationship of experiment and analysis. It is a rare event, even as of this writing, when an individual can design a ballistic component or device, either digitally or on paper, and have it function “as designed” in the field. Some form of testing is always required and consequent tweaking of the design. This inseparable linkage between design and test is due to three things: the stochastic nature of ballistic events, the infinite number of conditions into which a gun–projectile–charge combination can be introduced, and the lack of understanding of the phenomena. The stochastic behavior that dominates all of the ballistic disciplines stems from the tremendous number of parameters that affect muzzle velocity, initial yaw, flight behavior, etc. These parameters can be as basic as how or when the propellant was produced to what was the actual diameter of the projectile measured to 0.0001 in. Even though, individually, we believe that we understand the effect of each parameter, when all parameters are brought together the problem becomes intractable. Because of this parameter overload condition, the behavior is assumed to be stochastic. The number of battlefield and test conditions that a gun–projectile–charge combination can be subjected to is truly infinite. For safety and performance estimates, the U.S. Army is often criticized for demanding test conditions, which could not possibly occur. While this may be true, it is simply a means of over-testing a design to assure that the weapon system is safe and reliable when the time comes to use it. This philosophy stems from the fact that you cannot test every condition and also because soldiers are an ingenious bunch and will invent new ways to employ a system beyond its design envelope. Lack of understanding of the phenomena may seem rather strong wording even though there are instances where this is literally true. In most cases, we know that parameters are present that affect the design. We also know how they should affect the design. Some of these parameters cannot be tested because there is some other, more fundamental variable that affects the test setup to a far greater degree. The overall effect of ballistic uncertainty, as described earlier, is that it will be very unusual for you to see the words “always” or “never” when describing ballistic phenomena in this work. 3 © 2014 by Taylor & Francis Group, LLC 4 Ballistics: Theory and Design of Guns and Ammunition 1.1 Ballistic Disciplines The field of ballistics can be broadly classified into three major disciplines: interior ballistics, exterior ballistics, and terminal ballistics. In some instances, a fourth category named intermediate ballistics has been used. Interior ballistics deals with the interaction of the gun, projectile, and propelling charge before emergence of the projectile from the muzzle of the gun. This category would include the ignition process of the propellant, the burning of propellant in the chamber, pressurization of the chamber, the first-motion event of the projectile, engraving of any rotating band and obturation of the chamber, in-bore dynamics of the projectile, and tube dynamics during the firing cycle. Intermediate ballistics is sometimes lumped together with interior ballistics, but has come into its own category of late. Intermediate ballistics deals with the initial motion of the projectile as it is exiting the muzzle of the tube. This generally includes initial tip-off, tube and projectile jump, muzzle device effects (such as flash suppression and muzzle brake venting), and sabot discard. Exterior ballistics encompasses the period from when the projectile has left the muzzle until impact with the target. One can see the overlap here with intermediate ballistics. In general, all that the exterior ballistician is required to know is the muzzle velocity and tip-off and spin rates from the interior ballistician, and the physical properties (shape and mass distribution) from the projectile designer. In exterior ballistics, one generally is concerned with projectile dynamics and stability, the predicted flight path and time of flight, and angle, velocity and location of impact. More often, now than in previous years, the exterior ballistician (usually called an aero-ballistician) is also responsible for designing or analyzing guidance algorithms carried onboard the projectiles. Terminal ballistics covers all aspects of events that occur when the projectile reaches the target. This means penetration mechanics, behind armor effects, fragment spray patterns and associated lethality, blast overpressure, nonlethal effects, and effects on living tissue. This last topic is becoming more and more important because of the great interest in lessthan-lethal armaments and, indeed, it has been categorized into its own discipline known as wound ballistics. 1.2 Terminology Throughout this work we will be using the word “gun” in its generic sense. A gun can be loosely defined as a one-stroke internal combustion engine. In this case, the projectile is the piston and the propellant is the air–fuel mixture. Guns themselves can be classified in four broad categories: a “true” gun, a howitzer, a mortar, and a recoilless rifle. A true gun is a direct-fire weapon that predominantly fires a projectile along a relatively flat trajectory. Later on we will decide what is truly flat and what is not. Notice the word “predominantly” crept in here. A gun, say on a battleship, can fire at a high trajectory sometimes. It is just usually used in the direct-fire mode. A gun can be further classified as rifled or smooth bore, depending upon its primary ammunition. Guns exhibit a relatively high muzzle velocity commensurate with their direct-fire mission. Examples of guns include tank cannon, machine guns, and rifles. © 2014 by Taylor & Francis Group, LLC Introductory Concepts 5 A howitzer is an indirect-fire weapon that predominantly fires projectiles along a curved trajectory in an attempt to obtain improved lethal effects at well-emplaced targets. Again, howitzers can and have been used in a direct-fire role; it is simply not one at which they normally excel. A mortar is a tube that is usually man-portable used to fire at extremely high trajectories to provide direct and indirect support to the infantry. Mortars generally have much shorter ranges than howitzers and cannot fire a flat trajectory at all. A recoilless rifle is a gun designed with very little weight. They are usually mounted on light vehicles or man emplaced. They are used where there is insufficient mass to counteract the recoil forces of a projectile firing. This is accomplished by venting the high-pressure gas out of a rear nozzle in the breech of the weapon in such a way as to counter the normal recoil force. A large listing of terminology unique to the field of ballistics is included in the glossary in Appendix A. 1.3 Units and Symbols The equations included in the text may be used with any system of units. That being said, one must be careful of the units chosen. The literature that encompasses the ballistic field uses every possible system and is very confusing for the initiate engineer. The U.S. practice of mixing the International System of Units (SI), United States Customary System (USCS), and Centimeter–Gram–Seconds (CGS) units is extremely challenging for even the most seasoned veteran of these calculations. Because of this an emphasis has been placed on the units in the worked-out examples and cautions are placed liberally in the text. Intensive and extensive properties (where applicable) are denoted by lowercase and uppercase symbols, respectively. In some instances, it is required to use the intensive properties on a molar basis. These will be denoted by an overscore tilde. In all cases, the reader is advised to always be sure of the units. © 2014 by Taylor & Francis Group, LLC © 2014 by Taylor & Francis Group, LLC 2 Physical Foundation of Interior Ballistics 2.1 Ideal Gas Law The fundamental means of exchanging the stored chemical energy of a propellant into the kinetic energy of the projectile is through the generation of gas and the accompanying pressure rise. We shall proceed in a disciplined approach, whereby we introduce concepts at their simplest level and then add the complications associated with the real world. Every material exists in some physical state of either solid, liquid, or gas. There are several variables that we can directly measure and some that we cannot but that are related to one another through some functional relationship. This functional relationship varies from substance to substance and is known as an equation of state. Thermodynamically, the number of independent properties required to define the state of a substance is given by the so-called state postulate, which is described in Ref. [1]. For all of the substances examined in this text, we shall assume they behave in a simple manner. This essentially means that the equilibrium state of all of our substances can be defined by specification of two independent, intrinsic properties. In this sense, an intrinsic property is a property that is characteristic of (in other words, governed by) molecular behavior. The ideal gas law is essentially a combination of three relationships [2]. Charles’s law states that volume of a gas is directly proportional to its temperature. Avogadro’s principle states that the volume of a gas is directly proportional to the number of moles of gas present. Boyle’s law states that volume is inversely proportional to pressure. If we combine these three relationships, we arrive at the famous ideal gas law, which states in extensive form: pvɶ = NℜT (2.1) where p is the pressure of the gas ṽ is the molar specific volume N is the number of moles of the gas ℜ is the universal gas constant T is the absolute temperature The units of Equation 2.1 are not always convenient to work with. For this reason, the form of the ideal gas law that we shall use most often in this text is pv = RT (2.2) 7 © 2014 by Taylor & Francis Group, LLC 8 Ballistics: Theory and Design of Guns and Ammunition In this case p is the pressure of the gas v is the specific volume (in mass units as we are used to) R is the specific gas constant, unique to each gas T is again the absolute temperature The specific gas constant can be determined from the universal gas constant by dividing the latter by the molar mass: R= ℜ M (2.3) where M is the molar mass of the gas (e.g., 15.994 lbm/lb-mol for oxygen). There are many other variants of the ideal gas law, which differ only in units. The other two versions that we occasionally utilize are pV = mg RT (2.4) p = ρ RT (2.5) and In these equations V (nonitalicized) is the volume the gas occupies mg is the mass of the gas ρ is the gas density One should always check units when using these equations. The pressure in a vessel filled with gas is caused by innumerable collisions of the gas molecules on the walls of the vessel [2]. The more tightly packed the molecules are, the more collisions occur—the higher the pressure is. Similarly, temperature excites the gas molecules so that they move faster, collide more—thus also increasing pressure. It is these collisions, among other things, that must be handled somehow by our equation of state. The ideal gas law relies upon the fact that the gas molecules are very far apart relative to one another [3]. If the molecules linger in the neighborhood of one another they will be influenced by strong intermolecular forces, which can either attract or repel them from one another. Thus, the ideal gas law ignores this effect. The ideal gas law further assumes intermolecular collisions occur completely elastically (i.e., like billiard balls). These assumptions must be kept in mind when using the ideal gas law. We shall soon see that under the pressures and temperatures in a gun that these assumptions are invalid. Nevertheless, they provide us with a point of departure and a useful stepping-stone for our studies. To use the ideal gas law to determine the state of the gas in a gun, we need to invoke classic thermodynamic relationships. The second law of thermodynamics can be stated as follows: Q = ∆U + W + losses (2.6) where Q is the energy added to the system ∆U is the change in internal energy W is the work done on the system, and the losses term contains all of the energy that cannot be recovered if, say, we pushed the projectile back to its starting position in the gun tube © 2014 by Taylor & Francis Group, LLC Physical Foundation of Interior Ballistics 9 Our sign convention shall be that a Q will be positive when energy is added to the system, ∆U will be positive if the internal energy of the system is increased, and W will be positive if work is done on the system. Losses always remove energy from the system. If we tailor Equation 2.6 to a gun launch situation, then Q would be the energy released by burning our propellant, ∆U would be the change in internal energy of the propellant, and W would be the work done on the projectile. Let us further define the work term in the classical sense. It is typical of a first year engineering curriculum to define the work as follows: ∫ W = F ⋅ dx (2.7) In Equation 2.7, work is defined as a scalar that results from the vector dot product of force, F with the distance over which the force acts, also a vector, dx (note that all vectors are characterized by bold type in this book). If we restrict our analysis to a gun system, we can see that, given pressure acting on the base of a projectile, it only has one direction to travel due to the constraints of the gun tube. If we imagine that this gun tube is perfectly straight (it never is) and we align a coordinate system with the axis of the tube, then the displacement vector, dx, must be aligned with force vector, F (i.e., the angle between F and dx is zero therefore the cosine of the angle between them is unity); and our relation for a dot, or more formally, the scalar product of these two vectors gives us F ⋅ dx = F ⋅ dx ⋅ cos(0) = Fdx (2.8) Our work definition for this case is then ∫ W = Fdx (2.9) This relationship for work has to be refined somewhat to fulfill our needs. We will need to put the force acting on the projectile in terms of the pressure and sometimes would like the volume to be included in the equation. If we look at the ideal gas equation of state in the form of Equation 2.4, we do not see a force in there but we do see a pressure term and a volume term. We know from the mechanics of materials [4] that F = pA (2.10) This has not been written in vector form so as to keep things simple (we will write it differently later). Equation 2.10 states that the resultant force, F, on a body is equal to the average pressure, p, on that body times the area, A, over which the pressure acts. So we can rewrite Equation 2.9 using this result as W= ∫ pA dx (2.11) We now need to get volume in there somehow. We shall use the fact that, except for the chamber of a gun (and a few notable exceptions with the bore), the area over which the pressure acts is constant and equal to the bore cross-sectional area that we have defined © 2014 by Taylor & Francis Group, LLC Ballistics: Theory and Design of Guns and Ammunition 10 as A earlier. The area of the rifling grooves does contribute here if the tube is rifled, but let us assume a nice smooth cylindrical bore for now. If A is the cross-sectional area and dx is a differential element of length, then the differential element of volume, dV, can be defined as dV = Adx (2.12) We can now write Equation 2.11 in terms of pressure and volume as W= ∫ p dV (2.13) You may recall this form of the definition of work from thermodynamics [5]. We now have two equations and a definition at our disposal as a pedagogical device that can help illustrate the energy exchange mechanism in a gun. The equations are an ideal gas equation of state Equation 2.4 and the second law of thermodynamics, Equation 2.6; and the definition of how we defined work in Equation 2.13. Let us imagine that we have a simple gun as depicted in Figure 2.1. We shall assume that we have somehow placed a mass, mg, of a gas that behaves according to the ideal gas equation of state in the tube and compressed it, adiabatically, using the projectile and no leakage has occurred. We shall further assume that there is no friction between the projectile and the tube wall. Thus, in the situation depicted by Figure 2.1, we have an ideal gas trapped between the projectile and the breech, compressed to some pressure, p, at some absolute temperature, T. We shall further assume that the projectile of mass, mp, is somehow held at position x = 0 and no gas or energy can escape. In this situation, the volume the gas occupies, which we shall call the chamber volume, Vc, is given by Vc = πd 2 l 4 (2.14) What we have done essentially is compressed the projectile against an imaginary spring (the gas), which now has a potential energy associated with it. From a thermodynamic standpoint, we can reduce Equation 2.6 to 0 = ∆U + W (2.15) Recapping, we note that Q = 0 because there was no heat lost through the tube wall (adiabatic compression) and there is no propellant per se that will burn to generate heat. The losses were zero because we have no friction. mp l d mg x FIGURE 2.1 Simple gun system. © 2014 by Taylor & Francis Group, LLC L Physical Foundation of Interior Ballistics 11 Now that everything is set, we need to release our projectile and see what happens. If we substitute Equation 2.4 into Equation 2.13, we can write ∫ W = mg RT dV V (2.16) This equation now shows how much work is being done on the projectile as a function of the volume. It is noteworthy here that we are assuming the gas that is actually pushing on the projectile is massless. By this we mean that no energy is being applied to accelerate the mass of the gas. We will remove this assumption later in our studies. What we do not like about Equation 2.16 is that temperature still appears as a variable. By our earlier assumptions, we stated that the process was frictionless and adiabatic. Recall, again from thermodynamics, that this actually defines an isentropic process [1]. For a closed system (one with constant mass), it can be shown [6] that the absolute temperature, T, of our system is related to the initial temperature of the gas, Ti, through V  T = Ti  c  V (γ −1) (2.17) where V is the volume at a given time t Vc is the initial chamber volume γ is the specific heat ratio of the gas (defined later) If we substitute Equation 2.17 into Equation 2.16, we can write V (γ −1) c W = mg RTi V ∫V −γ dV (2.18) Vc This equation is easy to work with because we know most of the terms on the RHS (righthand side) when we set up our pedagogical gun. We know the mass, mg, of the gas. We know R and γ because we picked which gas it was. We know the initial temperature of the gas and we know the chamber volume. Now that we did all of this work with volumes, we want to convert these back to distances. A typical output desired by ballisticians is the pressure versus travel (i.e., distance) curve. This plot helps the gun designer determine where to make his tube thick and where he can get away with thinning the wall. If we again recognize that our gun has a constant inner diameter, we can use Equation 2.14 to write Equation 2.18 as L W = mg RTil (γ −1) ∫ (l + x ) −γ dx (2.19) 0 If we perform this integration, we obtain W= © 2014 by Taylor & Francis Group, LLC mg RTil(γ −1) (l + L)(1−γ ) − l(1−γ )  (1 − γ )  (2.20) Ballistics: Theory and Design of Guns and Ammunition 12 We need to recall from dynamics that the kinetic energy of the projectile can be written as K.E.projectile = 1 mpVm2 2 (2.21) where Vm is the muzzle velocity. If we assume that all of the energy of the gas is converted with no losses into kinetic energy of the projectile, then we can use Equation 2.15 to state that (2.22) K.E.projectile = W We can make use of Equations 2.20 and 2.21 to write this as mg RTil(γ −1) 1 [(l + L)(1−γ ) − l(1−γ ) ] mpVm2 = 2 (1 − γ ) (2.23) This is an important result as it relates muzzle velocity to the properties and amount of the gas used, the mass of the projectile, and includes the effect of tube length. We can use this equation to estimate muzzle velocity. So a convenient form of this equation is Vm = 2 mg RTil(γ −1) [(l + L)(1−γ ) − l(1−γ ) ] mp (1 − γ ) (2.24) In some instances, we would like to use these relationships to determine the state of the gas or velocity of the projectile at some point in the tube other than the muzzle. If this is the case, the procedure would be as follows: 1. Solve for the work term up to the position of interest, xproj, using xproj W (xproj ) = mg RTil (γ −1) ∫ (l + x ) −γ dx (2.25) 0 2. Determine the volume at the position of interest using V( xproj ) = πd 2 (l + xproj ) 4 (2.26) 3. Determine the gas temperature at this position from Equation 2.17 4. Determine pressure from the ideal gas Equation 2.4 This procedure is relatively straightforward. If, as an example, we look at an idealized 155 mm compressed air gun and assume the following parameters Projectile weight = 100 lbm Initial pressure = 45 MPa (approximately 6500 psi) Tube length = 6 m From Figures 2.2 through 2.4, we can depict the results of a calculation for temperature, pressure, and velocity versus distance for this idealized situation. © 2014 by Taylor & Francis Group, LLC Physical Foundation of Interior Ballistics 13 350 T (K) 300 Temperature (K) 250 200 150 100 50 0 0 1 2 3 4 Distance (m) 5 6 7 FIGURE 2.2 Temperature versus distance in an ideal gas gun. 50,000,000 p (Pa) 45,000,000 40,000,000 Pressure (Pa) 35,000,000 30,000,000 25,000,000 20,000,000 15,000,000 10,000,000 5,000,000 0 0 1 2 3 4 Distance (m) 5 6 7 FIGURE 2.3 Pressure versus distance in an ideal gas gun. Problem 1 Assume we have a quantity of 10 g of 11.1% nitrated nitrocellulose (C6H8N2O9) and it is heated to a temperature of 1000 K assuming it changes from solid to gas somehow without changing chemical composition. If the process takes place in an expulsion cup with a volume of 10 in.3, assuming ideal gas behavior, what will the final pressure be in pounds per square inch?  lbf  Answer: p = 292  2   in.  © 2014 by Taylor & Francis Group, LLC Ballistics: Theory and Design of Guns and Ammunition 14 250 Velocity (m/s) 200 150 100 50 V (m/s) 0 0 1 2 3 4 Distance (m) 5 6 7 FIGURE 2.4 Velocity versus distance in an ideal gas gun. 2.2 Other Gas Laws There are many times when ideal gas behavior is insufficient to model real gases. This is certainly true under the pressures and temperatures of gun launch. Although there are many models that attempt to account for the deviation of real gases from ideal or perfect behavior [2,3], we shall examine only two, the simplest of which we shall use. Ideal gas behavior is approached when the distance between molecules (known as the mean free path) is large. Thus, molecules do not collide or interact with one another very often. Temperature is a measure of the internal energy of the gas. Thus, when the temperature is high, the molecules are moving around faster and have more of an opportunity to interact with one another. Pressure is a result of how closely the molecules are packed together, thus a higher pressure tends to put the molecules in close proximity. It is for these reasons that we cannot normally use the ideal gas law in gun launch applications. The Noble–Abel equation of state is given by p( V − mg b) = mg RT (2.27) where p is the pressure of the gas V is the volume the gas occupies mg is the mass of the gas R is the specific gas constant T is the absolute temperature b is the co-volume of the gas The co-volume of the gas has been described as a parameter that takes into account the physical size of the molecules and any intermolecular forces created by their proximity © 2014 by Taylor & Francis Group, LLC Physical Foundation of Interior Ballistics 15 to one another. Think of it as not having physical meaning but as simply a number that allows for a better fit to observed experimental data. The units of the co-volume are cubic length per mass unit. Usually, the gas co-volume is provided in the literature but an estimation tool has been provided by Corner [7], which will not be repeated here since actual data exist. Occasionally, the Noble–Abel equation of state is insufficient to suit our needs. At these times, it is typical to use a Van der Waals equation of state given by p= ɶ RT a′ − vɶ − b′ vɶ2 (2.28) In this case p is again the pressure of the gas ṽ is the molar specific volume R̃ is the molar specific gas constant, unique to each gas T is again the absolute temperature a' and b' are constants particular to the gas The Noble–Abel equation of state is the basis for nearly all of our work in this text, therefore Equation 2.27 is very important. At times, we may write it a little differently but you will always be reminded of where it originated. Problem 2 Perform the same calculation as in Problem 1, but use the Noble–Abel equation of state and assume the co-volume to be 32.0 in.3/lbm  lbf  Answer: p = 314.2  2   in.  Problem 3 A hypothetical “air mortar” is to be made out of a tennis ball can using a tennis ball as the projectile. The can has a 2–1/2” inside diameter and is 8” long. If a tennis ball of the same diameter weighs 2 oz. and initially rests against the rear of the can, to what air pressure must one pressurize the can to in order to achieve a 30 ft/s launch of the tennis ball? Assume that the tennis ball can be held against this pressure until released, that it perfectly obturates and also assume an isentropic process and ideal gas behavior with γ = 1.4 for air. 2.3 Thermophysics and Thermochemistry The main energy exchange process of conventional interior ballistics is through combustion. Once ignited, the chemical energy of the propellant is released through an oxidation reaction. This energy release will be in the form of heat, which, in turn, increases the pressure in the volume behind the projectile (i.e., in a combustion chamber). The pressure exerts a force on the projectile, which accelerates it to the desired velocity. © 2014 by Taylor & Francis Group, LLC 16 Ballistics: Theory and Design of Guns and Ammunition In general, combustion requires three main ingredients to commence: a fuel, an oxygen source, and heat. In a common combustion reaction, such as an internal combustion engine like the one in your car, oxygen is supplied to the reaction independently of the fuel. The heat in this case is generated by a spark ignition and the burning of the air–fuel combination that ensues. A gun chamber has very little room for oxygen once it is stuffed with propellant. It is important to note that, for other reasons, there is always free volume in the chamber (called ullage)—we will explain this later. For now, we should understand that although there is some oxygen in the chamber, the amount is insufficient to completely combust the propellant. It is for this reason that propellants are formulated to contain both the fuel and the oxidizer. In general, the propellant burning is an under-oxidized reaction. This has some implications as the propellant gases leave the muzzle—again, we shall discuss this in more detail later. This brief introduction should make clear the reason to examine thermochemistry, thermophysics, and combustion phenomena. To proceed, we shall first define each field of study. The definitions of Ref. [8] shall be used here to describe the first two topics as they are extremely straightforward and clear. Thermophysics is defined as the quantification of changes in a substance’s energy state caused by changes in the physical state of the material. An example of this would be the determination of the amount of energy required to vaporize water in your teapot. Thermochemistry is then the quantification of changes in a substance’s energy state caused by changes in the chemical composition of the material’s molecules. An example of this would be the energy required to dissociate (break up) water molecules into hydrogen and oxygen. Combustion is defined in Ref. [1] as the quantification of the energy associated with oxidizer–fuel reactions. Thus, combustion is a natural outgrowth of thermophysics and thermochemistry. Now that we have categorized these three fields of study, we shall attack them in a somewhat jumbled order. The reason for this is that, from our perspective, we really need not distinguish between any of them and all of them appear in our gun launch physics. It is also important to realize that whether the energy change comes from a chemical reaction or a phase change from solid to gas, as long as we can calculate the extent of the energy change, we can perform a valuable analysis. Energy to all intents and purposes consists of two types: potential and kinetic. Potential energy can be considered as stored energy. There are many ways to store energy. We can store energy by compressing a steel bar or spring, by lifting a mass to a higher elevation in the earth’s gravitational field, and by chemically preparing a compound that, whether by combustion or chemical reaction, will release energy. Each of these forms of potential energy: elastic strain; gravitational potential and chemical potential energy, has a different method of storing and releasing the energy but they are all potential energies. There are other forms of potential energy but we need not deal with them in this context. Kinetic energy is the energy of a mass in motion. It can be observed in objects that are in translational motion or in rotational motion. To extract some or all of this energy, it is necessary to slow or stop the moving mass that has the kinetic energy. The energy in a spinning flywheel is an example of rotational kinetic energy. The field of thermodynamics is the study of energy transformations. It quantifies the balance of energy between kinetic and potential. In thermodynamics, it is common to see two energy transformation mechanisms: heat and work. Heat transfer is essentially an exchange of energy through molecular motion. As we shall soon see, molecules of a substance are always in motion. The faster they are in motion, the hotter the substance is. These molecules can influence other molecules when © 2014 by Taylor & Francis Group, LLC Physical Foundation of Interior Ballistics 17 they are placed in contact with them, thus giving up some of their energy and increasing the energy of the contacted substance. Temperature is a sensible measure of an object’s internal energy. Work is a means of increasing an object’s energy by application of a force through a distance. This method of energy transfer can create either potential energy, as in compressing a spring, or kinetic energy as applied to a free, rigid mass. While the equations for heat transfer can be the subject of entire texts (e.g., [9]), work can be defined through the vector equation W = F ⋅ dx (2.29) where W is the work done on or by the system F is the force vector dx is the vector distance through which the force acts, known as the displacement vector We must note that this is a vector equation. The work term is a scalar because the dot product of two vectors results in a scalar. Because of the dot product term, the sign of W is dependent upon the cosine of the angle between F and dx. Recall the definition of a dot product as A ⋅ B = AB cos θ (2.30) where A and B are the scalar magnitudes of the vectors A and B (Figure 2.5). If we use Equation 2.30 with the variables of Equation 2.29, this tells us that if the angle between the force vector and the displacement vector is between 0° and 90° or 270° and 0°, the work is positive, i.e., it is work performed on the system. If, however the angle is between 90° and 270°, the work is negative, and therefore work performed by the system. Internal energy, U, of a substance can be considered a form of potential energy. Some authors [5] categorize the internal energy separately from potential and kinetic energies. This can clearly be done in general, but for the application of gun launch it seems proper to group it as a potential energy. The internal energy of a substance is manifested in the molecular motions within that substance. These motions generally are translational or vibrational in nature. The molecules of a substance are attracted to and repelled by one another and are in some degree of translational motion. Additionally, the attractive or repulsive forces within a molecule itself allow us to use an analogy of springs holding the atoms together. Imagine a structure of a water molecule, for instance as depicted in Figure 2.6. If the oxygen and hydrogen atoms are assumed to be steel balls and the molecular bond springs, we could pick this molecule up, A θ B FIGURE 2.5 Depiction of two vectors for scalar product definition. © 2014 by Taylor & Francis Group, LLC 18 Ballistics: Theory and Design of Guns and Ammunition Oxygen atom Hydrogen atoms FIGURE 2.6 Model of a water molecule. hold the oxygen atom, and shake it. If the springs were really stiff in bending and much less so in tension or compression, we would see the hydrogen atoms oscillating in and out at some frequency. The greater the frequency, the more energy we would need to put into the system. Even though the springs are stiff in bending, it does not mean that they cannot bend. This just takes more energy. Like springs, we can store energy in the molecules this way. This simple model of a molecule is a crude but useful approximation. Imagine now that we put our model on a frictionless surface, like an ice hockey rink. If we hit the molecule in a random way, we will excite these vibrational modes as well as create translational and rotational motion. Now, if we fill the ice hockey rink with models … well, you get the idea. As stated previously, the level of this interaction (collisions) must be represented somehow. The metric used is internal energy with the level of activity defined as zero at the temperature known as absolute zero (0° on the Kelvin or Rankine scales). The internal energy also includes the energy required to maintain a particular phase of the material such as solid, liquid, or gas. Additionally, certain phases associated with molecular structure such as face-centered cubic (FCC), body-centered cubic (BCC), etc., are accounted for in the internal energy. Quite often we shall see internal energy and what is commonly known as “pdV” work terms together in our energy balance equations. The term is called pdV work because it is special and separate from work generated by, say, a paddle wheel moving fluid around. This work term arises from pressure pushing on a given volume. If the volume changes by an infinitesimal amount, dV, we essentially have force acting through a distance. To prove this to yourself, look at the units. Because we see these terms together so often, it is convenient for us to group them into one term, which we will call enthalpy, H. Mathematically, the enthalpy is defined as H = U + pV (2.31) Notice here that we have removed the differential from the work term. The reason for this is that, considering both enthalpy and internal energy, we are concerned with changes in H and U. Therefore, the differential appears when we write the entire equation in differential form as dH = dU + pdV (2.32) For proof of this result, refer to any thermodynamics text (e.g., [1,5]). An example of the difference between internal energy and enthalpy is the rigid container or piston container. Consider a rigid container that has some amount of gas in it. Assume the container is sealed © 2014 by Taylor & Francis Group, LLC Physical Foundation of Interior Ballistics 19 so that matter cannot enter or leave. Let us also assume that the container will allow energy to be transferred to and from the gas. If we transfer heat (energy) to the gas, the temperature will rise as will the pressure. Since the volume of the container is fixed, no work can be done; thus all of the energy added to the gas is internal energy. From Equation 2.32, we see that in this case the change in enthalpy would be exactly equal to the change of internal energy. Now we assume that, instead of our container being rigid, the roof of the container is a sealed yet moveable piston. In this case, once again matter cannot escape; however, the volume is able to change. Now the only thing holding up the roof is the pressure of the gas acting to just counteract the weight of the roof itself. Let us add the same amount of heat that we added to the original, rigid, container. In this case, the temperature of the gas will increase (but less than before) and the volume will increase because the piston is moveable and the pressure must remain constant and just sufficient to counteract the weight of the roof. In this instance, the enthalpy would be greater than the internal energy because it includes the work done in lifting the piston. When a substance changes form, chemically or physically, energy is either absorbed or released. The method that we use to quantify this energy change is through heats of formation and the like. Though called a “heat,” what is really implied is an enthalpy change. We shall proceed through these different enthalpy changes, attempting to list some of the more common ones. For greater detail, the reader is encouraged to consult thermodynamics texts in addition to the descriptions provided in Ref. [8]. Specific values for text problems will be given as needed. It is not the intent of the authors to tabulate the different energy parameters of different materials. When a substance is formed, atomic bonds in the constituent molecules are destroyed and then recreated (at least this is a clean way to think of it from a bookkeeping perspective). The energy absorbed or generated by this process is commonly called the heat of reaction, ∆H r0 . The Δ reminds us that we always are concerned with changes in enthalpy from a particular reference state (usually standardized as 25°C and 1 atm). The “0” superscript is a convenient reminder that this is from a reference state of 1 atm. As the subscript, sometimes we see “298” meaning 298 K. Though 298 K and 25°C are the same value, one must always be wary of the reference state chosen by a particular author. The heat of formation, ∆H f0, is the energy required to form a particular substance from its individual component atoms. The heats of formation are the building blocks that determine the heat of reaction. Any elemental substance in its stable configuration at standard conditions has a heat of formation equal to zero at that state. For instance, diatomic nitrogen, N2, has ∆H f0 = 0 at 25°C and 1 atm. We will provide an example of the heat of formation calculation in a later section. Now that with the aforementioned quantities defined, we can write an equation for the heat of reaction ∆H r0 = ∑ products ∆H f0 − ∑ ∆H f0 (2.33) reactants Equation 2.33 states that the heat of reaction for a given substance is equal to the sum of the heats of formation of the final products of the reaction that created the substance minus the sum of the heats of formation of the materials that had to be reacted together to create the new substance. This is further reinforcement of the definition of the heat of reaction. Recall that we stated the atomic bonds of the molecules were destroyed and then remade. This is essentially what Equation 2.33 is saying. The energy it took to create each of the reactants has to be accounted for and then the energy it takes to create the new substances from the © 2014 by Taylor & Francis Group, LLC Ballistics: Theory and Design of Guns and Ammunition 20 constituents is calculated—energy is conserved. If the heat of reaction is a negative number, heat is liberated by the reaction otherwise it is absorbed. When a compound is specifically combusted with sufficient oxygen to attain its most oxidized state, the heat of reaction has a special name: the heat of combustion. The heat of combustion is identified by the symbol ∆H c0. The heat of combustion is typically what is obtained when propellant is burned in a closed bomb. The equation for the heat of combustion mirrors that of the heat of reaction, the only difference being as noted earlier: ∑ ∆H c0 = ∆H f0 − fully oxidized products ∑ ∆H f0 (2.34) reactants The heats of detonation and explosion have meanings that seem to be reversed. The heat of detonation is the heat of reaction taken when detonation products are formed from an explosive compound during a detonation event. The formula for the heat of detonation is given by ∆Hd0 = ∑ ∑ ∆H f0 − detonation products ∆H f0 (2.35) original explosive What is termed the heat of explosion is the amount of energy released when a propellant or explosive is burned (not detonated) and is given by 0 ∆H exp = ∑ ∆H f0 − burning products ∑ ∆H f0 (2.36) original propellants The heat of afterburn is another type of heat of reaction that occurs often in propellants and explosives. Because the composition of propellants and explosives usually force an under-oxidized reaction, the reaction products will often combine with the oxygen present in the air outside the gun or explosive device, given sufficient temperature and pressure. This secondary reaction results in a second pressure wave or blast and a fireball. The heat of afterburn can be described mathematically as 0 ∆H AB = ∑ fully oxidized products ∆H c0 − ∑ ∆Hd0 (2.37) remaining detonation products Not all energy changes involve chemical reactions. We mentioned earlier that changes in physical state and structure require energy. When a solid melts to form a liquid or a liquid solidifies, we call the energy required, the latent heat of fusion, λf. These values are tabulated in any chemistry book or thermodynamics text. Some authors use different symbols so one must, as always, be careful. In a similar vein, the energy required to vaporize a liquid to a gas or condense a gas to a liquid is known as the latent heat of vaporization and given by the symbol λfg. If a material changes the structure of its atoms, say from BCC to FCC, the energy is known as the heat of transition, λt. There are many other types of material transitions that require energy. The types described earlier cover the needs of this work. © 2014 by Taylor & Francis Group, LLC Physical Foundation of Interior Ballistics 21 2.4 Thermodynamics The combustion process that occurs in a gun is a thermodynamic process. The term thermodynamics is a bit misleading because it implies that the dynamics of the combustion process is examined. This is not quite true. Classical thermodynamics is based on the examination of the various processes through equilibrium states. This is somewhat akin to frames of a motion picture. We examine the state of the system before some event and we usually examine it at some point, later in time, we are interested in. Some of the concepts of thermodynamics were introduced in earlier sections, work and energy being the major ones. Here we shall look in detail at two ways of describing thermodynamic systems to proceed with our study. We shall define energy for an arbitrary system as E=U+ 1 mV 2 + mgz 2 (2.38) Equation 2.38 is our extensive form of the definition of the system energy, E. In this equation, U is the internal energy, m is the system mass, V is the system velocity, g is a gravitational constant, and z is some height above a reference datum. The second and third terms on the RHS of the equation are the kinetic and potential energies, respectively. If we examine this equation, it is easy to see why some authors group the internal energy as a separate energy type. However, in the case of a gun launch, the potential energy term is insignificant. This focuses us on the transfer of energy between internal and kinetic. We sometimes write Equation 2.38 in its intensive form as 1 e = u + V 2 + gz 2 (2.39) Recall from our earlier discussions that an intensive property is the associated extensive property divided by mass. We shall now examine the first law of thermodynamics as it is applied to two different types of systems: a fixed mass of material and a fixed volume of space through which material flows. The first type of analysis, where the material is a fixed mass, is known as a Lagrangian approach, while the fixed or control volume (CV) approach is known as Eulerian. Both are important from a ballistic analysis standpoint and are prevalent in interior, exterior, and terminal ballistic studies. For a fixed mass of material, undergoing some thermodynamic process, the first law of thermodynamics can be written as Q1− 2 + W1− 2 = ∆E1− 2 (2.40) where Q is the heat or energy added to the system W is the work performed on or by the system ∆E is the change in the energy state of the material The subscript 1–2 simply lets us know that the process began at some state 1 and ends at some state 2. The signs on the terms are very important. We assume a positive change in © 2014 by Taylor & Francis Group, LLC Ballistics: Theory and Design of Guns and Ammunition 22 energy comes about through adding heat to the system and doing work on the system. Thus, work performed on the system is positive and heat added is also positive. Different thermodynamics texts write the first law slightly different, but if you understand that the net result of work on the system or heat transfer to the system is to increase its energy, then few mistakes will be made. An interesting observation of Equation 2.40 is that the energy state change has an infinite number of paths that lead to the same result. For instance, if we wanted to add 24 kJ of energy to some arbitrary system, we could do it by adding 12 kJ of heat and performing 12 kJ of work on the system. We could obtain the same result by adding 36 kJ of heat and extracting 12 kJ of work from the system. The possibilities are limitless. This reinforces our assertion that thermodynamics is really only concerned with end states. Caution is warranted at this point. Equation 2.40 does not say how the energy, once added to the system, is partitioned between potential (internal) energy or kinetic energy. This reveals something. Heat and work are added to or removed from the system at the system boundaries while the distribution of energy between internal or kinetic energy is done within the system. We shall now write out Equation 2.40 explicitly for a Lagrangian system  1   1  Q1− 2 + W1− 2 = m  u2 + V22  −  u1 + V12   2   2   (2.41) Here we have neglected the gravitational potential energy terms and used the intensive form of the energy, multiplied by the system mass. As previously stated, many times we would like to use enthalpies instead of internal energies. If this is the case, we can rewrite Equation 2.41 using our relationship between the two from Equation 2.40. We shall use the intensive form of Equation 2.40 to yield  1   1  Q1− 2 + W1− 2 = m  h2 − pv2 + V22  −  h1 − pv1 + V12   2   2   (2.42) Here we note that h is the specific enthalpy and v is the specific volume. We shall now examine the first law of thermodynamics in the Eulerian frame of reference. Recall that in the Eulerian frame, we chose a CV (real or imaginary) and observed how the energy within the volume changes based upon the energy carried into or out of it by any entering or exiting substance as well as any heat or work done at the system boundaries. It is convenient for us to write the first law in terms of the time rate of change of energy, heat, and work. We start by writing Equation 2.40 as a rate equation dQ dW dE + = dt dt dt (2.43) Qɺ + Wɺ = Eɺ (2.44) or Here the dots over the heat and work terms indicate the time rate of change of the variable. Proper thermodynamics terminology would require us to use the “δ” instead of “d” in © 2014 by Taylor & Francis Group, LLC Physical Foundation of Interior Ballistics 23 Equation 2.43 because of path dependency considerations, but for our purposes we shall ignore this fact. The reader is advised to consult any thermodynamics text for a better understanding of the difference. The substitutions that were performed to arrive at Equation 2.41 are not as straightforward in this case. Because we have material entering and leaving the CV, we can imagine that this material can enter or leave with a different pressure and density as it interacts with our fixed CV. Because of this, we must account for the energy used to make these changes. Alternatively, one can envision the material coming in at a higher pressure or density and wanting to push our imaginary CV outward, but since we fixed our CV it cannot. The energy from this must go somewhere so it works on the fluid in and around our CV. Mathematically, this results in the energy term in Equation 2.44 having to include a pv term. This is sometimes known as flow work [10]. With this in mind, Equation 2.44 can be written as ɺ out (eout + pout vout ) − m ɺ in (ein + pin vin ) Qɺ + Wɺ = m (2.45) . Here, by multiplying the intensive properties by the mass flow rate, m, we have the rate of change of the energy terms. We have also arbitrarily assumed one inlet and one outlet. If more inlets or outlets in our CV were present and they had different mass flow rates or pressures, we would have to consider each with a term identical to our outlet or inlet terms given earlier. We now can make the substitution for our energy terms to yield 1 2 1 2  ɺ   ɺ out  uout + Vout Qɺ + Wɺ = m + pout vout  − m Vout + pinvin  in  uin + 2 2     (2.46) In this case, we have also assumed a uniform velocity over the inlets and outlets. With one inlet and outlet, the mass flow in must equal the mass flow out so we can write Equation 2.46 as  1 2 1 2    ɺ  uout + Vout + pinvin   Qɺ + Wɺ = m + pout vout  −  uin + Vout 2 2     (2.47) Substitution of enthalpy into the aforementioned equation puts it into a compact form:  1 2   1 2  ɺ  hout + Vout Qɺ + Wɺ = m  −  hin + Vout   2 2     (2.48) In many fluid dynamics texts, there are wonderful examples of how these equations are used with multiple inlets and outlets [11]. You may be asking yourself how useful are these equations if we only use one inlet or outlet? The answer is that they are very useful. Except for flows through muzzle devices or through internal ports like bore evacuators and ports for automatic weapons, a gun is a right circular tube that contains the propellant gas. Any flow field analysis we perform on the moving gases will have just one inlet (toward the breech) and one outlet (toward the projectile). Thus, as we develop our equations later for in-bore motion, we can use these simple equations in the aforementioned form. © 2014 by Taylor & Francis Group, LLC Ballistics: Theory and Design of Guns and Ammunition 24 As a review, we have two equations that state the first law of thermodynamics. For a fixed mass of material (Lagrangian frame), we have  1   1  Q1− 2 + W1− 2 = m  h2 − pv2 + V22  −  h1 − pv1 + V12   2 2     (2.42) and for a fixed volume that material can flow in and out of (Eulerian frame)  1 2   1 2  ɺ  hout + Vout Qɺ + Wɺ = m  −  hin + Vout   2 2     (2.48) These equations have been repeated here because of their critical importance to our work. In many instances, we will find that we require a relationship between internal energy or enthalpy and temperature. If we have a gas that is not reacting and intermolecular forces are small enough to ignore, we can consider the gas to be thermally perfect [12]. The implications of this are that internal energy and enthalpy are functions of the temperature alone. With this model, we can write expressions for internal energy and enthalpy as follows: du = cvdT (2.49) dh = cpdT (2.50) where cv is the specific heat at constant volume cp is the specific heat at constant pressure Normally, cp and cv vary with temperature. In many practical cases, this variation is small and we can further assume that the gas is calorically perfect, which results in the aforementioned equations being written as u = c vT (2.51) h = cpT (2.52) For a thermally or calorically perfect gas (not a reacting gas), there is a relationship between cp, cv, and R. If we define γ as the ratio of specific heats where γ = cp cv (2.53) then we can write the aforementioned relationships as © 2014 by Taylor & Francis Group, LLC cp − c v = R (2.54) cp = γR γ −1 (2.55) cv = R γ −1 (2.56) Physical Foundation of Interior Ballistics 25 The second law of thermodynamics defines the concept of entropy for us [1]. We know from the second law of thermodynamics that Tds = du + pdv (2.57) or, if we insert the definition of enthalpy Tds = dh − vdp (2.58) If we evaluate Equations 2.57 and 2.58 under the assumptions of a calorically perfect gas, we obtain p  T  s2 − s1 = cp ln  2  − R ln  2   T1   p1  (2.59) T  v  s2 − s1 = cv ln  2  + R ln  2  T  1  v1  (2.60) In these expressions, the subscripts “1” and “2” indicate the initial and final states of the substance, respectively. An isentropic process is a process in which there is no entropy change. This is also known as a reversible process. In a real system, entropy must always increase or, at best, stay constant. Many processes have slight enough entropy increases as to be considered isentropic. Isentropic processes also are excellent to examine as theoretical limits on real processes. If we examine Equations 2.59 and 2.60 under an isentropic assumption, we see that the left-hand side (LHS) is zero in both. This has implications that allow us to write (for an isentropic process) γ p2  ρ 2   v2  =  = p1  ρ1   v1  −γ γ  T  γ −1 = 2   T1  (2.61) Problem 4 The M898 SADARM projectile weighs 102.5 lb. The projectile was fired from a 56 caliber, 155 mm weapon and a pressure–time trace was obtained. The area under the pressure–time curve was (after converting the time to distance) calculated to be 231,482 psi-m. Calculate the muzzle energy of the projectile in megajoules. Assume the bore area to be 29.83 in.2 Answer: E = 30.7 [MJ]. Problem 5 An 8 in. Mk. 14 Mod. 2 Navy cannon is used at NSWC Dahlgren, VA for “canister” firings. These firings are used to gun harden electronics that are carried in an 8 in. projectile. The projectile used weighs 260 lb. The measured muzzle velocity is around 2800 ft/s. Calculate the muzzle energy of the projectile in megajoules. Assume the bore area to be 51.53 in.2 The rifled length of the tube (distance of projectile travel) is 373.65 in. Answer: E ≈ 43 [MJ]. © 2014 by Taylor & Francis Group, LLC Ballistics: Theory and Design of Guns and Ammunition 26 2.5 Combustion As stated in the previous two sections, combustion is the process through which the energy of the solid propellant is converted to useful work. The purpose of this section is to quantify the oxidation reaction. The tactic we shall employ is to examine the more common, everyday combustion processes that combine (relatively) simple fuels with air to produce work. In this way, we shall, hopefully, bring to mind the combustion thermodynamics that has been taught at an undergraduate level and perhaps has been forgotten or not exercised since it was first learned. If we utilize the concept of a fixed CV, we can imagine a combustion chamber as depicted in Figure 2.7. In this CV, we can envision a mass of fuel entering as well as some mass of air. The two are then combusted with one another and the gaseous products leave as a mixture. We can write the first law of thermodynamics for this system then as in Equation 2.46, which we shall repeat here with subscripts that reflect Figure 2.7: 1 2 1 2  1 2   ɺ  ɺ products  hproducts + Vproducts ɺ fuel  hfuel + Vfuel Qɺ + Wɺ = m  − mair  hair + Vair  − m  2 2 2       (2.62) In Equation 2.62, we can see how the heat and energy generated are affected by the amount of mass flow, the enthalpies, and the velocities of the fuel, the oxidizer (air in this case), and the product gases. We require some means of determining the energy converted through the chemical reaction. We achieve this through the balancing of the chemical reaction. We shall return to Equation 2.62 once we have discussed chemical reactions. One of the most important compounds in the study of combustion is air. We shall adopt a convention that is standard in many thermodynamics texts [5,13,14] that models air as 21% diatomic oxygen (O2) and 79% diatomic nitrogen (N2). This means that every mole of oxygen carries with it 3.76 moles of nitrogen. This relationship comes about because  moles N 2  0.79   mole air  = 3.76  moles N 2   mole O   moles O 2  2   0.21    mole air  (2.63) As can be seen in Appendix B.1, the molecular weight for our simple model of air is 28.97 kg/kg-mol. • mair • • mfuel FIGURE 2.7 Fixed control volume (CV) combustion chamber. © 2014 by Taylor & Francis Group, LLC mproducts Physical Foundation of Interior Ballistics 27 The balancing of a chemical reaction determines what the molecular composition of the combustion products will be and furthermore helps us to quantify the amount of energy absorbed or released. If energy is absorbed in a chemical reaction, in other words, if we had to add energy to force the reaction to completion, the reaction is said to be endothermic. If heat is liberated, the reaction is said to be exothermic [15]. A reaction can be said to be theoretically or stoichiometrically balanced if the reaction goes to completion and there is no excess oxygen in the products [1]. We shall define a complete reaction as one in which all of the oxygen combines first with all of the hydrogen to form steam and then with all the carbon to form carbon dioxide. Oxygen has a greater affinity for combining with hydrogen than with carbon [1]. The only time that carbon monoxide (CO) will be formed is if there is insufficient oxygen. We must keep in mind that in any real reaction there will usually be some amounts of carbon monoxide and other compounds such as nitric oxide (NO) in the combustion products. We shall return to this issue later. For the time being, we shall assume that the only reaction products in the stoichiometric reaction are CO2 and H2O. The balancing of these chemical reactions is an important part of our study of the combustion process, which we shall now examine. We shall use two convenient forms of chemical equations: a molar-based equation and a mass-based equation. In the molar-based equation, we shall usually combust one mole of fuel with some amount of air. The result may be multiplied by the number of moles of fuel actually burned to obtain a final answer. When the mass-based equation is employed, we generally use one mass unit of fuel (lbm or kg) and some amount of air, again multiplying the solution by whatever the actual mass of fuel happens to be. The techniques just described are applicable to a system where the mass is fixed. The same equations can be used with mass or molar flow rates if the system happens to be a steady flow or open system. It is informative to balance the chemical reactions in the context of everyday systems that combust a fuel with air. Usually, this fuel is a hydrocarbon composition. The stoichiometric amount of air required would be enough so that all of the carbon combusts with sufficient oxygen to form CO2 and all of the hydrogen combusts to form water or steam. If we had a hydrocarbon fuel of chemical composition CxHy, we would like to find the number of moles, a, of air required to completely combust the fuel and we would write the balanced chemical reaction as C x H y + a(O 2 + 3.76N 2 ) → xCO 2 + y H 2O + 3.76 aN 2 2 (2.64) We could solve for a to yield a= x+ y 4 (2.65) As an example, let us say we have one mole of benzene (C6H6) that we would like to burn in air. The balanced, stoichiometric equation would be found by first determining a from Equation 2.65 a = 6+ © 2014 by Taylor & Francis Group, LLC 6 = 7.5 4 (2.66) Ballistics: Theory and Design of Guns and Ammunition 28 Now the balanced equation is found using Equation 2.64 C6H6 + 7.5(O 2 + 3.76N 2 ) → 6CO 2 + 3H 2O + 28.2N 2 (2.67) This is an example of a stoichiometrically balanced equation using a molar basis. There are times when a particular fuel is burned with too much air (over oxidized) or too little air (under oxidized). The latter is usually the case with propellants in the chamber of a gun. When a fuel is over oxidized, we usually categorize it by stating how much excess air is included in the reaction. For instance, 50% excess air used in the reaction of Equation 2.67 would alter the balanced equation to be written as C6H6 + (1.5)(7.5)(O 2 + 3.76N 2 ) → 6CO 2 + 3H 2O + 3.75O 2 + 42.3N 2 (2.68) If the fuel were burned with 50% deficient air, we would have C6H6 + (0.5)(7.5)(O 2 + 3.76N 2 ) → 4.5CO + 3H 2O + 1.5C + 14.1N 2 (2.69) In this case, we have used the rules set forth earlier where steam is formed first then carbon monoxide. At this point, all of the oxygen has been used up so solid carbon is formed. From this simple example, you can see that the amount of air used in the combustion is critical to determination of the products. We can now define an air–fuel ratio as the ratio mass of air combusted to the mass of fuel combusted. This is given mathematically by A−F = ɺ mair m = air ɺ fuel mfuel m (2.70) If we continue using our three examples, we could find the mass fuel ratio for each of the reactions defined in Equations 2.67 through 2.69. If we note here that the molar mass of Benzene is 78.11 lbm/lb-mol and the molar mass of air is 28.97 lbm/lb-mol, we have for the stoichiometric reaction A − FStoich  lbm  (7.5) [mol air ]( 4.76)(28.97 )   lb-mol  = 13.24  lbm air  = 13.24 =    lbm   lbm C6 H6  (1) [mol C6 H6 ](78.11)   lb-mol  (2.71) For the reaction with 50% excess air, we have A − F50%excess  lbm  (1.5)(7.5) [mol air ]( 4.76)(28.97 )   lb-mol  = 19.85  lbm air  = 19.85 =    lbm   lbm C6 H6  (1) [mol C6 H6 ](78.11)   lb-mol  © 2014 by Taylor & Francis Group, LLC (2.72) Physical Foundation of Interior Ballistics 29 For the reaction with 50% deficient air, we have A − F50%deficient  lbm  (0.5)(7.5) [ mol air ] ( 4.76)(28.97 )   lb-mol  = 6.61  lbm air =   lbm   lbm C6 H6 (1) [ mol C6 H6 ] (78.11)   lb-mol     = 6.61  (2.73) Now that we have introduced the process of chemical equation balancing and some of the mathematics required, we must quantify the energy released (or absorbed) by the chemical reaction. We have already introduced the concept of enthalpy as well as defined the enthalpy of formation. We shall pause here to examine how a heat of formation is obtained. We shall consider carbon dioxide for our example. If we have a combustion chamber in which we react pure oxygen with solid carbon, we can put the two substances into the container at 25°C and start the reaction somehow. The balanced equation on a molar basis would be C(s) + O 2 → CO 2 (2.74) The first law of thermodynamics states that Q + W = N products hproducts − N reactants hreactants (2.75) Here we have used specific values so that everything is on a molar basis. Since the container is rigid, there is no work performed on or by the system, thus Equation 2.75 reduces to Q = N products hproducts − N reactants hreactants (2.76) If we were to perform this experiment, we would find that the container would get hot. Theoretically, we could extract this heat from the container until the temperature returned to 25°C; if we were to do this, we would find that 393,546 kJ/kg-mol of energy would have been produced. Examination of Appendix B.1 reveals that this is exactly the value of the heat of formation of carbon dioxide recalling that a negative value denotes heat given off by the reaction. The enthalpy of a substance allows us to quantify the energy state of a material. The enthalpy of formation was defined as the energy required to form a particular composition from its basic elements resulting in the compound as a product at some reference temperature and pressure (we shall use 25°C or 298 K and 1 atm as this reference condition). If we were to take this compound and arbitrarily increase its temperature or pressure by some amount and if there were no phase change or change in composition, we will have increased its enthalpy. If we restrict our analysis to an ideal gas, it can be shown [1] that the enthalpy is a function of temperature only. With this, we can write for a composition hT = hf0 + ∆h298 →T where hT is the enthalpy of the material at temperature T, hf0 is the enthalpy of formation ∆h298→T is the change in enthalpy from the reference state to the temperature, T © 2014 by Taylor & Francis Group, LLC (2.77) Ballistics: Theory and Design of Guns and Ammunition 30 We define ∆h298→T as 0 ∆h298 →T = h (T ) − ( h298 ) (2.78) Tables of enthalpies are located in Appendix B at the end of the book. As an example, consider carbon monoxide at 2000 K. The enthalpy of this compound using Appendices B.1 and B.2 would be  kJ   kJ   kJ  hCO2000 K = −110, 541   + 56, 737   = −53, 804    kg-mol   kg-mol   kg-mol  (2.79) Now that we have worked with enthalpies a bit, we can begin to apply what we have learned. We shall look at an example of these principles applied first to a closed bomb where there is no work performed and then to a gun where there is. For a closed bomb, we shall tailor Equation 2.42 to our needs. If we consider a closed vessel, we realize that there is no velocity into or out of the CV, and there is no work performed on or by the system. This allows us to write Equation 2.42 as Q1− 2 = m[( h2 − pv2 ) − ( h1 − pv1 )] = m(u2 − u1 ) (2.80) If we write this equation on a molar basis as limit to ideal gas behavior, we can state that Q= ∑ N (h i i prod − RuTprod ) − ∑ N (h i reac − RuTreac ) (2.81) i This relationship is important because it tells us that the heat given off by the closed bomb is affected by the enthalpy change of the chemical reaction and the temperature of the products. We shall examine a pressure vessel containing 0.001 kg of methane (CH4) and 0.002 kg of air. The enthalpy of formation for methane is −74,850 kJ/kg-mol and its molecular weight is 16.04 kg/kg-mol. The reaction will begin at 298 K and we shall remove enough heat from the vessel that the final temperature becomes 1500 K. We would like to determine how much heat is given off. We need to balance the chemical reaction on a molar basis, so we shall determine how many moles of methane and air we have in the container. For methane, we have N CH4 = (0.001) kg CH4  = 6.23 × 10 −5 kg-mol CH4   kg  (16.04)    kg-mol  (2.82) For the air, we have N air = (0.002) kg air  = 6.90 × 10 −5 kg-mol air   kg  (28.97 )    kg-mol  © 2014 by Taylor & Francis Group, LLC (2.83) Physical Foundation of Interior Ballistics 31 Our balanced reaction is then (6.23 × 10 −5 )CH 4 + (6.90 × 10 −5 ) ( (0.21)O 2 + (0.79)N 2 ) → (2.9 × 10 −5 )H 2O + (9.56 × 10 −5 )H 2 + (6.23 × 10 −5 )C(s) + (5.45 × 10 −5 )N 2 (2.84) We shall examine the reactants first. For the methane, we have ( N CH4 hf0 + ∆h298 →T − RuTCH4 )    kJ    kJ + 0 − (8.314)  = (6.23 × 10 −5 ) [kg-mol]  −74, 850  (298) [K]       kg-mol   kg-mol ⋅ K    ( ) N CH4 hf0 + ∆h298 →T − RuTCH4 = −4.82 [kJ] For the oxygen and nitrogen, we have     kJ (298) [K]  N O2 hf0 + ∆h298 →T − RuTO2 = (1.45 × 10 −5 ) [kg-mol]  0 + 0 − (8.314)      kg-mol ⋅ K    ( ) ( ) N O2 hf0 + ∆h298 →T − RuTO2 = −0.036 [kJ]     kJ N N2 hf0 + ∆h298→T − RuTN2 = (0.79)(6.90 × 10−5 )[kg-mol]  0 + 0 − (8.314)   (298) [K] ⋅ kg-mol K     ( ) ( ) N N2 hf0 + ∆h298→T − RuTN2 = −0.135 [kJ] The enthalpies of the reactants are therefore ∑ N (h i reac ) − RuTreac = −4.82 [kJ] − 0.036 [kJ] − 0.135 [kJ] = −4.99 [kJ] i For the products we have (using the tables in the appendix) ( N H2 O hf0 + ∆h298 →T − RuTH2 O )     kJ = (2.9 × 10 −5 ) [kg-mol]  −241, 845 + 48, 181 − (8.314)   (1, 500) [K]    kg-mol ⋅ K    ( ) N H2 O hf0 + ∆h298 →T − RuTH2 O = −5.98 [kJ] © 2014 by Taylor & Francis Group, LLC Ballistics: Theory and Design of Guns and Ammunition 32 ( N H2 hf0 + ∆h298 →T − RuTH2 )     kJ (1, 500) [K]  = (9.56 × 10 −5 ) [kg-mol]  0 + 36, 307 − (8.314)      kg-mol ⋅ K    ( ) N H2 hf0 + ∆h298 →T − RuTH2 = +2.28 [kJ] ( N C hf0 + ∆h298 →T − RuTC )      kJ  kJ − (8.314)  = (6.23 × 10 −5 ) [kg-mol]  0 + 23, 253  (1, 500) [K]       kg-mol ⋅ K   kg-mol    ( ) N C hf0 + ∆h298 →T − RuTC = 0.67 [kJ] ( N N2 hf0 + ∆h298 →T − RuTN2 )    kJ    kJ − (8.314)  = (5.45 × 10 −5 ) [kg-mol]  0 + 38, 404  (1, 500) [K]       kg-mol   kg-mol ⋅ K    ( ) N N2 hf0 + ∆h298 →T − RuTN2 = 1.41 [kJ] The enthalpies of the products are then given by ∑ N (h i prod − RuTprod ) = −5.98 [kJ] + 2.28 [kJ] + 0.67 [kJ] + 1.41 [kJ] = −1.62 [kJ] i The heat given off by the reaction is then calculated through Equation 2.81 as Q = (−1.62) [kJ] − (−4.99) [kJ] = +3.37 [kJ] (2.85) This illustrates the process of calculating the amount of energy given off by a closed bomb reaction as well as the effect of temperature on the reaction products. In this case, energy has to be added to get the reaction to go to completion. It must be noted that had we decided to lower the temperature of the products, energy would eventually have to have been removed. This will be examined as a problem at the end of the chapter. If we apply the same principles to a gun launch, we can determine the amount of energy imparted to the projectile and in so doing, obtain a feeling for the process of energy conversion between propellant chemical energy and projectile kinetic energy. Unlike the fixed boundary examined in the closed-bomb problem, a gun launch involves a boundary that is moving (the base of the projectile). This problem is similar to a piston of an internal combustion engine that undergoes one stroke. We have defined work earlier as a form of energy and if we assume all of the energy of the propellant goes into heating of © 2014 by Taylor & Francis Group, LLC Physical Foundation of Interior Ballistics 33 the gaseous products, kinetic energy of the projectile, and a loss term (including friction, swelling of the gun tube, etc.), we can write the first law of thermodynamics as given in Equation 2.75. Rewriting this by assuming the velocity of the seated projectile is zero, we obtain our thermodynamic equation for a gun launch as Q+ 1 mV 2 = 2 ∑ N (h i i prod )− ∑ N (h i reac ) + losses (2.86) i We have neglected potential energy changes here because they are usually quite small relative to the other terms. We shall examine an example in the form of a potato gun to illustrate the use of Equation 2.86 and the other methods of this chapter. A potato gun is a device that people use to project potatoes at targets. These devices can be very dangerous to the operator as well as the target. We would like to calculate the muzzle velocity of a half-pound potato projectile used in a particular gun. This gun is made of 2 in. diameter PVC pipe (a very good insulator). The projectile rests on a stop when loaded through the muzzle so that there is a 6 in. long chamber. The device in question was injected with 0.005 oz (mass) of lighter fluid as a gas (n-butane—C4H10 (g) hf0 = −124, 733 kJ/kg-mol , n = 58.123 kg/kg-mol) to fire the potato. We shall assume the potato obturates perfectly and that there is no bore friction. The travel of the potato in the gun tube is 24 in. The weapon is fired under standard conditions of 77°F and 14.7 psi. Assume the reactants and the products both exist at these conditions. We would like to determine the velocity of the potato at the completion of combustion in feet per second assuming no losses. The chamber was 6 in. long and 2 in. in diameter, so our chamber volume is Vi = Al = π (2)2 [in.2 ](6) [in.] = 18.85 [in.3 ] 4 (2.87) The air weighs 28.97 lbm/lb-mol and if we assume ideal gas behavior, the density of air is calculated from pv = RT → ρ = p RT (2.88)  lbf   lbm  (14.7 )  2  (28.97 )   lbm  in.    lb-mol  ρ= = 0.0000428  3   ft-lbf   in.   in.  (12)   (537 ) [R ] (1545)    lb-mol − R   ft  So the amount of air we actually have is  lbm  mair = ρ Vi = (0.0000428)  3  (18.85) [in.3 ] = 0.0008068 [lbm]  in.  The amount of fuel was given in ounces  lbm  mfuel = (0.005) [oz](0.0625)  = 0.0003125 [lbm]  oz  © 2014 by Taylor & Francis Group, LLC (2.89) Ballistics: Theory and Design of Guns and Ammunition 34 For the actual combustion, we need to use our mass information and convert it to molar values, recognizing that the molar mass is the same whether it is kg/kg-mol or lbm/lb-mol. For the fuel and air, we have N fuel = mfuel = (0.0003125) [lbm] nfuel 1 = 0.0000054 [lb-mol]  lbm  (58.123)   lb-mol  N air = mair = (0.0008068) [lbm] nair 1 = 0.00000278 [lb-mol]  lbm  (28.97 )   lb-mol  (2.90) (2.91) For each lb-mol of air, we know that 1/4.76 lb-mol of it is oxygen so we have N O2 = 1 (0.0000278) [lb-mol] = 0.0000058 [lb-mol] 4.76 N N2 = 3.76 (0.0000278) [lb-mol] = 0.0000220 [lb-mol] 4.76 Now we can write our combustion equation as (0.0000054)C 4H10 ( g ) + (0.0000058)O 2 + (0.0000220)N 2 → (0.0000116)H 2O + (0.0000216)C + (0.0000154)H 2 + (0.0000220)N 2 To determine the muzzle velocity, we start with our first law of thermodynamics equation, simplified by the fact that there is no heat transfer and no shaft work. Then the energy of the fuel–air mixture equals the work done on the projectile plus the energy of the products of combustion. H R = H p + Wp (2.92) Let us look at the internal energies for each of the reactants Reactant Enthalpy of Formation (kJ/kg-mol) Enthalpy of Formation (in.-lbf/lb-mol) C4H10 (g) O2 N2 −124,733 0 0 −500,728,155 0 0 The conversion used here is as follows:  BTU     kJ   ft-lbf   in.   in.-lbf  lb-mol ( x)  (12)   → 4014.4 x   (0.4299)  kJ  (778..16)    BTU   ft   lb-mol     kg-mol   kg-mol  © 2014 by Taylor & Francis Group, LLC (2.93) Physical Foundation of Interior Ballistics 35 For the products, we have Product Enthalpy of Formation (kJ/kg-mol) Enthalpy of Formation (in.-lbf/lb-mol) H2O (g) N2 C2 H2 −241,845 0 0 0 −970,862,568 0 0 0 We will rearrange our first law equation as follows: Wp = H R − H p We calculate HR first ( ) ( ) ( H R = N C4 H10 hf0 + ∆h298 →T + N O2 hf0 + ∆h298 →T + N N2 hf0 + ∆h298 →T ) Plugging in the numbers we have, we get  in.-lbf  H R = (0.0000054) [lb-mol](−500, 728 , 155 + 0)   lb-mol   in.-lbf  + (0.0000058) [lb-mol](0 + 0)   lb-mol   in.-lbf  + (0.0000220) [lb-mol](0 + 0)   lb-mol  H R = −2704 [in.-lbf ] We calculate Hp in a similar manner ( ) ( ) ( ) ( H p = N H2 O hf0 + ∆h298 →T + N H2 hf0 + ∆h298 →T + N N2 hf0 + ∆h298 →T + N C hf0 + ∆h298 →T  in.-lbf  H p = (0.0000116) [lb-mol](−970, 862, 568 + 0)   lb-mol   in.-lbf  + (0.0000154) [lb-mol](0 + 0)   lb-mol   in.-lbf  + (0.0000220) [lb-mol](0 + 0)   lb-mol   in.-lbf  + (0.0000216) [lb-mol](0 + 0)   lb-mol  H p = −11, 262 [in.-lbf ] © 2014 by Taylor & Francis Group, LLC ) Ballistics: Theory and Design of Guns and Ammunition 36 Then the work done on the projectile is Wp = −2, 704 [in.-lbf ] − (−11, 262) [in.-lbf ] = 8, 558 [in.-lbf ] Since this work equals the muzzle energy of the projectile Wp = 1 mV 2 = 8, 558 [in.-lbf ] 2 Therefore, V= 2Wp = m  lbm-ft  (2)(8558) [in.-lbf ](32.2)  2  lbf-s  = 303  ft   s   in.  (0.5) [lbm](12)    ft  Wow! That is pretty fast but we used a lot of butane, assumed the products return to ambient conditions quickly, and neglected things. Also note that the length of the tube did not come into play. We would definitely have to account for this as we shall later see. One important parameter in determining the amount of energy transferred to the projectile is the temperature of the product gases. As you can see from our example, an increase in the temperature of the product gases will result in a decrease in the projectile velocity because Hp goes up. Typically, we can assume the product gases exit at a temperature between 0.6T0 and 0.7T0, where T0 is the adiabatic flame temperature of the product gases [7]. The adiabatic flame temperature of a gas is the temperature that is achieved if the gases burn to completion in the absence of any heat transfer or work being performed [1]. The calculation of the adiabatic flame temperature is relatively straightforward but requires iteration. This is beyond the scope of this text but the reader is referred to the references at the end of this chapter for a complete description of the procedure. In addition, there are several commercially available codes (including some that come with the purchase of textbooks now, for instance [13]). To achieve our objectives, the temperature of the reaction products will always be given. Problem 6 Calculate the A–F ratio for the combustion of the following fuels. Calculate the ratio with both theoretical air and 10% excess air. 1. Benzene (C6H6) Answer: 13.24 and 14.56 2. n-Butane (C4H10) Answer: 15.42 and 16.96 3. Ethyl alcohol (C2H5OH) Answer: 8.98 and 9.88 Problem 7 Let us examine a pressure vessel identical to the example problem in the text containing 0.001 kg of methane (CH4) and 0.002 kg of air. The enthalpy of formation for methane is −74,850 kJ/kg-mol and its molecular weight is 16.04 kg/kg-mol. The reaction will begin at 298 K and we shall remove enough heat from the vessel that the final temperature becomes 1000 K. © 2014 by Taylor & Francis Group, LLC Physical Foundation of Interior Ballistics 37 1. Determine the maximum heat given off. Answer: Q = +0.606 [kJ]. 2. Compare the result in (1) with the example problem in this chapter. Answer: This situation removes 2.764 kJ more energy than the example. Problem 8 A really interesting person takes the tennis ball mortar we built in problem 3 and modifies it—squirting in and igniting 0.003 oz. of acetylene gas (C2H2(g)). If we assume the combustion kinetics are fast enough such that the energy release occurs before the ball can move, we want to determine the muzzle velocity of the tennis ball. Proceed along the following steps: a. Balance the stoichiometric reaction equation for acetylene. b. Balance the actual equation neglecting the volume the acetylene occupies in the chamber. Assume the air initially in the chamber is at 14.7 psia and 77°F. c. Determine the increase in internal energy of the gas. d. Assuming the gas is calorically perfect (∆U = mgcv∆T) and that cv = 0.33 BTU/lbm-R for the mixture, determine the increase in temperature of the gas. you will have to do (c) and (d) by iteration, first assuming a final reaction temperature, carrying out the calculation for ∆U and seeing if the ∆T you get matches—if not iterate again – once you get an answer within say 10% that is good enough. No t e : e. Based on the result of (d), determine the initial pressure on the tennis ball assuming the specific gas constant of the products is R = 80 ft-lbf/lbm-R. f. Use the result of (e) and possibly your results from problem 2 to determine the muzzle velocity of the tennis ball. Assume γ = 1.4 g. Determine the temperature of the gases at shot exit. For acetylene, n = 26.038 lbm/lbmol, ∆h0f = +97,477 BTU/lbmol. Problem 9 A potato is stuffed into the 3” diameter exhaust pipe of a car that is not running too well. 1 g of incompletely combusted combustion products (assume gaseous Heptane) mixes with a stoichiometric amount of air behind the potato and ignites. If the potato is wedged 4” into the exhaust (i.e., it has 4” of travel) and weighs 0.25 lbm, and assuming the combustion takes place before the potato moves, determine the theoretical maximum “muzzle” velocity of the potato. Also calculate the muzzle velocity assuming an isentropic expansion. For gas expansion purposes you can assume the volume available initially behind the potato is equal to the volume between the point of obturation and the end of the exhaust and assume a “smeared” specific heat ratio of 1.3 for the product gases. Also assume the combustion begins at 500 K and completes at 1500 K. Assume the total enthalpy at 500 K for n-Heptane (C7H16) is −120,000 kJ/kg-mol. Problem 10 Assume we have a quantity of 29 lb of 11.1% nitrated nitrocellulose (C6H8N2O9) and it is placed in an empty chamber of a gun at 77°F and 14.7 psia. The chamber is 1160 in.3 © 2014 by Taylor & Francis Group, LLC Ballistics: Theory and Design of Guns and Ammunition 38 in volume. The propellant density is 0.060 lbm/in3. If the air in the chamber is NOT neglected and assuming the volume is fixed: a. Write the balanced equation for this combustion (assume the oxygen goes preferentially into CO2 instead CO this time—you will find the difference later). b. Using the tables in the textbook, estimate the adiabatic flame temperature of the resultant gas (Hint: recall the definition of adiabatic flame temperature). 2.6 Solid Propellant Combustion Now that we have examined the background of the thermochemistry and thermodynamics of combustion, we shall see how this applies to the behavior of a burning solid propellant. We shall endeavor, in this section, to come up with definitions and relationships that will allow us to define the state of the propellant behind a projectile at any given time. The process we will use is somewhat simplified because the real situation behind a moving projectile is generally a two phase, reacting flow field. Some of our assumptions, though not necessarily valid in the purest sense, are good enough to predict bulk behavior of the propelling gas. In the previous sections, we have discussed how energy is evolved by the propellant. We saw that thermodynamic properties were not dynamic at all, merely means of accounting for energy knowing the initial and end states and making assumptions on the process between them. This section will allow us to add in some time dependency to the equations to somewhat understand the rates at which combustion is occurring. Solid propellants are generally nitrocellulose compounds that are manufactured by nitrating through immersion in acid. The details of this process for various materials can be examined in detail in Refs. [7,8,16–18]. This material is then chopped and worked into a doughy substance and pushed though dies to form various shapes. The material then has solvents removed and it is dried. When this process is complete, the propellant has the consistency of uncooked (i.e., hard and somewhat brittle) pasta. Though this statement is general, there are, as always, exceptions. The burning of solid propellant is a surface phenomenon. The rate of gas evolution is dependent upon the amount of surface area of the propellant. Because of this, the shape that the propellant takes is extremely important. Burning is the mechanism of transforming the solid propellant to a gas. The burn rate of a propellant is highly dependent upon the pressure at which the burning reaction takes place. Essentially, the greater the pressure, the faster the propellant burns. These two behavioral observations tell us that if we can control the geometry and confinement of a given propellant, we can, to a large degree, control the rate of gas evolution. We shall examine a single propellant grain to gain an understanding of how the geometry affects the rate of evolution of gas. Consider a long cylinder of solid propellant that is commonly referred to as a grain. If the cylinder were long enough, we could see that most of the surface area would be located along the circumference and length. In other words, we can neglect the two small surface areas that comprise the ends. This is illustrated in Figure 2.8. If we neglect the burning of the end surfaces, it allows us to examine the geometry through simple mathematical relationships. As our grain begins to burn, solid material will be evolved into gas. Thus, we can imagine the solid surfaces shrinking toward the centerline of the grain. If we examine our grain from the end looking down its axis, we would see a circular section as depicted in © 2014 by Taylor & Francis Group, LLC Physical Foundation of Interior Ballistics 39 This is where most of the surface area is located FIGURE 2.8 Long cylindrical propellant grain. d FIGURE 2.9 Propellant grain cross section. Figure 2.9. We could then write an expression for the surface area of our grain as a function of its diameter and length: A(t) = πd(t)l (2.94) In this expression, A(t) is the surface area of the grain, d(t) is the diameter, and l is the length. We have denoted the surface area and diameter as functions of time to remind us of our assumption of no burning at the ends of the grain. After some time, t, the grain surface will have regressed such that our diameter has decreased. This is depicted in Figure 2.10. This graphically shows us that at time t1 the grain clearly has more surface area than at time t2; therefore, as burning progresses, the rate of evolution of gas slows down. This is commonly called regressive burning. Propellant geometry is characterized by a quantity known as the web thickness or simply the web. The symbol used for the web is D. The web is the smallest thickness of the initial propellant grain. In the case of our cylindrical grain, it would be the initial diameter. d(t1) d(t2) FIGURE 2.10 Propellant grain cross section at two times. © 2014 by Taylor & Francis Group, LLC Ballistics: Theory and Design of Guns and Ammunition 40 In the interior ballistics analysis of a gun system, we need to track how much gas is evolved and also how much solid is remaining. This is important because we have seen that all of our equations of state are dependent upon volume as well as pressure and temperature, and these, in turn, affect the burning rate. The amount of solid propellant remaining is tracked through use of the web fraction, f. The web fraction is the fraction of web remaining at a given time, t. Through use of this web fraction, we can write an expression for the amount of propellant remaining at any time as a function of the web: d(t) = fD (2.95) This is illustrated for a grain with a single perforation (known colloquially as a perf) in Figure 2.11. It is important to note here that for a single perf grain, the web is defined as the outside radius minus the inside radius. This sometimes is confusing for new ballisticians since we use D as the web thickness. Also one can see from the figure that an advantage of a single perf grain is that it burns from both the inside out and the outside in, thus decreasing the surface on the outside while increasing the surface on the inside—known as neutral burning behavior. Use of the web fraction is convenient because, mathematically, it is a function that varies from unity to zero. The manner in which it varies may be somewhat complex, but at least the end states are well defined. An example plot of web fraction versus time is shown in Figure 2.12. In this figure, tB is the time at which all of the propellants have evolved into gas—the burnout time. fD D FIGURE 2.11 Burning of a single perforated propellant grain. f 1 t f = fraction of remaining web f = f (t) FIGURE 2.12 Fraction of remaining web versus time. © 2014 by Taylor & Francis Group, LLC f (0) = 1 f (tB) = 0 tB Physical Foundation of Interior Ballistics 41 Many times, we are interested more in the volume of the propellant that has evolved into gas rather than the fraction of the web remaining. It should be clear that the two quantities are related since the gas had to come from the solid material and conservation of mass states that we can neither destroy nor create mass. This is handled through use of the fraction of propellant burnt, ϕ. Since ϕ is a function of f and f is a function of time, we see that ϕ must be also a function of time. Since propellant geometries can be fairly complicated, ϕ can be a rather complicated function of f. For simple shapes, this relationship is straightforward. For instance, a single perforated grain has the functional relationship that φ (t) = 1 − f (t) (2.96) Most shapes can be simplified to express ϕ as a quadratic function of f through use of a shape function, θ: φ (t) = [1 − f (t)][1 + θ f (t)] (2.97) This expression allows us to cover almost any simple geometry, the most notable exception being a sphere. Figure 2.13 depicts how variation in the shape function affects the relationship between f and ϕ. With the formulations earlier, we have been able to mathematically define the effect of propellant geometry on the rate of gas evolution. The second important parameter in this generation of gas was stated to be the effect of pressure on burning. Whenever a propellant burns, say in a fixed volume, two competing processes are happening: the volume into which the gaseous propellant is moving is increasing because there is less solid material—this decreases the pressure, and the more and more propellant gas is being pushed into a confined space—this increases the pressure. The rate at which the Form functions for various θ and spherical grains 1.0 0.9 0.8 0.7 (t) 0.6 0.5 θ = –1 0.4 θ = –0.4 0.3 θ=0 0.2 θ = 0.4 θ=1 0.1 Spherical grain 0.0 1.0 0.9 0.8 0.7 0.6 0.5 f(t) FIGURE 2.13 Effect of different values of θ on ϕ and f. © 2014 by Taylor & Francis Group, LLC 0.4 0.3 0.2 0.1 0.0 Ballistics: Theory and Design of Guns and Ammunition 42 surface area decreases affects this relationship. The simplest model for the relationship between burn rate and pressure is given by D df = − β pB (t) dt (2.98) where D(df/dt) is the time rate of change of the web (i.e., the burning rate) β is a burn rate coefficient pB is the pressure (we will discuss the subscript later) The negative sign comes about because the amount of propellant would be increasing if D(df/dt) were to result in a positive number. This simple relationship facilitates our analysis of propellant behavior in a gun. Other relationships can more accurately describe propellant behavior, but their complexity is such that computer codes must be used to obtain answers with them. Two very common burn relationships are df = − β [ pB (t)]α dt (2.99) df = − β [ pB (t) − P1 ] dt (2.100) D D Equation 2.99 is by far the most commonly used in computer codes. Caution must be exercised when using burn rate data from the literature as the units will be an indicator of the proper burn rate form of the governing equation. If we examine the units of D df/dt, we see that they are in terms of [length]/[time]. This type of data are usually obtained from a strand burner. A strand burner is a device that can accurately measure the rate of linear burning in a propellant. Reference [19] contains an excellent diagram of a strand burner. If we consider a pressure vessel so thick as to be rigid and the amount of propellant so small such that we can neglect its contribution to the volume, we can describe the burning of the propellant as a constant volume process. This is the essence of closed-bomb testing. We can further assume that this pressure vessel can be isolated thermally and the gas behavior is ideal. In this case, our closed bomb, with internal volume, V, would resemble Figure 2.14. V FIGURE 2.14 Diagram of a closed bomb. © 2014 by Taylor & Francis Group, LLC Physical Foundation of Interior Ballistics 43 Since we assumed ideal gas behavior, we can write an expression for pressure as a function of volume and temperature pB V = mg RT (2.101) where pB is the pressure V is the volume mg is the mass of the gas R is the specific gas constant T is the temperature When we place our solid propellant into the closed bomb, it has an initial weight that we would call c. So initially, we can write c = ρ Vsolid (2.102) where ρ is the density of the solid propellant Vsolid is its volume If we now assume that the propellant is cylindrical, we can write its volume as the product of its cross-sectional area and its length. The initial diameter is the web for a cylindrical grain, so Vcyl.grain = π D2 l 4 (2.103) [d(t)]2 l 4 (2.104) This volume at any time, t, can be expressed as V(t)cyl.grain = π Because mass is conserved, the amount of solid propellant burned is equal to the amount of gas generated. This is an important concept. If we started with 1 lbm of propellant and completely burned it, we would be left with 1 lbm of gas. Based on this, we can write for the mass of the gas as mg (t) = ρ  Vcyl.grain − V(t)cyl.grain  = ρ π l D2 − [d(t)]2 4 { } (2.105) We discussed the fraction of propellant burnt, ϕ, earlier. We are now in a position to formally define it as φ (t) ≡ © 2014 by Taylor & Francis Group, LLC mg (t) c (2.106) Ballistics: Theory and Design of Guns and Ammunition 44 If we substitute Equations 2.102, 2.103, and 2.105 into Equation 2.106, we obtain φ (t) = ρ π l D2 − [d(t)]2  [d(t)]2  4 = 1 −  π D2   ρ lD2 4 { } (2.107) Now we insert Equation 2.95 into Equation 2.107 to yield 2  fD  2 φ (t) = 1 −   = 1 − f = (1 − f )(1 + f ) D   (2.108) Comparing this expression (derived for a cylindrical grain) to Equation 2.97 shows that the shape factor θ = 1 for a cylindrical grain. Also, by comparison to Equation 2.96 we see that the shape factor θ = 0 for a single perforated grain. Essentially, any shape factor can be derived using this same procedure. So up to this point, we have determined that the shape factor θ = 0 for single perforated grains θ = 1 for cylindrical grains An interesting thing has happened. We started this section attempting to find a relationship for the mass of gas evolved from the solid propellant and we have come around to finding the relationship between ϕ and f again. The key procedure here is now to rearrange Equation 2.106: mg (t) = cφ (t) (2.109) This is the relationship that governs the amount of gas evolved from the burning propellant. It looks rather simple, but consider that ϕ is a function of f and t, and f is a function of pB and t. We shall return to this later. The burning propellant in our closed bomb must generate pressure. To take this further, we need to rearrange Equation 2.98 into pB (t) = − D df β dt (2.110) In this expression, we know that D is the initial web and therefore a constant, and we shall assume that β is a constant (β actually increases somewhat with pressure). Because we want to work with masses of substances, f is not a convenient variable. We shall use a relationship to express it in terms of ϕ. At this point, caution must be exercised. Recall that the relationship between ϕ and f varies with propellant geometry. We shall proceed using our cylindrical grain relationship (Equation 2.108). Rearranging Equation 2.108, we obtain f (t) = 1 − φ (t) (2.111) If we differentiate this relationship with respect to time, we obtain df 1 dφ =− dt 2 1 − φ (t) dt © 2014 by Taylor & Francis Group, LLC (2.112) Physical Foundation of Interior Ballistics 45 This form allows us to rewrite Equation 2.110 as pB (t) = − dφ D 2β 1 − φ (t) dt (2.113) We now have all the expressions we need to bring this together. We have an equation of state pB (t)V = mg (t)RT (t) (2.114) We have an expression for conservation of mass (relationship between mg and ϕ) mg (t) = cφ (t) (2.115) and we have an expression that relates the amount of pressure generated to the amount of propellant burnt (burn rate equation) pB (t) = − dφ D 2β 1 − φ (t) dt (2.116) All these expressions are in terms of constants we know beforehand or f, ϕ, and T. To describe the temperature of gas, we need to define a parameter used often in interior ballistics, the propellant force, λ. Propellant force is a constant that is defined as the amount of energy released from a propellant under adiabatic conditions. In other words, it is the most energy one can obtain by burning a propellant. Mathematically, we express it as λ ≡ RT0 (2.117) where R is the specific gas constant T0 is the adiabatic flame temperature of the gas This constant has units of energy per unit mass. Sometimes, T0 is referred to as the uncooled explosion temperature. In our development, we shall assume that all gases are evolved at the adiabatic flame temperature. There are many theories that describe combustion. Introductory treatments are provided in Refs. [20,21], but all of the references in the end of this section cover the topic to some degree. References [22–24] treat the topic in great detail. If we utilize this reactive assumption, we can rewrite Equation 2.114 using Equation 2.117 to give us pB (t)V = λ mg (t) (2.118) Now we can combine Equations 2.118 and 2.116 to yield (for a cylindrical grain) λ mg (t) D dφ =− V 2β 1 − φ (t) dt © 2014 by Taylor & Francis Group, LLC (2.119) Ballistics: Theory and Design of Guns and Ammunition 46 We then substitute Equation 2.115 into the aforementioned expression, resulting in dφ λ cφ (t) D =− V 2β 1 − φ (t) dt (2.120) 1 dφ 2 β λc =− DV φ (t) 1 − φ (t) dt (2.121) This can be rearranged to yield This is a separable, first-order, nonlinear, differential equation, which can be written in integral form as 1 ∫ 0 tB 2 β λc dφ dt =− DV φ (t) 1 − φ (t) ∫ (2.122) 0 The solution of which is 1  1−φ − 1  2βλc tB t ln  =−  1 − φ + 1  DV 0 0  (2.123) This expression is somewhat problematic because of its singular behavior at ϕ = 1 and ϕ = 0. The equation was approximated numerically to yield tB ≈ 350 DV β λc (2.124) In this case, the solution to this expression was problematic; however, in many cases, it can be evaluated more readily. The techniques that will follow are much simpler from a hand calculation standpoint. Even though the closed bomb may seem academic, it is actually quite a useful device for determining propellant parameters. If we consider Equations 2.110 and 2.113, we see that since we know the initial web, D, and we can measure the pressure, the only thing missing is β and ϕ or f. Equation 2.114 tells us that if we measure pB and T and know V, we can get ϕ or f. Thus, the closed bomb is useful for determining the burn rate coefficient, β. Problem 11 M1 propellant is measured in a closed bomb. Its adiabatic flame temperature is 3906°F. Its molar mass is 22.065 lbm/lb-mol. What is the effective mean force constant in ft-lbf/lbm?  ft-lbf  Answer: λ = 305, 709   lbm  © 2014 by Taylor & Francis Group, LLC Physical Foundation of Interior Ballistics 47 Problem 12 M15 propellant was tested in a strand burner to determine the linear burning rate. The average pressure evolved was 10,000 psi. If the burning exponent, α, was known to be 0.693 and the pressure coefficient, β, was known to be 0.00330 in./s/psi0.693. Determine the average linear burning rate, B in inch per second.  in.  Answer: B( p) = 1.952    s  Problem 13 Derive the functional form of ϕ in terms of f for a flake propellant. Assume cylindrical geometry. Hint: Flake propellant consists of grains that have thicknesses much smaller than any other characteristic dimension. Answer: ϕ(t) = 1 − f Problem 14 An M60 projectile is to be fired from a 105 mm M204 Howitzer. The propellant used in this semi-fixed piece of ammunition is 5.5 lbm of M1 propellant. M1 propellant consists of single perforated grains (θ = 0) with a web thickness of 0.0165 in., if the average pressure (over the launch of this projectile) developed in the weapon is 20,455 psi. Calculate the average burning rate coefficient in in.3/lbf-s if the burn rate is (we use a negative sign in the burn rate to make the form come out right later): df = −185.9 [s −1 ] dt  in.3  Answer: β = 1.50 × 10 −4    lbf-s  Problem 15 β is actually a function of pressure and temperature (it is really given in tables at 25°F at this value). For simplification (and illustration), we will assume it is constant. Given this assumption, calculate the functional form of the web fraction, f from Problem 14. β pavg Answer: f = 1 − t D Problem 16 Given the data provided in Problems 14 and 15, determine the proper form of the fraction of charge burnt. Answer: ϕ(t) = 185.9t Problem 17 You are asked to characterize a commercial propellant. In order to do this, you take one grain of the propellant and place it in a closed bomb of 0.5 in.3 volume, initially evacuated. You have a temperature and pressure sensor in the device. After 0.063 s you decide that the propellant has fully combusted. You read the data—pressure was measured to be 3.706 psi © 2014 by Taylor & Francis Group, LLC Ballistics: Theory and Design of Guns and Ammunition 48 (this is not a big value but it was only one small grain of propellant) but it looks as though the temperature sensor is broken. The initial propellant grain weighed 0.003189 grains and it was 0.1 in. long by 0.01 in. diameter. Based on this data only— a. Estimate the propellant force, λ, in ft-lbf/lbm. b. Estimate the linear burn rate coefficient, β in in/s/psi. c. List all assumptions and explain why you believe these estimates are too high or too low (certain assumptions may make the estimates high while others make them low)—There are at least four assumptions used here. 2.7 Fluid Mechanics The entire field of ballistics is steeped in the principles of fluid mechanics. The flow of propellant gases in the gun tube, the flow of the propellant gases through a muzzle brake upon shot exit, the flow of the air around the projectile in flight, and even, as we shall see, the flow of target material during a penetration event can many times be modeled as a fluid. This section is devoted to a basic treatment of fluid mechanics principles. Some of these we will use very soon, others will be used at a later time. All of them are important in the study of ballistics. A fluid differs from a solid in its behavior when placed in shear. In general, fluids can support little or no shear loads or tensile stress. Fluids are generally characterized by their behavior under shear stress. Because a fluid will, in general, flow readily under a shear stress, this behavior is normally plotted in a graph of rate of deformation versus shear stress as depicted in Figure 2.15. A fluid is considered to exhibit Newtonian behavior if there is a linear relationship between shear stress and rate of deformation. A fluid is non-Newtonian otherwise. Some fluids such as an ideal plastic or a thixotropic material actually do exhibit a yield stress. Newtonian Rate of deformation Ideal fluid Non-Newtonian Ideal plastic Thixotropic Yield stress FIGURE 2.15 Rate of deformation versus shear stress. © 2014 by Taylor & Francis Group, LLC Shear stress Physical Foundation of Interior Ballistics 49 In the case of an ideal plastic, after a certain yield stress is achieved, the material exhibits a linear relationship between stress and deformation rate. A thixotropic material exhibits a nonlinear relationship after yield stress is reached. An ideal fluid is one where the material will flow and continue to accelerate regardless of the amount of shear stress applied. Many of the fluids we will deal with are Newtonian. Mathematically, the relationship between applied shear stress and deformation rate is given by τ =µ ∂u ∂y (2.125) where τ is the applied shear stress μ is the dynamic viscosity of the fluid 𝜕u/𝜕y is the deformation gradient (change in velocity with respect to a spatial coordinate) The ratio of the dynamic viscosity to the fluid density occurs so often that it is customary to define a kinematic viscosity as ν= µ ρ (2.126) where ν is the kinematic viscosity ρ is the density of the fluid In the section on thermodynamics, we introduced the concept of a Lagrangian or control mass approach and an Eulerian or control volume approach to solving transport problems. In examination of a fluid’s behavior, we need to develop both of these techniques. Our plan of attack will be to develop these equations in a CV and provide equations to change the reference frame afterwards. For a more complete treatment, the reader is referred to Refs. [11,12,25–28]. The basis for our development of the following equations are the tenets that (a) mass must be conserved and (b) Newton’s second law must hold true. Newton’s second law can be written as ∑ F = dt (mV) d (2.127) where ΣF is the vector sum of all the forces acting on a body (or blob of fluid or CV) m is the mass of the body V is the vector velocity of the body It is important to note that throughout this work, V is volume (a scalar), V is velocity (as a scalar quantity), and V is velocity (as a vector quantity). © 2014 by Taylor & Francis Group, LLC Ballistics: Theory and Design of Guns and Ammunition 50 Vt V CV Vt Vn V Vn FIGURE 2.16 Depiction of normal and tangential velocity components with respect to an arbitrary CV. Since CVs can be oriented in an arbitrary manner, it is important to understand that only that component of velocity normal to the control surface (CS) (i.e., the boundary of the CV) transports material or energy into the CV. If we examine Figure 2.16 where we have broken the velocity vectors into normal and tangential components (denoted Vn and Vt, respectively), we can clearly see why this is so. Consider an arbitrary property, N, of a substance. We would like to see how this property is transported into and out of a CV. If we define an intensive property, η, such that N m or N = η m ∫ ηρdV + ∫ ηρ V ⋅ dA + η= (2.128) Then we can write dN ∂ = dt ∂t CV outflow area ∫ ηρ V ⋅ dA (2.129) inflow area This equation defines how a property of interest is transported into and out of the CV. If we look at each of the terms, we see that this is an intuitively satisfying equation. The term on the LHS is the time rate of change (decrease) of any property of the CV over a time of interest. The first term on the RHS tells us how much of that property is stored in the CV over this time. The second term on the RHS tells us how much material has left the CV, while the third term tells us how much material has entered. Now wait a minute! If we look at the signs on the second and third terms, they seem to be incorrect—should not the stuff leaving have a negative sign and the stuff entering have a positive sign? The answer to this is yes, but Equation 2.129 is written correctly. The key to this seemingly inconsistent sign convention lies in the fact that the dot product in the second term is positive when we define the area as a vector that points outward and is normal to the surface. Similarly, the inflow term will always lead to a negative number since the velocity vector points inward and the area vector points outward. We shall now examine the flow of propellant gases in a suitable CV located somewhere behind a projectile at an instant in time. This will serve to foster understanding of the CV approach. Consider a CV in a gun tube located somewhere behind a moving projectile as depicted in Figure 2.17. There will be a velocity associated with the propelling gases (we will see this later) such that the gases are flowing in one side and out the other, but no gases flow through the walls. The ends of this cylindrical CV are designated as CSs. The inlet side is CS1 and the outlet side is CS2. If we would like to write an equation for how mass is transferred into or out of © 2014 by Taylor & Francis Group, LLC Physical Foundation of Interior Ballistics CS1 51 CS2 No flow through tube walls FIGURE 2.17 Typical gun tube CV (2D representation). this CV, we set N, the flow variable in Equation 2.129, equal to m, the property of interest. When this is done, Equation 2.128 tells us that η= N m = =1 m m (2.130) So for this case, we can write dm ∂ = dt ∂t ∫ ρdV + ∫ CV ρ V ⋅ dA + outflow area ∫ ρ V ⋅ dA (2.131) inflow area We know mass can neither be created nor destroyed, so dm/dt = 0, then we arrive at what is commonly called the equation of conservation of mass, or the continuity equation. In a general form, it is given as ∂ ∂t ∫ ρdV + ∫ CV ρ V ⋅ dA + outflow area ∫ ρ V ⋅ dA = 0 (2.132) inflow area The first term on the LHS states how the mass in the CV is changing with time. The second term is the amount of mass exiting the CV and the third term is the amount of mass entering the CV. The flow inside a gun tube is never steady or uniform. Nevertheless, it is informative to look at this expression using these two assumptions to gain some physical insight into the nature of the terms. The steady flow assumption means that there is no increase or decrease in material flow into or out of our CV. This implies that the first term is zero. So for the special case of steady flow, we have ∫ outflow area ρ V ⋅ dA + ∫ ρ V ⋅ dA = 0 (2.133) inflow area Simply put, this equation states that what comes into the CV equals what goes out of the CV. Uniform flow is a special case where fluid viscosity effects are neglected. This results in a constant velocity across the CSs. In essence, the velocity at the wall of the gun tube is the same as the velocity on the centerline of the tube. We will discuss this and its implications in more detail later. When we apply this assumption to Equation 2.133 and note that V · dA is negative at CS1 (because the vectors have opposite directions) and positive at CS2, we obtain the following simple relationship: ρ1V1 A1 = ρ 2V2 A2 = mɺ . Thus, under the steady flow assumption, the mass flow rate, m, is constant. © 2014 by Taylor & Francis Group, LLC (2.134) Ballistics: Theory and Design of Guns and Ammunition 52 We shall now examine the use of momentum, mV, as our flow variable. Use of Equation 2.128 with this flow variable yields η= N mV = =V m m (2.135) Now we can include this into Equation 2.129 to obtain d(mV ) ∂ = dt ∂t ∫ VρdV + ∫ CV Vρ V ⋅ dA + outflow area ∫ Vρ V ⋅ dA (2.136) inflow area Through Newton’s second law, we know that the term on the LHS (time rate of change of momentum) equals the forces on the system. The first term on the RHS is the change in the system’s momentum through storage in the CV. The second and third terms are the momentum leaving and momentum entering the CV, respectively. It is again informative to examine the steady flow case, which reduces our equation to ∫ F= Vρ V ⋅ dA + outflow area ∫ Vρ V ⋅ dA (2.137) inflow area Here we have replaced the time rate of change of momentum term with the force. Once again we shall use the uniform flow assumption to facilitate our understanding of this equation. Consider the same gun tube CV as earlier, drawn slightly differently in Figure 2.18. As discussed earlier, the velocity and area scalar products result in a negative sign on the inflow and a positive sign on the outflow side. With this uniform flow assumption (recall we also included a steady flow assumption to reduce the equation to the form of Equation 2.137), our Equation 2.137 would become F = ρ 2V2V2 A2 − ρ1V1V1 A1 (2.138) Note that this is still a vector equation with the vectors V1 and V2 determining the direction of F. If we had already worked out or it was obvious what direction the resultant force would be in, then we could write F = ρ 2V22 A2 − ρ 1V12 A1 (2.139) Equation 2.138 only tells us part of the story. It tells us the inertial reaction of the CV to the forces arising from a fluid passing through it. There are two types of forces that occur on the LHS in response to or independent of this, body forces and surface tractions. V2 V1 A1 CS1 A2 CS2 No flow through tube walls FIGURE 2.18 Typical gun tube CV (3D representation). © 2014 by Taylor & Francis Group, LLC Physical Foundation of Interior Ballistics 53 p2 p1 A1 A2 CS2 CS1 FIGURE 2.19 CV with no viscous forces acting. Body forces are those that act through the bulk of the material (i.e., directly affecting every molecule). Examples of this are gravitational loads, electromagnetic loads, etc. It is customary to write these loads on a unit mass basis to be consistent with the rest of the equation. In many cases, these are small and are neglected. Surface tractions are forces that act on the CS. These forces tend to be large and can be categorized into normal forces and shear forces. As the name implies, normal forces act normal to the CS. Pressure is the most common normal force. Because pressure cannot be negative, it always acts opposite to the surface area vector. Shear stresses are a result of the fluid’s propensity to stick to a solid (or other fluid) surface. The fluid viscosity, as defined earlier, is a measure of the intensity of these stresses. Shear stresses always act opposite to the direction of flow and along the CS. If a fluid is modeled as inviscid, there can be no shear stresses. Picking up from Equation 2.138, if we model a flow as steady with no viscosity, there will still be pressure forces present. This is depicted in Figure 2.19. Since pressure forces always act opposite to the area vector, it is customary to define the pressure forces as Fp = − ∫ outflow area pdA − ∫ pdA (2.140) inflow area In Equation 2.140, the signs of the area vectors would define the direction of the force. Before we establish a CV with viscous forces acting, it is instructive to describe these viscous forces and their effect on the flow field. As previously established, viscosity is a property of a fluid. The greater the viscosity of a fluid is, the more difficult it is to shear the material. If the viscosity is high enough or the flow velocity low enough, a fluid will exhibit what is known as laminar flow. Laminar flow is a very orderly shearing of the fluid from a solid surface where the fluid sticks to the boundary. In a tube or pipe, after some entrance length required for the flow to establish itself, the fluid will achieve a parabolic velocity distribution as depicted in Figure 2.20. V Tube wall FIGURE 2.20 Laminar velocity profile in a tube. © 2014 by Taylor & Francis Group, LLC 54 Ballistics: Theory and Design of Guns and Ammunition V Tube wall FIGURE 2.21 Uniform velocity profile in a tube. The laminar profile in Figure 2.20 is in stark contrast to the uniform profile that we had assumed in our previous discussions depicted in Figure 2.21. If the flow velocity is high enough or the viscosity low enough, the flow will transition from laminar flow to what is known as turbulent flow. Turbulent flow is characterized by a large number of eddies that swirl around in the flow. These eddies are important in that they tend to distribute momentum, energy, and matter throughout the fluid resulting in better mixing and very different transport properties. Many more flows are turbulent than laminar. The dimensionless parameter that governs this behavior is known as the Reynolds number and is given by Re = ρ Vd Vd = µ ν (2.141) where Re is the Reynolds number and is dimensionless ρ is the fluid density V is the fluid velocity d is a relevant characteristic length of the system (an internal diameter of a pipe, a length of a projectile, etc.) μ and ν are the dynamic and kinematic viscosities of the fluid, respectively If the Reynolds number is high enough, the flow will be turbulent. This demarcation is, in general, a range of values that also depends whether the flow is an internal one (such as the gas flow in a gun tube) or an external one (such as the flow about a projectile). The velocity profile of a turbulent flow is depicted in Figure 2.22. Here we can see that the effect of the eddies is to distribute the momentum, resulting in a profile that is flatter and more akin to our inviscid flow model of Figure 2.21. If we now return to our discussion on the surface tractions, we can discern that the effect of fluid viscosity is to create a shear stress at the boundary between the fluid inside a gun V Tube wall FIGURE 2.22 Turbulent velocity profile in a tube. © 2014 by Taylor & Francis Group, LLC Physical Foundation of Interior Ballistics 55 p2 A2 p1 A1 τw CS1 CS2 FIGURE 2.23 Surface tractions on a gun tube CV. tube and the solid tube itself (i.e., on our CS). If we consider the diagram of Figure 2.19, we can redraw this figure to include the effect of shear stresses as depicted in Figure 2.23. Since the shear stress, τw, acts all over the area of our CV, we can add a term in for this into Equation 2.140 to obtain an expression for all of the surface forces as follows: Fsurface = − ∫ ∫ pdA − outflow area ∫ pdA − inflow area τ wdA (2.142) surface area We can insert this into our expression for the conservation of momentum Equation 2.137 to obtain, for steady flow − ∫ ∫ pdA − outflow area pdA − inflow area ∫ ∫ τ wdA = surface area Vρ V ⋅ dA + outflow area ∫ Vρ V ⋅ dA inflow area (2.143) or, in a more general sense, ∫ − pdA − outflow area = ∂ ∂t ∫ pdA − ∫ τ wdA inflow area surface area ∫ VρdV + ∫ Vρ V ⋅ dA + CV outflow area ∫ Vρ V ⋅ dA (2.144) inflow area The next transport property we shall examine is that of energy. In Sections 2.4 and 2.5, this was discussed to a degree. The objective of this section is to demonstrate that we can use the same transport Equation 2.129 to come up with the energy equations we have used earlier. We start by recognizing that our transport variable is energy, E. With this in mind, Equation 2.128 can be rewritten as η= E =e m (2.145) Recall that lower case letters are intensive properties. Then we can write dE ∂ = dt ∂t ∫ eρdV + ∫ CV eρ V ⋅ dA + outflow area ∫ eρ V ⋅ dA (2.146) inflow area This states that the change in energy of a system is equal to the change in energy stored in the system minus that which is advected away plus that which is advected into the system. Recall from Equation 5.6 that dQ dW dE + = dt dt dt © 2014 by Taylor & Francis Group, LLC (2.147) Ballistics: Theory and Design of Guns and Ammunition 56 From our definition of work, we know that W= ∫ pdV (2.148) But volume is nothing more than a length times an area. This allows us to write ∫ px ⋅ dA W= (2.149) If we take the derivative of this expression with respect to time assuming pressure is an average value over the time increment, we can write dW = dt dx ∫ p dt ⋅ dA = ∫ pV ⋅ dA (2.150) There are many types of work terms. The aforementioned term happens to be called pdV work or pressure work. The other types of work, such as shaft work, are usually not present in a gun launch so we shall neglect them. Insertion of Equation 2.150 into Equation 2.146 and rearranging yields dQ ∂ = dt ∂t   p p  e +  ρ V ⋅ dA  e +  ρ V ⋅ dA + ρ ρ     inflow area outflow area ∫ ∫ eρdV + ∫ CV (2.151) In the section on thermodynamics, we defined the specific energy through Equation 2.39. If we insert this definition into the aforementioned expression, we obtain dQ ∂ = dt ∂t ∫ eρ dV + CV  p V2 + u +  ρ V ⋅ dA  gz + 2 ρ  CS ∫ (2.152) Here we have combined the last two terms on the RHS of Equation 2.151 with the understanding that the integral of the last term in Equation 2.152, being an integral over the entire CS, accounts for the difference between inflow and outflow. It is informative to look at this equation with respect to a gun launch. The term on the LHS represents the transfer of heat to or from the system. The first term on the RHS represents the change in stored energy of the system (such as energy released by propellant combustion). The last term on the RHS is the change in energy of the system. Since gravitational potential energy, the product gz, is small relative to the other energy terms, it is usually neglected allowing us to rewrite the expression as dQ ∂ = dt ∂t ∫ CV eρ dV +  V2 p + u +  ρ V ⋅ dA  ρ  2 CS ∫ (2.153) Earlier in this section, we introduced the common practice of characterizing a fluid based on its behavior under shear stress. This allowed us to come up with a relationship between applied © 2014 by Taylor & Francis Group, LLC Physical Foundation of Interior Ballistics 57 shear stress and deformation rate. Another distinction has to be made between fluids with respect to the density. If the density is considered constant in a fluid or solid that we model, we call this material incompressible. If the density varies, we must analyze the problem with the assumption of compressible material. This has many ramifications. The most significant ramification is that if the material is incompressible, then the energy equation is decoupled from the momentum equation and we can solve them independently [25]. This makes problem solving much simpler. We do not have this luxury when the density varies significantly. In fluid flows, such as those that we shall study later, a dimensionless parameter known as the Mach number, Ma, is used as a measure to determine the effect of compressibility, among its other uses. The Mach number is given by Ma = V a (2.154) where V is some characteristic velocity in the material a is the speed of sound in the material In general, if the Mach number is below 0.3, the deviation from incompressible flow is small so the assumption of incompressibility leads to an acceptably small error [16]. In an ideal gas, the speed of sound is given by the relation a = γ RT (2.155) where γ is the specific heat ratio R is the specific gas constant T is the absolute temperature (i.e., in degrees Rankine or Kelvin) The speed of sound in any material is formally defined as a= ∂p ∂ρ (2.156) s That is to say that the speed of sound in a material is equal to the square root of the partial derivative of pressure with respect to density evaluated with constant entropy. The interested reader is referred to any of Refs. [15,16,25] for the detailed proof of this equation. The speed of sound is essentially the fastest speed at which a disturbance can be propagated by molecular interaction. If a disturbance is created that is strong enough, a shock will form. This shock must always move faster than the speed of sound in the material. We will discuss this in detail later. In the study of compressible flows, it is common practice to utilize stagnation values in many of our calculations. Stagnation values are the values of the enthalpy, pressure, temperature, and density that are achieved by adiabatically slowing a flow down to zero velocity. The assumption of adiabatic behavior is warranted in many of the situations we will examine, particularly in exterior ballistics. The stagnation enthalpy is given by 1 h0 = h + V 2 2 © 2014 by Taylor & Francis Group, LLC (2.157) Ballistics: Theory and Design of Guns and Ammunition 58 In this and the following equations, the subscript “0” indicates the stagnation value, V is the velocity of the flowing fluid, and the values without the subscript are the static value, in the case of Equation 2.157, h is the static enthalpy. Equation 2.157 holds for any material. If the material is an ideal gas, we can define the stagnation temperature, pressure, and density as T0 = h + 1 V2 2 cp or γ −1 T0 = 1+ Ma 2 2 T (2.158) γ p0  γ −1  γ −1 Ma 2  = 1+ p  2  (2.159) 1 ρ0  γ − 1  γ −1 = 1+ Ma 2  ρ  2  (2.160) In each of these cases, thermodynamic relations have been used for an ideal gas (Equation 2.61). Shock waves are formed in materials when disturbances of sufficient strength propagate through the medium. “Sufficient” strength is a term that we throw about rather loosely to describe conditions where shocks are formed—it can be cast in terms of flow velocities or pressures (the two are linked as we shall see). Shocks can be classified as normal or oblique, depending upon the direction of material flow into them. They can also be analyzed as steady or transient. In general, shocks can take curved and rather complex shapes, but the simple analytical tools we have allow us to look at them only under simplified geometries. More complex geometries require the assistance of a computer. We shall only examine normal shocks in this brief review and direct the reader to Ref. [16] for the handling of oblique shocks. The best way to examine the behavior of a shock is to look at a shock tube. This simple device will allow us to introduce all of the material necessary for the introductory study of ballistics and set the stage for later work when we discuss stress waves in solids. Before we look at a shock tube, we need to discuss the principle of superposition as applied to shock waves. Consider two shocks as depicted in Figure 2.24. One of these cases is a stationary shock where we could consider ourselves “riding on the wave,” while in the other case, we can consider ourselves to be sitting on the ground watching the shock pass by. If, in both cases, the shock were passing into a stagnant medium, we would see some important correlations. The passage of a shock wave always induces motion that follows the wave. Consider the situation where we are sitting on the ground, the air about us is V2 Gas motion downstream V1 up > 0 Fluid in motion behind shock Gas motion upstream (a) FIGURE 2.24 (a) Stationary and (b) moving shock waves. © 2014 by Taylor & Francis Group, LLC x (b) U Stagnant gas ahead of shock x Physical Foundation of Interior Ballistics 59 stagnant and all of a sudden a shock passed by us just as is shown in Figure 2.24b. If the shock were moving at velocity, U, we would feel an induced motion, a wind, immediately afterwards moving at velocity up in the same direction that the shock was moving. If we experienced this same situation but instead were riding on the shock, we would feel a wind of velocity U coming toward our face. This would be analogous to the situation in Figure 2.24a. In this situation, the velocity V1 would be equal to U. Note the direction of the velocity vectors in the figure. The velocity vector of magnitude V2 is moving away from the wave. The figure is drawn correctly, but in the case that was just described, based on superposition, since U is larger than up (and it always is). If we were riding on the wave, we would see material leaving us at velocity (U − up). When we examine a shock wave in the frame of Figure 2.24b, we are said to be using an Eulerian frame of reference. If we analyze the very same situation as shown in Figure 2.24a, we are using a Lagrangian reference frame. The difference between Lagrangian and Eulerian reference frames is important because we sometimes prefer to solve a problem in one frame or the other because the mathematics are simpler. As long as the reference frame motion is accounted for, solving in one frame or the other leads to the same answer. We shall now use the Lagrangian approach to examine the governing equations for a stationary normal shock wave. Consider the situation in Figure 2.25 where a shock wave is moving to the left at velocity, U. Since we would like to examine the behavior of this shock, we will put ourselves in a reference frame attached to the shock itself. We form a CV enclosing the shock only. We observe, while riding on this shock, that fluid enters the CV at velocity U and leaves at velocity u2. We can write the conservation of mass, momentum, and energy equations for this system as follows: Conservation of mass (continuity equation) ρ1U = ρ 2u2 (2.161) p1 + ρ1U 2 = p2 + ρ 2u22 (2.162) 1 1 h1 + U 2 = h2 + u22 = h01 = h02 = h0 = constant 2 2 (2.163) Conservation of momentum Conservation of energy We see from the last equation that across a shock wave the stagnation enthalpy must remain constant. This falls out directly from the fact that we assumed the shock wave FIGURE 2.25 Stationary shock wave. © 2014 by Taylor & Francis Group, LLC Ballistics: Theory and Design of Guns and Ammunition 60 FIGURE 2.26 Stationary shock wave moving into a stagnant fluid. was adiabatic. These equations are coupled through a material model such as the ideal gas equation of state (relates p, V, and T) and the calorically perfect assumption (relates h to T). If we consider the special case where the shock under examination is moving into a stagnant fluid as depicted in Figure 2.26, we can write the aforementioned three equations as ρ1U = ρ 2 (U − up ) (2.164) p1 + ρ1U 2 = p2 + ρ 2 (U − up )2 (2.165) 1 1 h1 + U 2 = h2 + (U − up )2 = h0 = constant 2 2 (2.166) The conservation of mass, momentum, and energy equations can be combined as detailed in Refs. [10,16] to yield the Rankine–Hugoniot relationship. This relationship determines how the energy changes across a normal shock wave. It is very important and will appear again when the terminal ballistics material is discussed. It can be written in terms of total specific energy, e, or if some of the energy components are negligible, it can be written in terms of enthalpy, h. At this stage, we will use the latter expression, but we shall switch when we discuss shock in the terminal ballistics section. Writing the Rankine–Hugoniot relationship in terms of enthalpy, we have h2 − h1 =  1 1 1  ( p2 − p1 )  −  2 ρ ρ 2 1   (2.167) The strength of a shock is normally assessed by the change in pressure across it. In other words, its strength is given by the ratio p2/p1. If we assume the material through which this shock is propagating is an ideal gas, Equations 2.164 through 2.166 can be combined with the relationships provided in Equation 2.61 to yield expressions that relate all of the values ahead of the shock to values after the passage of the shock. The details of this are available in Ref. [16]. These expressions are as follows:  γ + 1 p2  +   T2 p2  γ − 1 p1  = T1 p1  γ + 1  p2   1+   γ − 1  p1    © 2014 by Taylor & Francis Group, LLC (2.168) Physical Foundation of Interior Ballistics ρ1 = ρ2 61 γ + 1  p2    γ − 1  p1  γ + 1 p2 + γ − 1 p1 1+ (2.169) The real power of these equations lies in the fact that with just the strength of the shock known we can determine all of the other items of interest. In the aforementioned equations, we have seen that given the pressure ratio (i.e., the strength) of the shock, we know the temperature behind the wave and the increase in density across the wave. We can also determine the wave speed, U, and induced velocity, up, through U = a1  γ + 1  p2  − 1 + 1 2γ  p1  (2.170) 2γ   ρ1  a1  p2 γ +1 up = U  1 −  =  − 1  ρ 2  γ  p1   γ − 1 + p2 γ + 1 p1 (2.171) If we change reference frames to one in which we are stationary and the shock is moving, then the assumption of constant stagnation enthalpy, h0, is no longer valid. The reason is best illustrated by an example. Consider the gas ahead of the shock wave. It was initially motionless so h1 = h01. After the wave passes, we know that the temperature must increase so h2 > h1. Additionally, the gas is now moving at velocity, up, so that we can see h1 = h01 < h02 = h2 + 1 2 up 2 (2.172) It is by this very same logic that the stagnation pressure, temperature, and density must also increase. We have discussed some governing equations but let us break for a moment to discuss why a gas shocks up. If we examine Equation 2.170 closely, we see that a higher pressure causes a faster motion of the wave. If we imagine a shock wave as depicted in Figure 2.27 moving to the right, we can pick out three points that we shall follow for some time. Point A is essentially the beginning of the pressure increase and at the un-shocked initial pressure. p C C B C B A B A A t, x FIGURE 2.27 Formation of a shock wave. © 2014 by Taylor & Francis Group, LLC Ballistics: Theory and Design of Guns and Ammunition 62 Point B is at some pressure in between the peak pressure of the shock and the initial pressure of the material into which the shock is propagating. Point C is at the peak shock pressure. From Equation 2.170, we see that the local velocity of point B must be greater than point A and also that the local velocity of point C is greater still. This means that at some time, t, these points must converge thereby forming a step discontinuity in pressure. This step discontinuity is the way we model the shock—there is actually a very small distance over which a shock will develop so that the pressure increase is rapid, but continuous. With this information, we see that compression shocks are the only admissible shocks. Later we will introduce rarefactions that are the converse of shocks. Since the pressure decreases in a rarefaction wave, the wave will tend to spread out over time and distance. Now that we have the governing equations, we shall examine the behavior of a shock wave in a shock tube. A shock tube is a device as depicted in Figure 2.28 that contains two regions of gas. These regions are separated by a diaphragm that can be burst very quickly and contain one gas at high pressure and another at lower pressure. The gases could be different (thus all of their properties as well) as can their temperatures. Below the graphic of the shock tube is a pressure versus distance plot showing that the pressure in region 4 (the high-pressure region) is greater than that of region 1 and the diaphragm divides the two regions. If the diaphragm is burst, then a shock will propagate into the lower pressure region, increasing the pressure, and a rarefaction wave (to be discussed later) will propagate into the high-pressure region, decreasing the pressure. If we examine the shock tube after some very short time, t, the situation will appear as shown in Figure 2.29 with the corresponding pressure–distance profile. One of the most interesting aspects of compressible fluid flow is that if we know what the initial states of the ideal gases in the p4, T4, a4, γ4 p1, T1, a1, γ1 4 Diaphragm p4 p 1 p1 FIGURE 2.28 The shock tube in its initial state. (From Anderson, J.D., Modern Compressible Flow with Historical Perspective, 3rd edn., McGraw-Hill, New York, 2003. With permission.) up 4 2 3 U 1 Diaphragm burst p p4 p3 = p2 p1 x FIGURE 2.29 The shock tube after some short time, t. (From Anderson, J.D., Modern Compressible Flow with Historical Perspective, 3rd edn., McGraw-Hill, New York, 2003. With permission.) © 2014 by Taylor & Francis Group, LLC Physical Foundation of Interior Ballistics 63 shock tube are, we can predict the pressures and temperatures of the unsteady motion afterwards by Equations 2.161 through 2.171. In fact, we can predict the pressure behind the initial shock from −2γ 4   ( γ4 − 1)   a  p   (γ 4 − 1)  1   2 − 1    a p p 4 p2   4  1  = 1 −  p1 p1    p2   2γ 1  2γ 1 + (γ 1 + 1)  − 1       p1     (2.173) Equation 2.173 needs to be solved for the initial shock strength, p2/p1, but afterwards Equations 2.161 through 2.171 can be used directly to calculate the parameters of interest. The details of this derivation can be found in Ref. [16]. You can see from Figures 2.28 and 2.29 that the shock tube is not infinite in extent. At some point, the shock produced by the bursting of the diaphragm will reach the right end of the tube. When this occurs, the condition at the wall is such that no flow through it is possible. Consider that all the fluid behind the shock wave is moving with induced velocity, up, toward the wall. Clearly, this situation is at odds with the wall-imposed boundary condition of zero velocity. Nature handles this issue by creating a shock wave of strength UR that propagates back into the fluid that is heading toward the wall at velocity up. Notice that we have used a velocity here to define the strength of the shock—it should be clear by now that if we know either the velocity of the shock or the pressure ratio, we can find the other. The net effect of this reflected shock is that it stagnates the fluid between it and the fixed end of the shock tube as depicted in Figure 2.30. In this figure, we shall assume the tube is extremely long on the rarefaction side so we do not have to discuss rarefaction reflections, yet. If we look at our conservation Equations 2.161 through 2.163 and consider that the shock wave sees material coming into it at velocity UR + up, we can write equations for the reflected shock that are analogous to Equations 2.164 through 2.166 for the incident wave. These are ρ 2 (U R + up ) = ρ 5U R (2.174) p2 + ρ 2 (U R + up )2 = p5 + ρ 5U R2 (2.175) 1 1 h2 + (U R + up )2 = h5 + U R2 2 2 (2.176) up 4 3 2 UR 5 u5 = 0 p4 p p3 = p2 p5 x FIGURE 2.30 The shock tube after a reflection of the incident wave. © 2014 by Taylor & Francis Group, LLC Ballistics: Theory and Design of Guns and Ammunition 64 A simple method for determination of the speed of the reflected shock is to first determine the Mach number of the incident pulse, Mas, through Mas = U a1 (2.177) The relationship between the incident shock velocity and the reflected velocity is derived in Ref. [16] and given by MaR Mas 2(γ − 1) ( Mas2 − 1)  γ + 1 2  = 1+ 2 2 2 MaR − 1 Mas − 1 Mas  (γ + 1)  (2.178) Here MaR is the Mach number of the reflected shock that can be converted to a velocity through use of MaR = U R + up a2 (2.179) In our discussions on shock waves throughout the terminal ballistics sections, we will make use of time–distance diagrams, so-called t–x plots. It is prudent to introduce them here as reinforcement of the shock wave discussion. An t–x plot places distance on the abscissa and time on the ordinate. Because of this placement, which is opposite to normal function versus time plots, we need to adjust some of our logic that we are used to. For instance, slopes of straight lines on these diagrams are reciprocal velocities. If we consider the situation in Figure 2.30 and draw an t–x plot for it with the origin starting from the initial diaphragm location, we would have a plot as depicted in Figure 2.31. We shall examine the shocks in this diagram first. If we assume that the incident shock forms immediately (this is not really true, as we learned earlier, but close enough for our purposes), it propagates toward the wall that is located at point x2 in our figure. If we wanted to determine what the velocity distribution was t u5 = 0 t3 Reflected shock, slope = 1/UR Particle path, slope = 1/up t2 x = 0 is diaphragm location Wall t1 Incident shock, slope = 1/U u1 = 0 x1 x3 x2 x FIGURE 2.31 t–x plot for the reflection of a shock wave. (From Anderson, J.D., Modern Compressible Flow with Historical Perspective, 3rd edn., McGraw-Hill, New York, 2003. With permission.) © 2014 by Taylor & Francis Group, LLC Physical Foundation of Interior Ballistics 65 in this device at any time, t, we would examine a horizontal line in the figure. For instance, if we examined the situation at time, t1, we would see that the material in the unshaded region up to point x1 would have a velocity up and everything between x1 and x2 (the wall) would have zero velocity. Once the incident shock reflects off the wall, a new shock of velocity UR propagates back into the fluid. This is depicted by the upper line in the diagram. Note that the slope is greater on this reflected shock, indicative of a lower velocity than the incident wave. The material in the shaded region behind this wave has been stagnated to zero velocity. We can use an t–x diagram to determine how a particle moves over time. Consider a particle located initially at location x1. It remains stationary until the shock wave passes by at time t1, as indicated by a vertical line. At time t1, the incident shock passes it and induces a velocity, up to the particle. When the particle moves at velocity, up, it will trace out a line on the diagram that has a slope of 1/up. While this particle is moving at velocity up, the shock interacts with the wall and reflects at time t2. While the reflected shock is approaching, the observed particle has no idea anything is about to happen and continues to move at velocity up until the reflected shock passes by at time t3. This passage of the reflected shock stagnates the particle to zero velocity and its motion (or lack thereof) traces out a vertical line. A final point of interest regarding t–x plots is that we can actually see the compression of the material. If we consider all of the material initially between points x1 and x2, we see that, after the passage of the shock and its reflection, it has all been compressed to the region between x3 and x2. With this information, the basis for our future discussions using t–x plots is established. A rarefaction wave, sometimes known as an expansion or relief wave, is the means by which nature handles a sudden drop in pressure. As we stated earlier, compression waves (also known as condensations) eventually coalesce into shocks that are analyzed as step discontinuities in pressure. This coalescence was brought about by the fact that the local velocity increases with increasing pressure. In a rarefaction, the opposite is true. A rarefaction increases over time because the pressure at the head of the wave is greater than that at the tail of the wave. In the case of our shock tube, the head of the rarefaction will propagate at the local speed of sound in the material (a4 in Figure 2.29), while the tail will propagate at velocity (u3 − a3) that is equal to (up − a2). This is depicted schematically in Figure 2.32. Tail slope = 1/UT = 1/(u3–a3) t 3 u = u3 a = a3 x = 0 is diaphragm location Particle path, slope = 1/u3 = 1/up Head slope = 1/UH = 1/(u4–a4) = –1/a4 3 4 u = u4 = 0 a = a4 4 x FIGURE 2.32 (See color insert.) t–x plot for a rarefaction wave. © 2014 by Taylor & Francis Group, LLC 66 Ballistics: Theory and Design of Guns and Ammunition Throughout the rarefaction wave, the velocity continuously decreases between these two values. Because of this continuous decrease in velocity, it is common to model the decrease as a series of wavelets. The more wavelets we include, the smoother the curve. If we use Figure 2.32 to trace a particle path after the bursting of the diaphragm, we see that the particle would not move until the head of the rarefaction wave passed by it. After the passage of the head of the wave, the velocity would continuously increase until passage of the tail of the wave, after which it would be moving at velocity up. The length of the rarefaction can be determined at any time by scribing a horizontal line through the diagram. If we do this at two points in time on the diagram, we can see how the length of the wave increases. What is depicted in Figure 2.32 is a simple, centered rarefaction wave. A wave is considered simple if all of the characteristics (the rays emanating from the origin) are straight. Reflections of a rarefaction are somewhat more complicated than that of a shock. The reflection of the head of the rarefaction wave must pass through the characteristics of the rest of the wave being both affected by as well as affecting them. The result is that the characteristics tend to bend making the calculations somewhat more complex. We will handle this in a simplified fashion later, but the interested reader is directed to Ref. [16] for an outstanding treatment for handling these situations. We now have sufficient information to handle the fluid mechanics of interior and exterior ballistics. We shall treat the formation of shocks and rarefactions as necessary in the terminal ballistics section. Problem 18 The principle behind a muzzle brake on a gun is to utilize some of the forward momentum of the propelling gases to reduce the recoil on the carriage. In the simple model below (Figure 2.33), the brake is assumed to be a flat plate with the jet of gases impinging upon it. If the jet diameter is 105 mm and the velocity and density of the gas (assume air) are 750 m/s and 0.457 kg/m3, find the force on the weapon in Newtons assuming the gases are directed 90° to the tube and the flow is steady. Answer: −2225.9 [N] Problem 19 Some engineer gets the idea that if deflecting the muzzle gases to the side is a good idea, then deflecting it rearward would be better (until of course an angry gun crew gets hold of him). If the jet diameter is again 105 mm and the velocity and density of the gas (again assume air) are 750 m/s and 0.457 kg/m3, find the force on the weapon in Newtons assuming the gases are directed 150° to the tube and the flow is steady (Figure 2.34). Answer: −4153.5 [N] F FIGURE 2.33 Normal deflection of flow through a muzzle brake. © 2014 by Taylor & Francis Group, LLC Physical Foundation of Interior Ballistics 67 F 30° (typ.) FIGURE 2.34 Rearward deflection of flow through a muzzle brake. Problem 20 Consider a shock tube that is 6 ft long with a diaphragm at the center. Air is contained in both sections (γ = 1.4). The pressure in the high-pressure region is 2000 psi. The pressure in the low pressure region is 14.7 psi. The temperature in both sections is initially 68°F. When the diaphragm is burst, determine the following: 1. The velocity that the shock wave propagates into the low–pressure region. Answer: 2798 [ft/s] 2. The induced velocity behind the wave. Answer: 1946 [ft/s] 3. The velocity of a wave reflected normally off the wall (relative to the laboratory). Answer: 1232 [ft/s] 4. The temperature behind the incident wave. Answer: 657 [°F] 5. Draw an t–x diagram of the event. Include the path of a particle located 2 ft from the diaphragm. Problem 21 An explosion generates a shock wave in still air. Assume we are far enough from the initial explosion that we can model the wave as a one-dimensional shock. Assume that the pressure generated by the explosion was 10,000 psi and the ambient atmospheric pressure, density, and temperature are 14.7 psi, 0.06 lbm/ft3 and 68°F, respectively. Determine 1. The static pressure behind the wave (assume γ = 1.4 and since we are far away from the effects of the explosion assume a1/a4 ≈ 0.5) Answer: p2 = 376.6 [psi] 2. The velocity that the wave propagates in still air Answer: U = 5294 [ft/s] 3. The induced velocity that a building would see after the wave passes Answer: up = 4212 [ft/s] 4. The velocity of a wave reflected normally off a building Answer: UR = 1921 [ft/s] © 2014 by Taylor & Francis Group, LLC 68 Ballistics: Theory and Design of Guns and Ammunition References 1. Wark, K., Thermodynamics, 5th edn., McGraw-Hill, New York, 1988. 2. Jones, L. and Atkins, P., Chemistry, Molecules, Matter, and Change, 4th edn., W.H. Freeman and Co., New York, 2003. 3. Masterson, W., Slowinski, E., and Stanitski, C., Chemical Principles, 6th edn., Saunders College Publishing, Philadelphia, PA, 1985. 4. Beer, F.P. and Johnson, R., Mechanics of Materials, 2nd edn., McGraw-Hill, New York, 1992. 5. Van Wylen, G.J. and Sonntag, R.E., Fundamentals of Classical Thermodynamics, 3rd edn., John Wiley and Sons, New York, 1986. 6. Fermi, E., Thermodynamics, Dover Publications, New York, 1956. 7. Corner, J., Theory of the Interior Ballistics of Guns, John Wiley and Sons, New York, 1950. 8. Cooper, P.W., Explosives Engineering, Wiley-VCH Inc., New York, 1996. 9. Sucec, J., Heat Transfer, William C. Brown Publishers, Dubuque, IA, 1985. 10. Kays, W.M. and Crawford, M.E., Convective Heat and Mass Transfer, 3rd edn., McGraw-Hill, New York, 1993. 11. Fox, R.W. and McDonald, A.T., Introduction to Fluid Mechanics, 4th edn., John Wiley and Sons, New York, 1992. 12. Anderson, J.D., Modern Compressible Flow with Historical Perspective, 3rd edn., McGraw-Hill, New York, 2003. 13. Cengel, Y.A. and Boles, M.A., Thermodynamics and Engineering Approach, 4th edn., McGraw-Hill, New York, 2002. 14. Moran, M.J. and Shapiro, H.N., Fundamentals of Engineering Thermodynamics, 5th edn., John Wiley and Sons, New York, 2004. 15. Masterson, W.L., Slowinski, E.J., and Stanitski, C.L., Chemical Principles, 5th edn., Saunders College Publishing, Philadelphia, PA, 1981. 16. Cooper, P.W. and Kurowski, S.R., Introduction to the Technology of Explosives, Wiley-VCH, New York, 1996. 17. Hayes, T.J., Elements of Ordnance, John Wiley and Sons, New York, 1938. 18. Eringen, A.C., Liebowitz, H., Koh, S.L., and Crowley, J.M., Eds., Mechanics and Chemistry of Solid Propellants, Proceedings of the Fourth Symposium on Naval Structural Mechanics, Pergamon Press, London, U.K., 1965. 19. Kubota, N., Propellants and Explosives, Wiley-VCH, New York, 2002. 20. Turns, S.R., An Introduction to Combustion, 2nd edn., McGraw-Hill, New York, 2000. 21. Borman, G.L. and Ragland, K.W., Combustion Engineering, WCB-McGraw-Hill, New York, 1998. 22. Yang, V., Brill, T.B., and Ren, W.-Z., Solid propellant chemistry, combustion, and motor interior ballistics, Progress in Astronautics and Aeronautics, Vol. 185, American Institute of Astronautics and Aeronautics, Reston, VA, 2000. 23. Kuhl, A.L., Leyer, J.C., Borisov, A.A., and Sirignano, W.A., Dynamics of deflagrations and reactive systems flames, Progress in Astronautics and Aeronautics, Vol. 131, American Institute of Astronautics and Aeronautics, Washington, DC, 1989. 24. Kuhl, A.L., Leyer, J.C., Borisov, A.A., and Sirignano, W.A., Dynamics of gaseous combustion, Progress in Astronautics and Aeronautics, Vol. 151, American Institute of Astronautics and Aeronautics, Washington, DC, 1993. 25. White, F.M., Fluid Mechanics, 5th edn., McGraw-Hill, New York, 2003. 26. Panton, R.L., Incompressible Flow, 2nd edn., John Wiley and Sons, New York, 1995. 27. Currie, I.G., Fundamental Mechanics of Fluids, 2nd edn., McGraw-Hill, New York, 1993. 28. White, F.M., Viscous Fluid Flow, 3rd edn., McGraw-Hill, New York, 2006. © 2014 by Taylor & Francis Group, LLC 3 Analytic and Computational Ballistics Chapter 2 has provided us with the necessary background to discuss procedures that calculate the behavior of projectiles and propellant in the gun tube. The chapter had to be brief because detailed treatment of any one of the subjects could be (and are) collected into complete texts in their own right. The reader is directed to the references at the end of the chapter if a more complete background in the individual subject is felt to be necessary. Much like other introductory texts on difficult subjects, this chapter shall begin with fundamental treatments that will allow the reader to perform meaningful calculations of interior ballistic problems. This simplified treatment will, by its very nature, not provide exact answers but answers that are reasonable from an engineering viewpoint. As will be discussed, more exact methods require a varying degree of computer assets. 3.1 Computational Goal The interior ballistician is charged with devising a propellant charge that will deliver the projectile of interest to the gun muzzle intact, with the desired muzzle velocity, with no damage to the weapon from excess pressure, and with high probability that successive charges propelling the same projectiles will produce the same results. To do this, the ballistician must be able to predict a priori what the charge will do, i.e., what pressures will both the gun and the projectile experience during travel down the bore and what the velocity and acceleration profile would be during the travel to the muzzle. Over the centuries, ballisticians, including some quite eminent mathematicians and physicists, have devised computational schemes that can be used to make such predictions. We intend to explore a few of these analytic tools in sufficient depth so that the physics and mathematics become clear to the user, who would then also be able to discern reasonable answers from patently erroneous ones. It is important to understand how predictions of pressure and velocity are verified experimentally in real guns. Such understanding has led to the development of pressure ratios that allow the gun and projectile designers to know what pressures are acting on the gun and on the projectile at locations that practical instrumentation has some difficulty capturing. Pressure is most readily measured at the base of the gun chamber, where the gas flow is minimal or nonexistent. When pressure taps are introduced along the bore to take measurements while the projectile is traveling and the gases are flowing, it has been found that turbulent flow and shock waves make such measurements difficult to interpret. Copper crusher gauges are used in which small copper cylinders are crushed to a barrel shape in the gauge by the applied pressure and the distortion of the cylinders measured. These gauges are placed in the base of the charge and recovered after firing. Distortion is checked against a calibration chart and the pressure is quickly read. Of course, pressure measured in this way is representative only of the maximum pressure sensed by the gauge, which gives no indication of its profile in time or in travel. 69 © 2014 by Taylor & Francis Group, LLC Ballistics: Theory and Design of Guns and Ammunition 70 Even such a primitive measurement was and still is of use; because the designer would know the maximum pressure, the projectile and gun would have to contend with an indication that piezo-type pressure gauges are functioning properly. These gauges are still widely used to check the pressure consistency of already developed charges. Knowledge of how that copper pressure was related to pressures at other locations during the travel was a great advance. When the pressure ratios were devised that related chamber pressure to the pressure at the base of the projectile during its travel down the bore, these were greatly appreciated by the designers. Even better was the introduction of electronic piezo gauges installed through the breech that allowed the measurement of pressure over time so that a pressure–time profile could be available. The study of a few of the computational theories that develop these ratios follows in succeeding sections. 3.2 Lagrange Gradient To determine the time-dependent motion of the projectile, we need to make some assumptions about the behavior of the gas pushing it out of the gun. These assumptions will involve the pressure, mass, and density distribution of the gas (see Figure 3.1). We shall continue to use x as the distance from the projectile base position at the seating location to its position at all later times with the time derivative defined as dx = xɺ = V dt (3.1) xpmax p pmax pmuz x=L pB p FIGURE 3.1 Pressure–distance relationship in a typical gun firing. © 2014 by Taylor & Francis Group, LLC pS x Analytic and Computational Ballistics 71 We will first assume that the gas density is uniform in the volume behind the projectile at time t. We can then write, for any time, t, that ρ = ρ ( xg , t ) (3.2) In this equation, xg is the x-location of the gas mass center behind the projectile. We shall also assume that there is no spatial gradient in density at any time, thus ∂ρ =0 ∂xg t (3.3) We can also write the continuity equation for a compressible fluid as ∂ρ ∂ ρVx g = 0 + ∂t ∂xg ( ) (3.4) We can expand the continuity equation (3.4) as ∂Vx g ∂ρ ∂ρ Vx g + ρ + =0 ∂t ∂xg ∂xg (3.5) Inserting our assumption of the absence of a spatial density gradient allows us to simplify this expression to ∂Vx g ∂ρ +ρ =0 ∂t ∂xg (3.6) Now because we stated that the density was not a function of x, we can remove the partial derivative notation from the temporal term and rearrange to yield ∂Vx g 1 dρ =− ρ dt ∂xg (3.7) Assume at this point that the solid propellant in the charge has all turned to gas, then what was initially a solid propellant of charge weight, c, is now a gas of identical weight, c. So the gas density is this weight divided by the volume the gas occupies, or ρ (t)|c = c V(t) (3.8) where the subscript “c” refers to conditions after the charge has burned out, i.e., all the solid has evolved into gas. If the base of the projectile has moved a distance, x, and the bore area is A, then the volume behind the projectile containing gas is V(t) = Ax(t) © 2014 by Taylor & Francis Group, LLC (3.9) 72 Ballistics: Theory and Design of Guns and Ammunition If we insert Equation 3.9 into Equation 3.8 and then take the derivative with respect to time, the result can be simplified to Equation 3.10: 1 dx ∂Vx g = x dt ∂xg (3.10) Note that there is a difference here between x and xg: x is the location of the base of the projectile xg is the location of the mass center of the gas If we integrate Equation 3.10 with respect to xg and use the boundary conditions of Vxg = 0 when xg = 0, then we get xg dx = Vxg ( xg ) x dt (3.11) Now, since x is the position of the base of the projectile at time t we see that dx/dt is the velocity of the projectile at time t, so we can write V Vxg = x xg (3.12) This implies that the gas particle velocity varies linearly from the breech face to the projectile base, and is a fundamental tenet of the Lagrange* approximation. We can describe the kinetic energy of the gas stream as KEg = 1 mgVx2g 2 (3.13) But, as described earlier, the mass of the gas is its density times the volume it occupies at time t, therefore x KEg = 1 ∫ 2 ρ AV 2 xg dxg (3.14) 0 Moving the spatially constant terms, ρA/2, outside the integral and performing the integration gives us ρ A V 2 xg3 KEg = 2 x2 3 x = 0 1 ρ AxV 2 6 (3.15) But we know from our earlier work that ρ Ax = c (3.16) So we can write KEg = 1 2 cV 6 * Joseph-Louis Lagrange, 1736–1813, Italian/French mathematician. © 2014 by Taylor & Francis Group, LLC (3.17) Analytic and Computational Ballistics 73 The total kinetic energy of the system (neglecting recoil) is KE g = 3 ρ A V 2 xg 2 x2 3 x 1 cV 6 = 0 (3.18) But the kinetic energy of the projectile is KEshot = 1 wpV 2 2 (3.19) where wp is the projectile mass. So the Lagrange approximation for kinetic energy is KEtot = 1 1 1 c wpV 2 + cV 2 =  wp +  V 2 2 6 2 3 (3.20) In this development the volume of gas is assumed to be a cylinder of cross-sectional area A. In reality, it is not; while the bore is cylindrical, the chamber is not. Chamber diameters can be much greater than bore diameters. To account for this, an effort to modify the Lagrange gradient approximations has been performed [1]. This will be explored subsequently. The changes from the Lagrange gradient will be found to be small but not insignificant and the so-called chambrage gradient will be explained in Section 3.3 and incorporated in the discussion of numerical methods in Section 3.4. We can describe the linear momentum of the gas stream as Mom g = mgVxg (3.21) But, again, the mass of the gas is its density times the volume it occupies at time t, therefore x ∫ Mom g = ρ AVxg dxg (3.22) 0 We can use our continuity relationship in Equation 3.11 to write x x  xg dx   xg  Mom g =ρ A   dxg = ρ A  x V  dxg x t d     0 0 ∫ ∫ (3.23) 1 ρ AxV 2 (3.24) Performing the integration gives us Mom g = ρ A © 2014 by Taylor & Francis Group, LLC V xg2 x 2 x = 0 Ballistics: Theory and Design of Guns and Ammunition 74 If we recall Equation 3.16, we can write Mom g = 1 cV 2 (3.25) The total linear momentum of the system (neglecting the weapon) is Mom tot = Mom shot + Mom g (3.26) The linear momentum of the projectile is Mom shot = wpV (3.27) So the Lagrange approximation for linear momentum is Mom tot = wpV + 1 c  cV =  wp +  V 2 2   (3.28) Because we are looking for the parameters, we can readily measure breech pressure and muzzle velocity, and we must develop predictive equations for them, i.e., equations for pressure in terms of charge parameters and equations of motion of the projectile. To do this, we adopt a Lagrangian approach to track the motion of a particle of gas. What follows is a derivation for the equation of motion for an element of gas. For a rigorous, complete treatment, see any text on fluid mechanics, (e.g. [2]). For differentiation that tracks a fluid element (the Lagrangian approach), the following differential operator (called the substantial derivative or material derivative) is used: D ∂ ∂ ∂ ∂ = +u +v +w Dt ∂t ∂x ∂y ∂z (3.29) where u, v, and w are the velocity components in the x, y, and z directions, respectively. If we consider a one-dimensional flow operating on the velocity Vxg ( x) (here Vxg is the axial velocity and replaces u given earlier) DVxg ∂Vxg ∂Vxg = + Vxg ∂t ∂x Dt (3.30) In vector notation, the gradient of a function is ∇=i ∂ ∂ ∂ +j +k ∂x ∂y ∂z (3.31) Force is the time rate of change of momentum F= © 2014 by Taylor & Francis Group, LLC ∂ ( mv ) ∂t (3.32) Analytic and Computational Ballistics 75 It can be shown using Gauss’s theorem [3] that the rate of change of linear momentum of the fluid inside a surface S in changing to surface S’ in time, dt, is dv ∫ ρ dt dV (3.33) V From the equations of motion for an inviscid fluid we know that the total force equals the pressure on the boundary element integrated over the boundary plus the body force F integrated over the mass in S, or ∫ pndS +∫ FρdV = −∫ ∇pdV +∫ FρdV V S V (3.34) V Because by Gauss’s theorem ∫ pndS = − ∫ ∇pdV (3.35) V S Setting the RHS of Equation 3.35 equal to Equation 3.33, we get  dv  ∫ Fρ − ∇p − ρ dt  dV = 0 (3.36) V Since V is chosen arbitrarily, the sum in brackets must equal zero Fρ − ∇p − ρ dv =0 dt (3.37) In the absence of a body force F, we can rewrite this as dv 1 = − ∇p dt ρ (3.38) We can write Equation 3.38 as follows for one-dimensional flow and negligible body forces ∂Vxg   ∂Vxg dp + Vxg = −ρ   ∂xg  ∂ dxg t  (3.39) Note here that we have used the substantial derivative for the velocity of the gas stream. If we insert the relationship for the gas stream velocity we obtained through the continuity equation (3.11) into Equation 3.39, we can write  ∂  xg dx  ∂Vxg  dp = −ρ     + Vxg ∂ ∂xg  t x t dxg d    © 2014 by Taylor & Francis Group, LLC (3.40) Ballistics: Theory and Design of Guns and Ammunition 76 or  ∂  xg dx   xg dx  ∂  xg dx   dp = −ρ   +    dxg  ∂t  x dt   x dt  ∂xg  x dt   (3.41) We can combine terms in Equation 3.41 as follows:  xg  dx  2 xg d 2 x xg  dx  2  dp +  = − ρ − 2     + x dt 2 x 2  dt   dxg  x  dt  (3.42) Simplifying the expression gives us xg d 2 x dp = −ρ x dt 2 dxg or xg dp = −ρ xɺɺ x dxg (3.43) If we use our relationship between density and charge weight in Equation 3.8, we can write cxg dp =− xɺɺ Ax 2 dxg (3.44) We can integrate this expression with respect to the gas mass center as xg ∫ 0 xg dp c dxg = − xɺɺ xgdxg dxg Ax 2 ∫ (3.45) 0 Performing the integration yields p=− cxg2 xɺɺ + constant 2 Ax 2 (3.46) Let us now define pS = pressure at the projectile base pB = pressure at the breech p = mean pressure in volume behind projectile pR = pressure resisting projectile motion (force/bore area) We will develop the equations of motion both with a resistive force in the bore (such as friction and the air being compressed in front of the projectile) and neglecting the resistance. If we write Newton’s second law for a projectile being acted upon by propellant gases, we have wxɺɺ = ApS © 2014 by Taylor & Francis Group, LLC (3.47) Analytic and Computational Ballistics 77 Writing this in terms of the acceleration we get xɺɺ = A pS w (3.48) where w is the projectile mass. Since the base of our projectile is at location x and the local pressure on the base is pS, we can substitute these values into Equation 3.46 for xg and p to obtain pS = − c xɺɺ + constant 2A (3.49) Keep in mind that this is a local condition that we applied to the gas in the vicinity of the base (that gas’s mass center is approximately at x). We can rearrange Equation 3.49 to yield our constant of integration: c xɺɺ 2A (3.50) c A c   pS =  1 + p 2A w 2w   (3.51) constant = pS + If we use Equation 3.48, we obtain constant = pS + Inserting this constant back into our Equation 3.46 gives us p=− cxg2  A  cxg2 c  c    ɺɺ = − + + x p 1 S   pS +  1 +    pS 2 Ax 2 2w  2 Ax 2  w  2w    (3.52) or  xg2  c p = pS + pS  1 − 2  x  2w  (3.53) This equation relates the pressure at the base of the projectile to that at the location of the gas mass center. By similar logic, at the breech, xg = 0, and the pressure, p = pB, so we can substitute the values into Equation 3.53 to obtain a relationship between the breech pressure and the pressure at the projectile base pB = pS + pS c c   = pS  1 +  2w 2w   (3.54) The space-mean pressure is formally defined as x 1 p= pdxg x ∫ (3.55) 0 If we insert Equation 3.53 into this equation, we get x p= 1 x ∫ 0 © 2014 by Taylor & Francis Group, LLC   xg2  c   pS + pS  1 − 2   dxg x  2w    (3.56) Ballistics: Theory and Design of Guns and Ammunition 78 Solving this integral, inserting the limits of integration, and simplifying yields x 1 c c xg3  p =  pS xg + pS xg − pS  2w 2w 3 x 2  0 x (3.57) Inserting the limits of integration gives us p = pS + pS 1 c c − pS 2w 3 2w (3.58) Simplifying we get c   p = pS  1 +  3 w  (3.59) This equation relates the space-mean pressure to the base pressure acting on the projectile. We now have equations that relate breech pressure to base pressure (Equation 3.54) and space-mean pressure to base pressure (Equation 3.59). What is missing is a relationship between breech pressure and space-mean pressure. We can arrive at the desired result by dividing Equation 3.59 by Equation 3.54, simplifying to yield c   pS  1 +  p 3w   = c   pB pS  1 +  w 2  (3.60) For easier manipulation, it is sometimes desirable to expand Equation 3.60 in a Taylor series, which, neglecting higher order terms, would be p c = 1− +⋯ pB 6w (3.61) To account for the effects of bore resistance, we again write Newton’s second law for a projectile being acted upon by propellant gases and bore friction as w1xɺɺ = A( pS − pR ) (3.62) Here we have used w1 to represent the mass of the projectile (you will see why later) and have included a resistive pressure, pR, that fights the gas pressure. Note that the resistive pressure is simply the resistive force divided by the bore cross-sectional area so that the terms in the aforementioned equation can be conveniently grouped—it is not actually a pressure at all. Writing this in terms of the acceleration we get xɺɺ = © 2014 by Taylor & Francis Group, LLC A ( pS − pR ) w1 (3.63) Analytic and Computational Ballistics 79 Again, since the base of our projectile is at location x and the local pressure on the base is pS, we can substitute these values into Equation 3.46 for xg and p to obtain pS = − c xɺɺ + constant 2A (3.64) Remember that this is a local condition that we applied to the gas in the vicinity of the base where the gas’s mass center is approximately at x. Following the same procedure that we used to arrive at a general expression for pressure, but now with bore resistance, we rearrange Equation 3.64 to find the constant of integration, and with simplification arrive at c xɺɺ 2A (3.65) c A c  c   ( pS − pR ) =  1 +  pS − 2w pR w 2 A  w1 2 1  1   (3.66) constant = pS + If we use Equation 3.63, we obtain constant = pS + Inserting this constant back into Equation 3.64 gives us cxg2 c  c  xɺɺ +  1 + pS − pR 2Ax 2 2w1  2w1  (3.67) cxg2  A  c  c  pR ( pS − pR ) +  1 +  pS −  2  2 Ax  w1  2w  2w1  (3.68) p=− or p=− or p = pS + ( pS − pR ) xg2  c  1− 2  x  2w1  (3.69) which relates the pressure at the base of the projectile to the pressure at the gas mass center, but with the effect of bore friction included. Having this general equation we can again proceed as we did earlier to find equations that relate breech to base pressure, space-mean to base pressure, and space-mean to breech pressure for the bore friction case. These are c c pS − pR 2w1 2w1 (3.70) c  c  p = pS  1 + − pR  3w1  3 w1  (3.71) pB = pS + © 2014 by Taylor & Francis Group, LLC Ballistics: Theory and Design of Guns and Ammunition 80  p 1+  1 − R pS p  =  pB p 1+  1 − R pS   c   3w1  c   2w1 (3.72) If we plot breech, space-mean, and base pressure versus x, the position of the projectile base, we shall see that a gradient of pressure exists in which the breech pressure is always the greatest and the base pressure is always the smallest. This is the so-called Lagrange gradient and is fundamental to our modeling of the propellant gas. There are instances where this gradient is reversed and this usually means that we have a problem—a socalled negative delta-p. This is indicative of a fragmented propellant charge caused by poor ignition. A charge designed to move with the accelerating projectile, the traveling charge, is a notable exception. We are essentially prepared now to treat the F in the equation F = ma, which is in its simplest form, the base pressure times the base area. We now need to determine what generates the pressure, what the acceleration of the projectile will be, and how the acceleration and the ever-increasing volume behind the projectile affect the pressure. To do this, we shall review the equations from our initial discussions of propellant burning as well as revisiting our notation before moving on to combining everything into the equations of motion of the projectile. We have previously defined the following quantities and shall simply list them here for .. ease of reference. The first quantity is the projectile’s acceleration, x . The pressure acting on the base of the projectile is the stimulus that causes the acceleration pS (t) = pressure at the base of the projectile at time t. We usually measure pressure at the breech of the weapon and it is this pressure that we are determining when we examine the burning of the propellant. We need to constantly refer this breech pressure to the base pressure. We do this by invoking the Lagrange gradient assumption, keeping in mind that we begin by neglecting bore resistance c   pB = pS  1 +  2w   (3.73) We can write Newton’s second law for the force on the projectile base as wxɺɺ = pS A (3.74) If we substitute our Lagrange gradient into this equation to put it in terms of the breech pressure and the projectile velocity, we can write w © 2014 by Taylor & Francis Group, LLC pB dV = A c  dt  1 +   2w   (3.75) Analytic and Computational Ballistics 81 If we want to include losses, w can be replaced by w1, an effective projectile mass that can be thought of as an added mass due to the combination of resistance of bore friction, engraving by the rifling, resistance due to compression of the air ahead of the shot, etc. Then we have c  dV  = pB A  w1 +  2   dt (3.76) The burning of the propellant generates the pressure that pushes on the projectile. Let us now recall the equation that relates the amount of propellant turned to gas φ = (1 − f ) (1 + θ f ) (3.77) Also recall that the rate of gas evolution (burning) is a function of the pressure D df = − β p ≈ β pB dt (3.78) In our earlier study of solid propellant combustion, we developed an equation of state for the gas that related ϕ to the pressure and the distance the projectile traveled cλφ pB ( x + l ) = A c   1 + 2w 1   1+ c 3w1       (3.79) Finally, we have our equation of motion for the projectile c  dV  = pB A  w1 +  2   dt (3.80) whose initial conditions are x = 0, V = 0, f = 1 at t = 0. These equations may be manipulated to determine the parameters of interest as functions of the fraction of the remaining web f = f(t) x = projectile travel V = projectile velocity pB = breech pressure If we combine Equations 3.79 and 3.80, eliminating the breech pressure between them, we can write − © 2014 by Taylor & Francis Group, LLC c  dV D df w1  = 1+  β dt A  2w1  dt (3.81) Ballistics: Theory and Design of Guns and Ammunition 82 We can rearrange this to get the equation in terms of the projectile acceleration   DA  1  df dV = −   β w1  1 + c  dt dt 2w1   (3.82) This can be integrated resulting in V=− AD f + constant c   β w1  1 + 2w1   (3.83) If we insert the initial conditions that V = 0 when f = 1, Equation 3.83 yields constant = AD c   β w1  1 + 2w1   (3.84) This gives us V (t) = AD (1 − f (t)) c   β w1  1 + 2w1   (3.85) From the aforementioned equation, we can rearrange Equation 3.82 as follows: dV =− dt df AD c  dt  β w1  1 + 2w1   (3.86) We can now substitute our relationship between velocity and fraction of web remaining (Equation 3.86) into our projectile equation of motion (Equation 3.80), algebraically simplifying it and inserting the relationship for base pressure (Equation 3.79) to yield c   1+  cλφ D df 2w1  − =   β dt A( x + l)  1 + c  3w1   (3.87) This may be rearranged to obtain c   1+ df cλφβ  2w1  =−   AD ( x + l)  1 + c  dt 3w1   © 2014 by Taylor & Francis Group, LLC (3.88) Analytic and Computational Ballistics 83 Using the chain rule transformation between distance and time df df dx df = =V dt dx dt dx (3.89) df 1 df = dx V dt (3.90) This can be written as Now let us substitute Equations 3.85 and 3.88 into Equation 3.90, simplify the result and yield 2   c   1 +    df w1cλφβ 2 2w1     =− 2 2  dx A D ( x + l)(1 − f )  1 + c  w 3 1    (3.91) To examine the rate of change of f, the fraction of web remaining, with the travel distance, x, we take the reciprocal of Equation 3.91   c   1+ A D (1 − f )  dx 3w1  ( x + l) =− 2 df w1cλφβ 2   c    1 + 2w1     2 2 (3.92) Here l is an initial chamber length, to be described subsequently. By inserting the relationship between ϕ and f, from Equation 3.77 we get   c   1+ dx AD  3w1  ( x + l) =− 2 df w1cλβ 2   c   (1 + θ f )  1+ 2w1     2 2 (3.93) Equation 3.93 is cumbersome and following Corner [4] we find that we can define a dimensionless central ballistic parameter, M, that is a function of the gun, the charge, and the projectile, i.e., the system   c   1+ AD  3w1  M= 2 w1cλβ 2   c    1+ 2w1     2 © 2014 by Taylor & Francis Group, LLC 2 (3.94) Ballistics: Theory and Design of Guns and Ammunition 84 This simplifies our distance–web fraction relationship to dx ( x + l) = −M df (1 + θ f ) (3.95) The dimensionless nature of M can be shown if we note that c and w1 are mass units. We can also write the units of the burning rate coefficient as  D df   L2T  [β ] =  ⇒ [ β ] =     M   pB dt  (3.96) The units of the propellant force, λ, are 2  energy   ML L   L  [λ ] =  = ×  =   = [velocity ]2 2 M T   mass   T (3.97) Using these in our definition of the central ballistic parameter, we can show     L6   [ M] =  = [0] 2 2 2   L T L    M×M     M   T    (3.98) thereby demonstrating that M is dimensionless. Equation 3.95, repeated here, dx ( x + l) = −M df (1 + θ f ) (3.95) shows how M relates the burning of the propellant, f, with the expansion of the volume represented by x, the travel. A similar concept appears in all interior ballistic theories. We are now in a position to compute the parameters that the interior ballistician really seeks, the projectile’s velocity and the concomitant instantaneous breech pressure for each point along its travel down the tube. If we wish to know the pressure on the base of the projectile or the space-mean pressure in the volume behind the projectile, we need only apply the appropriate Lagrange approximation to the breech pressure. This is an extraordinary result. By simply understanding the amount of propellant burnt and some gun or propellant or projectile data, we have determined everything we need to know about the interior ballistics. We can now take the distance–web fraction relationship and integrate it directly. But we must examine two distinct cases for θ, the form factor of the grain. One where θ ≠ 0 and one where θ = 0. Let us separate the variables in Equation 3.95 to obtain df dx = −M ( x + l) (1 + θ f ) © 2014 by Taylor & Francis Group, LLC (3.99) Analytic and Computational Ballistics 85 Then we can write for θ ≠ 0 x ∫ 0 f df dx = −M ( x + l) (1 + θ f ) ∫ (3.100) 0 or, for θ = 0 f x ∫ 0 dx = − M df ( x + l) ∫ (3.101) 0 Evaluation of the integral equation (3.100) for θ ≠ 0 gives us ln( x + l) = − M −M ln(1 + θ f ) + ln(K ) = ln  K (1 + θ f ) θ     θ (3.102) Solving for K with the initial conditions, f = 1 at x = 0 we get M K = l(1 + θ ) θ (3.103) This constant, when inserted in the original Equation 3.102, gives us M  1+θ  θ x + l = l   1+θ f  (3.104) In a similar fashion, we can evaluate Equation 3.101 to give us the distance-remaining web fraction relation for θ = 0 x + l = le M ( 1 − f ) (3.105) We now know how the web fraction, f, varies with distance, and have, incidentally, shown the algebraic simplification inherent in the central ballistic parameter, M. We can now pursue a relationship between pressure and web fraction. If we look at Equation 3.88, we see the quotient on the RHS and note that this occurs frequently. We define it as our Lagrange ratio, RL, another simplification. c 2w1 RL = c 1+ 3w1 1+ (3.106) This will allow us to rewrite Equation 3.79 in simpler form as pB ( x + l ) = © 2014 by Taylor & Francis Group, LLC cλφ RL A (3.107) Ballistics: Theory and Design of Guns and Ammunition 86 We will make an assumption that the chamber and bore diameters are the same and relate the volume behind the projectile to a fictitious chamber length, l. (We will correct this subsequently when we examine the chambrage gradient.) Vi = Al = U − c δ (3.108) In this expression, U is the empty chamber volume and c/δ is the volume occupied by the solid propellant charge. We continue by substituting Equations 3.108 and 3.77 into Equation 3.107 and rearranging to give our relationship between the breech pressure and the fraction of remaining web for θ ≠ 0: M λ cRL  1 −θ f  θ pB = (1 − f ) (1 + θ f )   Vi  1 −θ  for θ ≠ 0 (3.109) We can also proceed in similar fashion for θ = 0 by substituting Equations 3.105 and 3.77 into Equation 3.107 to find the relationship between the breech pressure and the fraction of remaining web: pB = λ cRL (1 − f ) (1 + θ f )exp  − M(1 − f ) for θ = 0 Vi (3.110) Summarizing, we now have the definition of the central ballistic parameter (Equation 3.94) and equations that relate velocity as a function of remaining web (Equation 3.85) and travel as a function of remaining web for different form functions (Equations 3.104 and 3.105) as well as breech pressure as a function of remaining web for different form functions (Equations 3.109 and 3.110). With these we can now integrate the governing equations and find solutions for velocity at peak pressure, at all-burnt point of travel, and at muzzle exit. Equations 3.109 and 3.110 are somewhat cumbersome to work with, so we shall define a parameter, Q, as follows: c   1+  λc  2w1  λ c Q= RL = c  Vi Vi  1 +  3w1   (3.111) Then we can rewrite Equation 3.109 in a more compact way M  1+θ f  θ pB = Q(1 − f ) (1 + θ f )    1+θ  (3.112) The maximum or peak pressure attained is then found by taking the first derivative of pB with respect to f and setting it equal to zero M M +1  dpB M  = Q (1 − f )  + 1  θ (1 + θ f ) θ − (1 + θ f ) θ  = 0  df θ   © 2014 by Taylor & Francis Group, LLC (3.113) Analytic and Computational Ballistics 87 Let us solve Equation 3.113 for f. By introducing the subscript “m” to denote maximum, we obtain the product of two terms (1 + θ f m ) ( M + θ )(1 − f m ) − (1 + θ f m ) = 0 (3.114) Solving this we have two choices here, either (1 + θ f m ) = 0 or ( M + θ )(1 − f m ) − (1 + θ f m ) = 0 (3.115) The first would only be admitted for the special case of θ = −M, thus, our criteria for determination of fm is ( M + θ )(1 − f m ) − (1 + θ f m ) = 0 (3.116) and fm = M +θ − 1 M + 2θ (3.117) Equation 3.117 works for all values of θ. If we want to determine ϕm, the fraction of propellant burnt at peak pressure, we call on our relationship between f and ϕ, Equation 3.77. Here we have denoted peak values with the subscript “m”: φ = (1 − f )(1 + θ f ) (3.118) Substitution of Equation 3.117 into the aforementioned equation yields  M +θ − 1    M + θ − 1  φm = 1 −   1 + θ    M + 2θ    M + 2θ   (3.119) This, when simplified, gives φm = (1 + θ ) [ M + θ + θ ( M + θ )] (1 + θ ) [( M + θ )(1 + θ )] = [ M + 2θ ]2 [ M + 2θ ]2 (3.120) or the following (valid for all θ): φm = ( M + θ )(1 + θ )2 [ M + 2θ ]2 (3.121) In designing a gun (and for other reasons), it is desirable to know where a projectile is in its travel down bore when the pressure is at a maximum. This involves substitution of Equation 3.117 into Equation 3.104 and for the case where θ ≠ 0 this yields M  θ   1+θ  xm + l = l  1+θ  M +θ − 1       M + 2θ   © 2014 by Taylor & Francis Group, LLC (3.122) Ballistics: Theory and Design of Guns and Ammunition 88 Simplifying, we finally get M  ( M + 2θ )  θ xm + l = l    (M + θ )  for θ ≠ 0 (3.123) For the case where θ = 0, we substitute Equation 3.117 into Equation 3.105, which we rewrite as follows: xm + l = l exp  M(1 − f m ) (3.124)   M + θ − 1  xm + l = l exp  M  1 −  M + 2θ     (3.125) On substitution we get Simplifying this result and substituting θ = 0 into it gives us xm + l = le for θ = 0 (3.126) for a zero form factor. Knowing now the position of the peak pressure in the bore, we can then ask what the breech pressure would be at this point. We can insert the value we have for the fraction of remaining web at peak pressure, fm, back into the breech pressure equation for θ ≠ 0 (Equation 3.109) M  1 + θ fm  θ pBm = Q(1 − f m )(1 + θ f m )    1+θ  (3.127) With considerable algebraic simplification including substituting the values for Q, and the Lagrange ratio, RL, for the case θ ≠ 0, we finally arrive at pB m c  1+ λc  2 w 1 =  Vi  1 + c  3w1     (1 + θ )2 ( M + θ ) Mθ +1    M +2    ( M + 2θ ) θ    (3.128) Following a similar procedure, we now insert the value we have for the fraction of remaining web at peak pressure, fm, into the breech pressure equation for θ = 0 (Equation 3.110) pBm = Q(1 − f m )exp  − M(1 − f m ) (3.129) Then substituting for Q and RL and simplifying, we see that we have characterized the breech pressure at the instant peak pressure is achieved down bore: pB m © 2014 by Taylor & Francis Group, LLC c  1+ λc  2w1 =  Vi  1 + c  3w1     1     M exp[1]    (3.130) Analytic and Computational Ballistics 89 Determining the breech pressure and travel when the solid grains have been completely consumed is also of considerable interest. We shall use the subscript “c” to represent charge burnout. If the charge is designed properly, it will burnout somewhere in the bore that allows us to extract most energy from the propellant and reduces the muzzle blast. Recall from our previous discussions that at t = 0, x = 0, f = 1, and ϕ = 0 but at all burnt (subscript “c”), t = tc, x = xc, f = 0, and ϕ = 1. If we substitute f = 0 in Equations 3.109 and 3.110, we obtain the breech pressure at the instant of charge burnout c   M 1+  λc  2w1   1  θ pB c = c   1 + θ  Vi   1 + 3w  1   for θ ≠ 0 (3.131) c   1+ λ c  2w1  exp[ − M] for θ = 0 pB c = c  Vi  1 +  3w1   (3.132) and The travel of the projectile at burnout is a data point we usually want to know because if this distance turns out to be longer than the barrel length, then the charge is not completely burnt when the projectile exits. If we substitute f = 0 in Equations 3.104 and 3.105, we obtain the position of the projectile at the instant of charge burnout M xc + l = l(1 + θ ) θ for θ ≠ 0 (3.133) xc + l = lexp[M] for θ = 0 (3.134) and It is a good idea to use these equations first to see whether the propellant burns out in the tube with the parameters we have designed into the grain. Still-burning grains leaving the tube signify a poorly designed charge. For completeness, however, if charge burnout happens outside the bore, the pressure at the breech location when the projectile leaves the muzzle may be calculated by evaluating f at the muzzle through Equation 3.104 or 3.105 and using this value to calculate pB from Equation 3.109 or 3.110. The muzzle velocity could then be obtained from Equation 3.85. If charge burnout is, as desired, in the bore, recall that there is still a net force (pressure) pushing on the projectile. A simple means of calculating this pressure is to assume that the process occurs so quickly that it is essentially adiabatic and that the gas behaves as an ideal gas. With these assumptions and the initial conditions that the pressure is pBc and the distance is xc, we have a closed form solution to the problem. It is vitally important to note that the expansion of the gas after charge burnout is neither adiabatic nor isentropic, © 2014 by Taylor & Francis Group, LLC Ballistics: Theory and Design of Guns and Ammunition 90 however the result is usually within about 5% with respect to pressure. The isentropic relationships for an ideal gas are  p ρ   =  = p0  ρ 0     γ 1 v 1 v0 γ    v0 γ  v −γ  =  =    v   v0    (3.135) This equation relates pressure to specific volume in a general way, but we need to involve the projectile travel as well. We can express the volume behind the projectile as a function of distance as V( x) = ( x + l)A (3.136) Then the specific volume of the gas is this value divided by the mass of the gas, which we know is still c after burnout. Thus, we can write Equation 3.136 in its intensive form as v( x) = ( x + l) A c (3.137) Furthermore, we can specialize this to the point at which the charge burns out and write v( xc ) = ( xc + l) A c (3.138) We can now tailor Equation 3.135 to our needs by substituting the conditions at burnout as our reference conditions p( x)  v( x)  =  pc  v( xc )  −γ  x+l  =   xc + l  −γ (3.139) This condition occurs often so we define r( x ) = x+l xc + l (3.140) which can be written in more compact form as p( x ) = r ( x )− γ pc (3.141) A sketch of this situation is depicted in Figure 3.2. These extensive preparations have finally brought us to the goal of interior ballistics and the design of a gun system—imparting a desired velocity to a projectile and being able to repeat that process at will. We have developed the means for predicting how the © 2014 by Taylor & Francis Group, LLC Analytic and Computational Ballistics 91 x xc l Chamber Propellant burnout FIGURE 3.2 Position of projectile at charge burnout. propellant burns over time, how the breech, space-mean, and base pressures vary with time, and where the projectile moves to in relation to these pressures. Now we will focus on the velocity of the projectile during this ballistic cycle. Recall that the kinetic energy of the projectile plus the gas losses was written as KEtot = 1 1 1 c wpV 2 + cV 2 =  wp +  V 2 2 6 2 3 (3.20) The work done on the projectile and the gas from charge burnout to the point of interest (usually muzzle exit) is x ∫ W = A p dx (3.142) xc Combining Equations 3.20 and 3.142 and inserting Equation 3.139 yields x −γ 1 c 2  x+l  2  w1 +  V ( x) − V ( xc ) = Apc   dx 2 3  xc + l  x ∫ (3.143) c We must keep in mind that we are using space-mean pressure here because the work is being done on both the projectile and the gas. We can use any of breech, space-mean, and base pressure (with the appropriate relationship) because we know each in terms of the others. Integrating and rearranging we get x 1 1 c 2 2 ( x + l )− γ + 1  w1 +  V ( x) − V ( xc ) = Apc 2 3 (1 − γ )(xc + l)−γ x (3.144) c Evaluation of the limits of integration yields 1 1 c 2 2 ( x + l)1−γ − ( xc + l)1−γ   w1 +  V ( x) − V ( xc ) = Apc 2 3 (1 − γ )(xc + l)−γ  © 2014 by Taylor & Francis Group, LLC (3.145) Ballistics: Theory and Design of Guns and Ammunition 92 Rearranging and inserting Equation 3.132 into the aforementioned equation, we get a velocity relationship after burnout for θ = 0 V 2 ( x ) − V 2 ( xc ) =  ( x + l)1−γ  2 Aλ ce − M ( xc + l)1−γ − 1  −γ Vi (1 − γ )( xc + l) (w1 + 3c )  ( xc + l)1−γ  (3.146) But recall that the volume Vi = Al and if we define 1− γ  2  x + l  Φ= − 1   (1 − γ )  xc + l   (3.147) We can write V 2 ( x ) − V 2 ( xc ) = λ c( xc + l)exp[ − M] Φ c  l  w1 +  3  (3.148) Here again we revert to the cases of the form factor being zero or not zero and examine the former first. Recall Equation 3.105. For conditions after charge burnout, there is no remaining web (f = 0), so we can write x c + l = le M for θ = 0 and f =0 (3.149) Rearranging this and substituting it into Equation 3.149 yields V 2 ( x ) − V 2 ( xc ) = λc c   w1 +  3  Φ (3.150) This allows us to calculate the velocity of a projectile after burnout of the web for θ = 0. The case for nonzero θ requires further examination and manipulation. In Equation 3.85, we had a general expression for velocity as a function of remaining web. After burnout this becomes V ( xc ) = AD (3.151) c  β  w1 +  2  Since we are working with kinetic energy, squaring this gives V 2 ( xc ) = A 2D 2 c  β  w1 +  2  2 (3.152) 2 We defined the central ballistic parameter, M, in Equation 3.94, and can rearrange it for our purposes into the form cλ M A 2D 2 = 2 c  c   w1 +  β 2  w1 +  3  2  © 2014 by Taylor & Francis Group, LLC (3.153) Analytic and Computational Ballistics 93 When this is compared with Equation 3.152, we conclude that V 2 ( xc ) = λ cM c   w1 +  3  (3.154) This is an important result—it says that just by knowing the physical parameters of the weapon, projectile, and charge one can predict the projectile velocity at charge burnout. With this result and continuing the examination for nonzero θ, we can then say that Equation 3.150 is valid for any θ. If we solve Equation 3.150 for the velocity at any point, V(x), insert Equation 3.154, and rearrange the terms, we get V 2 ( x) = cλ ( M + Φ ) for θ ≠ 0 c   w1 +  3  (3.155) This result along with our earlier work allows us now to determine projectile velocity at all points in the gun for charge grains of all form factors both before and after burnout. We have been through many derivations that have led us to the essentials of interior ballistics—breech pressure and velocity in terms of projectile travel. These results, furthermore, are in closed form, accessible to computation by hand calculator. Specialized pressures, spacemean pressure and projectile base pressure, may be computed from the breech pressure data using the Lagrange approximations. Projectile design and gun design proceed from these equations. In the following sections, we shall discuss refinements to the Lagrange formulation with an emphasis on the use of modern computer programs that take the drudgery out of hand calculation and provide the ability to iterate solutions for small changes in the parameters. Problem 1 You are asked to analyze the pressure of a charge zero (igniter) firing in an M31 boom for a 120 mm mortar projectile. You decide to examine it as a closed bomb first. Assume we have 59 g of M48 propellant (properties given later). The volume of the closed bomb is 5.822 in.3. The propellant grains are balls (roughly spherical) with a diameter (web) of 0.049 in. M48 propellant properties  lbm  Density ρ = 0.056  3   in.  Ratio of specific heats γ = 1.21  in.3  Co-volume b = 26.72    lbm  Isochoric flame temperature T0 = 3720°F Burn rate exponent α = 0.9145  in.  Average burn rate coefficient β = 0.0095    s ⋅ psi  Burn rate D df  in.  = 40.341   dt  s   ft-lbf  Force constant λ = 391, 000   lbm  © 2014 by Taylor & Francis Group, LLC Ballistics: Theory and Design of Guns and Ammunition 94 1. Come up with the equation for the web fraction, f, as a function of time. Answer: f = (1 − 823.29t) [%] 2. For a sphere, the fraction of propellant burnt has the functional form ϕ = 1 − f 3, write this in terms of time and ϕ. Answer: ϕ = 2,470t − 2,033,419t2 + 558,031,251t3 3. Determine how long it will take the propellant to burn halfway through and all the way through. Answer: Time to burn through halfway is 0.6 ms 4. Using the Noble–Abel equation of state, determine the pressure in the vessel when half of the propellant is burnt and when all of the propellant is burnt. Note that this cannot usually occur as the propellant is a charge zero firing that is vented into the main ullage volume behind the mortar bomb (significantly greater volume).  lbf  Answer: p = 258, 746  2   in.  Problem 2 If we use the Lagrange approximation in examination of a 155 mm projectile launch, what is the average pressure in the volume behind a 102 lbm projectile if the breech pressure is 55,000 psi? The propelling charge weighs 28 lbm.  lbf  Answer: p = 52, 787  2   in.  Problem 3 A 120 mm projectile is to be examined while in the bore of a tank cannon at a time 4 ms from shot start. Over this time period, the projectile has acquired an average velocity of 1000 ft/s. The propellant grain (M15) is single perf (θ = 0) with a 0.034 in. initial web. The co-volume of the propellant is 31.17 in.3/lbm. The density of the propellant is 0.06 lbm/in.3. If the projectile weighs 50.4 lbm, the propellant weighs 12.25 lbm and the chamber volume is 330 in.3. At this time, 0.02 in. of the web remains. The propellant force is 337,000 ft-lbf/lbm. Determine the breech pressure in the weapon. Be careful with the units!  lbf  Answer: pB = 21, 784  2   in.  Problem 4 The Paris gun was a monstrous 210 mm weapon designed by Germany during the First World War to bombard Paris from some 70 miles away. It was unique in that it fired the first exo-atmospheric projectile ever designed. The weapon had a chamber volume of 15,866 in.3. Very little of the projectile protrudes into the chamber after it seats (so ignore the volume the base occupies). The length of travel for the projectile from shot start to shot exit is 1182 in. The projectile weighs 234 lb. The propelling charge weighs 430.2 lb. The propellant used was specially designed and was similar to U.S. M26 propellant. It consisted of 64%–68% NC, 25%–29% NG with 7% Centralite (symmetrical diethyl diphenylurea C17H20N2O), and some other additives. The propellant was single perforated with a web thickness described later. Assume the propellant has the following properties. © 2014 by Taylor & Francis Group, LLC Analytic and Computational Ballistics 95 (Note that these are the authors’ guesses—a better estimate of the properties can be found in Ref. [5].) Adiabatic flame temperature T0 = 2881 K Specific heat ratio γ = 1.237 Co-volume b = 1.06 cm3/g Density of solid propellant ρ = 1.62 g/cm3 Propellant burn rate coefficient β = 0.0707 (cm/s)/(MPa) Web thickness D = 0.217 in. Propellant force λ = 1019 J/g 1. Using the aforementioned data, determine (a) the projectile base pressure in psi, (b) velocity in ft/s, and (c) distance down the bore of the weapon in inches for peak pressure.  lbf  Answer: (a) pSmax = 31, 548  2   in.   ft  (b) Vpmax = 2880   s (c) x pmax = 270.1 [in.] 2. Determine the pressure in psi at a point 3 in. behind the projectile base when the charge burns out.  lbf  Answer: px− 3 = 28, 637  2   in.  3. Assuming the gas behaves according to the Noble–Abel equation of state, determine the muzzle velocity of the projectile in ft/s.  ft  Answer: V = 5791   s Problem 5 A British 14 in. Mark VII gun has a chamber volume of 22,000 in.3. A 5 in. of the projectile protrude into the chamber after it seats. The length of travel for the projectile from shot start to shot exit is 515.68 in. The weapon has a uniform twist of 1 in 30. The projectile weighs 1590 lb. The propelling charge weighs 338.25 lb. The propellant used is called “SC” and consists of 49.5% NC (12.2% nitrated), 41.5% NG with 9% Centralite. Assume SC propellant has the following properties: Adiabatic flame temperature T0 = 3090 K Specific heat ratio γ = 1.248 Co-volume b = 26.5 in.3/lbm Density of solid propellant ρ = 0.0567 lbm/in.3 Propellant burn rate β = 0.000331 (in./s)/(psi) Web thickness D = 0.25 in. Specific molecular weight n = 0.04262 lb-mol/lbm 1. Determine the force constant, λ in ft-lbf/lbm. 2. Determine the central ballistic parameter for this gun–projectile combination. © 2014 by Taylor & Francis Group, LLC Ballistics: Theory and Design of Guns and Ammunition 96 3. Using the aforementioned data, determine the projectile base pressure, velocity, and distance down the bore of the weapon for both peak pressure and charge burnout assuming the grain is a cylindrical propellant (θ = 1). Answers:  ft-lbf  λ = 366, 246   lbm  M = 1.933  lbf  pBmax = 43, 281  2   in.   lbf  pBc = 26, 350  2   in.   lbf  psmax = 39, 120  2   in.   lbf  psc = 23, 817  2   in.   ft  Vpmax = 1082   s  ft  Vc = 2128   s x pmax = 75.6 [in.] xc = 293.7 [in.] Problem 6 Verify Equation 3.150 is valid for any θ. Problem 7 You are asked to design a gun to propel a fragment for an explosive initiation test at 3000 ft/s. The diameter of the chamber and bore is to be 0.50 in. You have on hand some M10 flake propellant with the ballistic properties given later. Because of space limitations the device (chamber and bore) cannot exceed 5 ft in length. The fragment plus sabot weighs 0.1 lbm. Assume the propellant has the following properties: Adiabatic flame temperature Specific heat ratio Co-volume Density of solid propellant Propellant burn rate coefficient Web thickness Propellant force © 2014 by Taylor & Francis Group, LLC T0 = 3000 K γ = 1.2342 b = 27.76 in.3/lbm δ = 0.0602 lbm/in.3 β = 0.00002468 (in./s)/(psi) D = 0.011 in. λ = 339,000 ft-lbf/lbm Analytic and Computational Ballistics 97 (a) Using the aforementioned data, determine the bore and chamber length for the weapon as well as the amount of propellant required—be careful to leave some air space in the chamber. (b) Once the system is established in (a), determine the central ballistic parameter, value of peak breech pressure, and location of the projectile when peak pressure occurs. Problem 8 A test of an experimental propellant grain yields a parametric relation for the fraction of web remaining of f (t) = C1t 2 + C2t + 1 In the equation above, C1 and C2 are constants. The burn rate model for this propellant was determined to be D df = − β pαB dt Determine the expression for the velocity of the projectile as a function of time for a given gun-projectile system. Describe the assumptions used and why they are relevant. Problem 9 A black powder charge is to be designed to throw a 3 in. diameter firework canister that weighs 2 lbs at 100 ± 2 ft/s. The tube is 3 in. in diameter and the chamber where the propellant charge sits is 6 in. long. The projectile travel distance can be up to 18 in. long but you may size the tube length. The propellant grains are spherical and obey the relation φ (t) = 1 − f 3 (1) The average diameter of the propellant grains (they vary quite a bit and they are not exactly spherical) is 0.02 in. The tube can handle a maximum pressure of 2000 psi. Assume the linear burn rate for the black powder is 2.22 in./s measured at 3,329 psi and the force constant is 105,000 ft · lbf/lbm. Assume a specific heat ratio of 1.21 and a solid density of 0.06 lbm/in.3. Listing all assumptions a. b. c. d. Develop the proper equations for the motion of the projectile. Size the charge so that the system will function as required—make sure it fits too! Determine the peak pressure in psi and location of peak pressure in inches. Determine the location of charge burnout in inches and the velocity at burnout in ft/s. Please note that if your design allows unburnt propellant to exit the tube please state that and calculate muzzle exit pressure and velocity in this case. e. Determine the muzzle velocity in ft/s. © 2014 by Taylor & Francis Group, LLC Ballistics: Theory and Design of Guns and Ammunition 98 Problem 10 A 7 perf grain is a very common geometry in weapons. The geometry is such that the web between the outer diameter of the grain and the I.D. of the outer perforations and the web between the perforations themselves is equal. Determine, up to the point of slivering, the following: a. The equation for ϕ(t) for any perforation size and any web b. If possible obtain an estimate for the value of θ assuming the perforations are ¼ of the web thickness, if not, show why not End effects may be neglected. Problem 11 A Japanese 18.1 in. Type 94 gun was the largest weapon ever mounted on a warship. The gun had a chamber volume of 41,496 in.3. An estimated 8 in. of the projectile protrude into the chamber after it seats. The length of travel for the projectile from shot start to shot exit is 806.3 in. The weapon has a uniform twist of 1 in 28. The type 91 AP projectile weighs 2,998 lbs. The propelling charge weighs 794 lbs. The propellant used is called “DC1” and consists of 51.8% NC (11.85% nitrated), 41.0% NG with 4.5% Centralite (symmetrical diethyl diphenylurea C17H20N2O), 2.0% orthotolyl urethane (added as an improvement to the centralite) and 0.7% mineral matter (salts for wear and flash reduction) [6]. Assume DC1 propellant has the following properties: Adiabatic flame temperature Specific heat ratio Co-volume Density of solid propellant Propellant burn rate Web thickness Propellant force T0 = 3000 K γ = 1.23 b = 27.0 in.3/lbm δ = 0.059 lbm/in.3 β = 0.000300 (in./s)/(psi) D = 0.184 in. λ = 284,000 ft lbf/lbm a. Determine the central ballistic parameter for this gun/projectile combination. b. Using the aforementioned data, determine the projectile base pressure, velocity, and distance down the bore of the weapon for both peak pressure and charge burnout assuming the grain is single perforated propellant. c. Determine the muzzle velocity of the weapon and the pressure acting on the projectile at muzzle exit. Problem 12 A British “2 Pounder” (so-called because the projectile weighed about two pounds) was their main anti-tank weapon for the first two and a half years of the Second World War. It fired a 40 mm projectile that weighed 1.94 lbm. The gun had a chamber volume of 23.0 in3. Since the shot had a flat base, when crimped to the cartridge case none of the projectile protrudes into the chamber. The length of travel for the projectile from shot start to shot exit is 70 in. The weapon has a uniform right-hand twist of 1 in 30. The propelling charge © 2014 by Taylor & Francis Group, LLC Analytic and Computational Ballistics 99 weighs 0.583 lbs. The propellant used was single perforated cordite. Assume the cordite propellant has the following properties: Adiabatic flame temperature Specific heat ratio Co-volume Density of solid propellant Propellant burn rate Web thickness Propellant force T0 = 2442 K γ = 1.21 b = 31.32 in.3/lbm δ = 0.059 lbm/in.3 β = 0.00024 (in./s)/(psi) D = 0.0197 in. λ = 318,000 ft lbf/lbm a. Determine the central ballistic parameter for this gun/projectile combination. b. Using the aforementioned data, determine the projectile base pressure, velocity, and distance down the bore of the weapon for both peak pressure and charge burnout. c. Determine the muzzle velocity of the weapon and the pressure acting on the projectile at muzzle exit. d. Create a pressure–travel and velocity–travel curve for this system. Annotate the location of charge burnout on the pressure–travel curve. Problem 13 Develop the equations required to model a propellant/charge/gun system where the propellant behaves according to a βpα burn rate. Do this for both θ = 0 and θ ≠ 0. Describe, in bulletized form, how you would solve these equations numerically. Hint: Start with Equations 3.80 and 3.77 through 3.79 making sure that you alter Equation 3.78 to the new burn rate. Problem 14 Write a code using any software you want to solve Problem 13. Only write the code so it solves the interior ballistics problem up to charge burnout. Check the code by setting α = 1 and show that you get answers close to that obtained in Problem 12. Problem 15 Use your code developed in Problem 14 to obtain a solution to the same problem assuming black powder is the propellant. Only use your code to take the problem to the all-burnt point. Assume the properties of black powder are as follows: Adiabatic flame temperature Specific heat ratio Co-volume Density of solid propellant Propellant burn rate Web thickness Propellant force © 2014 by Taylor & Francis Group, LLC T0 ≈ 2300 K γ ≈ 1.22 b ≈ 31 in.3/lbm δ = 0.060 lbm/in.3 β = 0.04044 (in./s)/(psi0.511) D = 0.0197 in. λ = 105,000 ft lbf/lbm 100 Ballistics: Theory and Design of Guns and Ammunition Problem 16 A Japanese (designed by the British firm of Vickers) 14 in./45 cannon is to be examined. It fired a 14 in. (36 cm) projectile that weighed 1485 lbm. The gun had a chamber volume of 17,996 in.3 [6]. Assume 4 in. of the projectile protrudes into the chamber. The length of travel for the projectile from shot start to shot exit is 540.8 in. [6]. The weapon has a uniform right-hand twist of 1 in 28. The propelling charge has four increments, where each weighs 78.45 lbs. The propellant used was DC that consisted of 64.8% NC, 30% NG, 4.5% centralite, and 0.7% mineral matter [6]. Assume the propellant geometry is such that θ = 0.1. Assume the DC propellant has the following properties: Adiabatic flame temperature Specific heat ratio Co-volume Density of solid propellant Propellant burn rate Web thickness Propellant force T0 = 3200 K γ = 1.23 b = 27.0 in.3/lbm δ = 0.059 lbm/in.3 β = 0.000298 (in./s)/(psi) D = 0.165 in. λ = 365,000 ft lbf/lbm The weapon was “zoned” to fire using 2, 3, and 4 bags that were called, “weak,” “reduced,” and “full” [7]. For each of these charge configurations a. Determine the central ballistic parameter for this gun/projectile/propellant combination. b. Using the aforementioned data, determine the projectile breech pressure for both peak pressure and charge burnout. c. Using the aforementioned data, determine the projectile base pressure, velocity, and distance down the bore of the weapon for both peak pressure and charge burnout. d. Determine the muzzle velocity of the weapon and the pressure acting on the projectile at muzzle exit. e. Plot the pressure versus distance based on the aforementioned results at the instant of peak pressure and muzzle exit. 3.3 Chambrage Gradient In our derivation of the Lagrange gradient approximations, we assumed that the chamber of the gun was simply an extension of the bore. The volume of the chamber was converted to a cylinder of bore diameter and the tube was lengthened appropriately behind the projectile. In doing this, we neglected the effects of short, larger diameter chambers (the definition of chambrage is the ratio of the diameter of the chamber to the bore inner diameter) and all calculations that are functions of distance from the breech, x–xs, are inaccurate in the distance term. If we account for these differences by deriving a chambrage gradient, we find that the two methods yield similar but close answers. Nevertheless, one should understand how the answers relate to each other and to the real problem. Fredrick W. © 2014 by Taylor & Francis Group, LLC Analytic and Computational Ballistics 101 Robbins of the Army Research Laboratory, who has allowed us to base this section on his excellent work, derived the chambrage gradient formulation that follows. The formulation of the chambrage gradient follows much the same pattern that was used in the development of the Lagrange gradient. It leads, however, to an algorithm that is best applied with the aid of a computer. Small increments of time (hence distance) are chosen and computations of pressure (breech, mean, and base), velocity, acceleration, and distance traveled are made for the end point of the interval. The calculation is then repeated for the next increment of time. This is done until the projectile exits the bore. A representation of the situation is shown in Figure 3.3 for a chosen time step. The definitions of the terms used in Figure 3.3 are shown in Figure 3.4. In Robbins’ derivation, certain integrals called J integral factors are developed and must be computed. They are x0 J 1 ( x0 ) = V(x) ∫ A(x) dx (3.156) 0 J 1 ( xs ) = J 1 ( x0 ) + 1 Ab Ab  2  V( x0 )( xs − x0 ) + 2 ( xs − x0 )  (3.157) 2 J 2 ( xs ) = [ V(x0 ) + Ab (xs − x0 )] (3.158) Ab2 Ab Vs A(x) V(x) x, V(x) xs , V(xs) FIGURE 3.3 Chamber with large chambrage. Velocity of propellant gas Velocity of the projectile at position, x measured from at the time of interest the breech at the time of interest V(x) = Ab VS V(xS) Cross-sectional area of the bore V(x) A(x) Cross-sectional area of the weapon at position, x at the time of interest Volume behind the projectile base at the time of interest FIGURE 3.4 Definitions of terms used in chambrage gradient development. © 2014 by Taylor & Francis Group, LLC Volume at position, x at the time of interest Ballistics: Theory and Design of Guns and Ammunition 102 J 3 ( xs ) = J 3 ( x0 ) + Ab J1( x0 )( xs − x0 ) + V( x0 ) A ( x s − x 0 )2 + b ( x s − x 0 )3 2 6  V( x0 ) + Ab ( xs − x0 )3  − [ V( x0 )] J 4 ( xs ) = J 4 ( x0 ) +  3 Ab2 (3.159) 3 (3.160) The acceleration at any point, a, appears in one of our algorithm factors explicitly a(t) = a1(t) + a2 (t)ps (3.161) where a1(t) =  AbVs2 cAb presist  +  2  mp  [ V(xs )]  V(xs ) cAb a2 (t) = − cAb2 mp [ V( xs )] 2 (3.162) (3.163) Another factor required in the algorithm is b(t) derived as b(t) = − cAb2Vs2 2 [ V( xs )] 3 (3.164) The way the algorithm is used is (roughly) as follows: At each time step • • • • • • The breech pressure is calculated from the burning rate equations J1 through J4 are calculated a(t) and b(t) are calculated The projectile acceleration, velocity, and distance down the bore are calculated The volume behind the projectile is updated The process moves to the next time step This gradient, while only slightly more accurate than the Lagrange gradient in the computed distance from the breech, is used in some modern interior ballistic computer codes. 3.4 Numerical Methods in Interior Ballistics In this section, we shall briefly discuss methods for solving the interior ballistics problem through use of computational tools. In recent decades, computational capabilities have increased at an astronomical rate. One of the most famous early uses of the computer to © 2014 by Taylor & Francis Group, LLC Analytic and Computational Ballistics 103 solve the exterior ballistics problem (firing tables) was the use of the electronic numerical integrator and computer (ENIAC) machine during and immediately after the Second World War. In this case, the computer was used to solve tedious exterior ballistics problems in rapid order. In the field of interior ballistics, the computer revolution has given the individual ballistician the tools (although some commercial packages can be expensive) to solve extremely complicated interior ballistics problems and optimize a system quickly. The complexity of these tools is driven by the physics that are incorporated in the particular code. We shall discuss some general categories of software, their uses, and their limitations. Many interior ballistics codes are of the zero-dimensional variety. In these types of codes, the density of the propellant gas (as stipulated by the Lagrange approximation) is considered constant in the volume between the breech and the projectile. The Lagrange pressure gradient is assumed to be in effect and results in a nice, always well-behaved launch. These codes are extremely useful for predictive applications because they run fast. One of the features of these codes that make them so useful is that we can easily include and track burn characteristics of multiple propellant types (both geometry and chemical composition). This allows us to tailor the burn characteristics so that a particular pressure–distance distribution is achieved while maintaining a particular muzzle velocity. Another excellent feature of this type of code is that heat transfer to the weapon can be accounted for in the energy balance. This provides a more realistic muzzle velocity than if it is neglected and can be of great value to the gun designer. Friction and blow-by effects can be fudged in and burn rate parameters varied to replicate actual tests. Additionally, the effects of the regression of all surfaces (recall that we neglected end effects in our hand calculation methods) can be simulated and accounted for. Zero-dimensional codes can also include the effects of inhibitors on the propellant grains as well as highly nonlinear pressure–burn rate relationships. Since zero-dimensional codes track the pressure, it is simple enough to use them to develop recoil models as well. All in all, zero-dimensional codes are probably the most effective tools at the disposal of the interior ballistician for basic ballistics design work. Once a set of experiments have been conducted to validate these codes, their accuracy is excellent. A quasi-one-dimensional code is one in which the density of the propellant gas behind the projectile is a known function of some other variable. An example of this would be a zero-dimensional code that incorporated the chambrage gradient. Essentially, beyond the ability to track the effect of variable chamber or bore area on the density, the limitations and benefits of this type of code are the same as discussed in the zero-dimensional section. A one-dimensional interior ballistics code allows density to vary based on the physical equations and conservation laws in the axial direction only. Thus, at a given cross section, the density is considered constant throughout the radial direction. These codes are very good at predicting pressure waves and therefore can estimate the pressure differential along the volume behind the projectile. The benefit of this is that, since propellant generally burns faster under higher pressure, the local burn rate and therefore the amount of gas evolved can be tracked. This allows the user to see pressure waves develop and propagate. The disadvantage is that, since the code can only track pressure waves in the axial direction, unless the charge fully fills the volume behind the projectile, it is difficult to completely match the physics of the firing. This occurs because the presence of solids and gases in the chamber is generally not uniform—the solids are usually at the bottom of the chamber. This affects the gas dynamics. Solids will also be entrained by the gas flow down the bore and some modeling of their motion has to be accomplished (or ignored). In most cases, the propellant bed is assumed to be a monolithic mass that © 2014 by Taylor & Francis Group, LLC 104 Ballistics: Theory and Design of Guns and Ammunition regresses and stretches as the propellant is burned. These codes are usually very good but the user should completely understand the assumptions on how the propellant is allowed to move before using them. A two-dimensional model is one where the density can vary in the radial direction as well. These models are better at predicting pressure waves but take somewhat longer time to run than one-dimensional models. Pressure can be tracked in the radial direction and the propellant motion included. The same issues with propellant motion are present as they were in the one-dimensional models though it is possible to track propellant motion. A three-dimensional model has it all. Because of this they usually take an excruciatingly long time to set up and run. This time constraint makes them generally reserved for failure investigations rather than predictive simulations. Individual propellant grains can regress and be tracked and one can imagine the difficulty with this in the sense of model validation. With suitable stress and failure models, grain fracture can also be examined. If erosion models are incorporated, the effect of gas wash on propellant burn rate can even be included. One has to ask oneself if all of this is really necessary. In some cases, these models are crucial, in other cases, they are certainly overkill. The usefulness of this type of model is still somewhat limited by computer speed, but as computers become faster the limitation will change to a lack of accurate physical models for motion, surface regression, propellant and gun tube erosion, grain fracture, etc. These issues are certainly solvable, but finding a proponent who will fund the research is difficult. Now that we have described the general types of models, it is important to explain their use further. In general, all of them are used in a similar manner. We shall use the zero-dimensional model as an example and leave the rest to the reader’s imagination (and budget restrictions). Typically, a propellant formulation and geometry is chosen as a point of departure given that we have a preliminary gun design and a projectile to work with. This propellant is then further developed in terms of geometry or chemical composition. Some zero-dimensional codes are provided with optimization subroutines so that particular characteristics of the ballistic cycle can be achieved. The pressure– time, acceleration–time, and pressure–distance curves are examined and, if suitable, some experimental charges are made up. The configuration is then fired and the results checked against the code. These results then can be used to adjust burn rates and resistive characteristics, and the model can be used to predict all future firings and design iterations. A particular example of the power of these codes is their usefulness in assessing the interior ballistics of systems that vary widely in matters of scale, e.g., in mass of projectile, diameter of bore, and muzzle velocity. In the 1960s, ballisticians J. Frankle and M. Baer at the Ballistics Research Laboratories at Aberdeen, Maryland [8] and others elsewhere devised codes largely based on Corner’s zero-dimensional analysis that we described in detail in Section 3.2. Among these the Frankle–Baer simulation, still in use today, which examined and expanded on the basic energy equation, Energy released by burning propellant = change in internal energy of the gases + work done on the projectile + secondary losses or Q = ∆U + W + Losses © 2014 by Taylor & Francis Group, LLC (3.165) Analytic and Computational Ballistics 105 developed equations of state of the propellant gases based on more recent thermodynamic theories and refined the losses term from new experimental data. This led to more refined ratios for breech, mean, and shot base pressures, and more accurate equations of motion for the projectile. To examine the effects of scale we computed the relevant pressure ratios for three widely different gun–projectile combinations. We show these combinations and the resultant ratio values in Tables 3.1 through 3.5. What is noteworthy is the applicability of the theory over the range of size, projectile mass, and propellant type and volume. Notice also the closeness of the pressure ratios for each projectile between the Corner and the Frankle– Baer simulations. TABLE 3.1 Inputs for Comparison of Corner and Frankle–Baer Parameter Expression or Value (J. Corner) M735     M1   M193    M735     M1   M193    Expression or Value (Frankle–Baer) Charge weight c  13.125     9.000   4.020 × 10 −3   c  13.125     9.000   4.020 × 10 −3   Projectile weight w  12.78     31.97  − 3 7.86 × 10   wp  12.78     31.97  − 3 7.86 × 10   Propellant type —  M30     M1   Ball    —  M30     M1   Ball   TABLE 3.2 Burn Characteristic Inputs for Numerical Comparison of Corner and Frankle–Baer Parameter Expression or Value (J. Corner) M735     M1   M193    Expression or Value (Frankle–Baer) M735     M1   M193    Propellant impetus (force) λ  3.64 × 10 5   5  3.05 × 10   3.32 × 10 5    λ  3.64 × 10 5   5  3.05 × 10   3.32 × 10 5    Specific heat ratio γ 1.2385    1.2592   1.26    γ 1.2385    1.2592   1.26    Polytropic index 1 γ −1 © 2014 by Taylor & Francis Group, LLC n Ballistics: Theory and Design of Guns and Ammunition 106 TABLE 3.3 Pressure Gradient Calculations for Numerical Comparison of Corner and Frankle–Baer Parameter Expression or Value (Frankle–Baer) Expression or Value (J. Corner) p pS 1+ 1 δ n/a 1 ab n/a  2n + 3 2(n + 1)  +   c / wp   δ p pB  1 c 1 − 6 w     1 c 1 −   δ  wp pB pS  1 c 1 + 2 w    −( n+1)   (1 − ab )  c 3w 1+ 1 c δ wp  1 + c1 β n   1  1 + α n   2n + 3   1 + c1n    n+1   (1 − ab )   TABLE 3.4 Specific Frankle–Baer Computations for the M735, M1, and M193 Projectiles Projectile M735 (105 mm KE) M1 (105 mm HE) M193 (5.56 mm ball) α β c1 1 δ 0.56 0.63 0.60 1.07 1.01 1.03 1.05 1.02 1.04 0.333 0.322 0.315 ε= c wp 1.027 0.282 0.511 1 ab (1−ab)n+1 11.002 37.811 22.325 0.683 0.876 0.800 TABLE 3.5 Specific Gradient Comparison of Corner and Frankle–Baer for the M735, M1, and M193 Projectiles p pB p pS Projectile M735 (105 mm KE) M1 (105 mm HE) M193 (5.56 mm ball) 1+ 1 c 2w Corner Frankle–Baer Corner Frankle–Baer Corner Frankle–Baer 1.342 1.094 1.170 1.342 1.091 1.161 0.829 0.953 0.915 0.917 0.956 0.929 1.514 (1.619) 1.141 (1.148) 1.256 1.464 (1.463) 1.142 (1.141) 1.250 (1.250) 3.5 Sensitivities and Efficiencies Having explored the detailed development of theories of interior ballistic events, we will now probe the outcome of varying some of the parameters that are under the control of the charge designer. To do this, we will be referring back to definitions and equations developed under Section 3.2. © 2014 by Taylor & Francis Group, LLC Analytic and Computational Ballistics 107 A useful quantity for our analysis is the dimensionless central ballistic parameter, M   c   1+ AD  3w1  M= 2 w1cλβ 2   c    1+ 2w1     2 2 (3.166) Of particular importance in this are the variables D and β, the original web dimension and the burning rate coefficient, respectively. If we examine Equation 3.167 pB m c  1+ λc  2 w 1 =  Vi  1 + c  3w1     1     M exp[1]    (3.167) We can see that, at least for the case of θ = 0, M is in the denominator and as the ratio D/β decreases, M decreases and from Equation 3.167, the peak pressure, pB, increases. That is, if the original web size is decreased, the peak pressure will increase. This is a parameter much under the control of the designer. Referring again to Equation 3.166, we see that if the charge mass (weight), c, is increased, then M decreases (the cs in the gradient term largely cancel out and c in the first term denominator governs). In Equation 3.167, c appears in the numerator and M in the denominator causing pB, the peak pressure, to again rise. Let us now examine the shift in location of xm, the point in travel where the peak pressure exists: M  ( M + 2θ )  θ xm + l = l    (M + θ )  (3.168) Equation 3.168 relates xm to M. If the ratio D/β or the charge mass, c, decreases, then M decreases and consequently xm is reduced (it moves toward the breech). This kind of shift is important in gun design since wall thickness and center of mass are important considerations for weapon mounting. The sensitivity of muzzle velocity, V, to changes in web size or charge weight can be seen in Equation 3.169: V 2 ( x ) − V 2 ( xc ) = λ c( xc + l)exp[ − M] Φ c  l  w1 +  3  (3.169) where M is the governing term and is decreased as we showed earlier, if charge mass is increased or web size reduced. Because M has a negative exponent in the equation, its reduction drives an increase in V. Finally, the influence of travel on muzzle velocity can be shown to be quite weak. The computation is complex and will not be shown here. But, e.g., by doubling the travel, velocity increases only by a factor of about a tenth, hardly worth the effort in the real world. © 2014 by Taylor & Francis Group, LLC Ballistics: Theory and Design of Guns and Ammunition 108 pB m p FIGURE 3.5 Average and maximum breech pressure for a typical gun firing. There are two measures of efficiency that are of interest to the interior ballistician: piezometric or pressure efficiency and ballistic or energy efficiency. Piezometric efficiency, εp, is the ratio of the average pressure during the entire ballistic cycle to the peak pressure during the cycle: εp = p pB m (3.170) An illustration of the space-mean pressure and maximum breech pressure is provided as Figure 3.5. Increasing εp implies that the muzzle pressure will be high (usually an undesirable trait), and that the charge burnout point will move toward the muzzle (hopefully never outside the muzzle). High piezometric efficiency usually means poor regularity, i.e., round-to-round muzzle velocity repeatability is poor (an undesirable trait). For powerful, high-velocity cannons, this efficiency is usually in the 50%–60% range. Other cannons are lower. High piezometric efficiency also implies that the expansion ratio, the ratio of total gun volume to chamber volume, will be low: powerful guns have large chambers and consume lots of propellant. Ballistic efficiency, ε b, is defined as the ratio of the kinetic energy of the projectile as it exits the muzzle to the total potential energy of the propellant charge: εb = muzzle KE = propellant PE wV 2 (γ − 1)wV 2 = λc 2λ c γ −1 1 2 (3.171) because the potential energy is defined as propellant PE = RT0 (γ − 1) and λ = RT0 (3.172) Increasing ε b tends to shift the all-burnt position toward the breech and increases the expansion ratio. Reducing the central ballistic parameter, M, by going to a smaller web will also increase ε b. The ballistic efficiency of most guns is approximately 0.33. © 2014 by Taylor & Francis Group, LLC Analytic and Computational Ballistics 109 References 1. 2. 3. 4. 5. Robbins, F., Interior Ballistics Course Notes, Self Published, Aberdeen, MD, 2002. Panton, R.L., Incompressible Flow, 2nd edn., John Wiley & Sons, New York, 1995. Currie, I.G., Fundamental Mechanics of Fluids, 2nd edn., McGraw-Hill, New York, 1993. Corner, J., Theory of the Interior Ballistics of Guns, John Wiley & Sons, New York, 1950. Bull, G.V., Murphy, C.H., Paris Kanonen—The Paris Guns (Wilhelmgeschutze) and Project HARP, Verlag, E.S. Mittler & Sohn GmbH, Herford und Bonn, 1988. 6. Campbell, J., Naval Weapons of World War Two, Naval Institute Press, Annapolis, MD, 1985. 7. Jordan, J. (Ed.), Warship 2012, Conway Publishing, London, U.K., 2012. 8. Frankle, J.M., Interior Ballistics of High Velocity Guns Experimental Program, Phase I, BRL Memorandum Report 1879, U.S. Army Ballistic Research Laboratory, Aberdeen Proving Ground, MD, November 1967. © 2014 by Taylor & Francis Group, LLC © 2014 by Taylor & Francis Group, LLC 4 Ammunition Design Practice Chapter 3 provided us with the information necessary to determine the forces acting on the projectile and gun. This chapter endeavors to describe techniques necessary for the projectile or weapon designer to be successful. Sections 4.1 and 4.2 describe topics in the field of mechanics of materials. This material will form the basis by which we will evaluate designs. Sections 4.3 through 4.8 apply these concepts to the design of projectiles and guns. This chapter ends with practices and techniques used to design modern ammunition that must be fired from a gun. 4.1 Stress and Strain Before proceeding with our examination of design practices, a discussion of the fundamentals of the general state of stress in materials is in order. Consider an arbitrary cube of material under load as depicted in Figure 4.1. The state of stress can be completely defined by six stress components: σx, σy, σz, τxy, τyz, and τzx. Here, we have used a Cartesian coordinate system where the normal stresses are denoted by σ and the shear stresses are denoted by τ. The first subscript represents the plane in which the stress acts (defined by its normal vector) while the second subscript indicates the direction of action. These components form the stress tensor, which is actually a 3 × 3 matrix of nine elements; except that we have assumed that τxy = τyx, τzy = τyz, and τxz = τzx. When written as a tensor, the state of stress in a material is defined as σ x  σ = τ xy τ zx τ xy σy τ yz τ zx   τ yz  σ z  (4.1) It can be shown that the coordinate system in which we measure the stresses can be rotated so that the shear stresses vanish. The three remaining stresses are normal stresses, known as the principal stresses, and are denoted as σ 1, σ 2, and σ 3. These stresses are important because, regardless of what coordinate system we view the component in, the stress state is uniquely determined. Also, in some materials, these stresses are associated with failure and fracture. These points are sometimes shown graphically through use of Mohr’s circle. The determination of the principal stresses will be discussed later in this section. 111 © 2014 by Taylor & Francis Group, LLC Ballistics: Theory and Design of Guns and Ammunition 112 σy σz τyx τyz τxy τzy σx σx τzx τxz σz σy FIGURE 4.1 Cartesian stress components. It is also very important to understand this when we try to examine the stress levels in a part experimentally with a strain gage. Stress is a point function defined by force per unit area expressed as σ= F A (4.2) where σ is the stress F is a force A is the cross-sectional area of the component The same equation also holds if we use the symbol τ signifying a shear stress. When we examine a structure, we normally are given the loads that are imposed on it. We then either choose a material or evaluate a given material to see how it will behave under the applied loads. This process requires us to convert the external loads to stress. These stresses will cause movement of the material in the form of either stretching (tension) or compression. This movement is the actual displacement of the material. There is an intermediate analytical step between these two where we need to define the strain of the material. The strain in the material is defined as the change in length of a part over its initial, unstressed length. Mathematically, this is expressed as ε= ∆l l (4.3) We require a relationship between stress and strain to evaluate material behavior under a load. The link between stress and strain is called a stress–strain relationship. The most common and simplest stress–strain relationship is that for a linear-elastic material. This is known as Hooke’s law and is given for small deformations and uniaxial loading by ε= © 2014 by Taylor & Francis Group, LLC σ E (4.4) Ammunition Design Practice 113 where E is the modulus of elasticity, sometimes known as Young’s modulus. In a linearelastic material, any loading and unloading of the structure occurs along a curve in stress– strain space that has a slope equal to the modulus of elasticity. Under the assumption of general loading, material will be “pulled in” in the transverse directions as it is stretched longitudinally. The ratio of lateral strain to axial strain is denoted as ν and called Poisson’s ratio and is given for an isotropic material as ν =− εy ε =− z εx εx (4.5) This assumption of general loading changes our Hooke’s law relation as follows: εx = σ x νσ y νσ z − − E E E (4.6) εy = − νσ x σ y νσ z + − E E E (4.7) εz = − νσ x νσ y σ z − + E E E (4.8) While we have defined ε to represent longitudinal strain in a material, a different type of strain can be examined—shear strain. Shear strain, γ, is defined as the angular deviation of a material from its original, undeformed shape. Shear strain is given by its own version of Hooke’s law as γ = τ G (4.9) where G is known as the shear modulus of the material. In an isotropic material, E, ν, and G are not independent. The relationship that links them is G= E 2(1 + v) (4.10) When we perform hand calculations, it is customary to convert the loads to stresses, then the stresses to strains, and finally strains to deformations. The process is somewhat different (i.e., reversed) in a finite element analysis (FEA). The determination of the principal stresses is important in several failure criteria. When a part is being examined experimentally during a gun launch, it is customary to utilize a strain gage. A strain gage measures the change in a parts length using the fact that resistance increases in a conductor as it is stretched. Strain gages are not always placed along the directions in which it is desired to compute stress however. Since strain gages only measure in-plane stress, it is common to transform this two-dimensional measurement © 2014 by Taylor & Francis Group, LLC Ballistics: Theory and Design of Guns and Ammunition 114 σy y y´ σy΄ τxy x´ θ τx΄y΄ σx΄ σx x FIGURE 4.2 Transformation of stress components. into a desired in-plane direction. To transform stress from the strain gage coordinate system to the desired coordinate system, we use the following equations: σ x′ = σ x cos 2 θ + σ y sin 2 θ + 2τ xy sin θ cos θ (4.11) σ y ′ = σ x sin 2 θ + σ y cos 2 θ − 2τ xy sin θ cos θ (4.12) τ x′y ′ = τ xy (cos 2 θ − sin 2 θ ) + (σ y − σ x )sin θ cos θ (4.13) In each of these equations, the primed variables are those in the desired direction and the unprimed variables are those measured by the strain gages. This is depicted in Figure 4.2. Rotation of coordinate systems in three dimensions is covered in excellent detail in Ref. [1]. It was stated earlier that a rotation can be made such that the shear stresses vanish and this results in what are known as principal stresses [2]. To determine the values of the principal stresses, we determine the stress invariants through solution of the eigenvalue problem. The three stress invariants are given by I1 = σ x + σ y + σ z (4.14) 2 2 I2 = σ xσ y + σ yσ z + σ zσ x − τ xy − τ yz − τ z2x (4.15) 2 2 2 I 3 = σ xσ yσ z − σ xτ yz − σ yτ zx − σ zτ xy + 2τ xyτ yzτ zx (4.16) Once these invariants are obtained, the principal stresses are obtained through I1 2 2 I1 − 3 I 2 cos φ + 3 3 (4.17) σ2 = I1 2 2 2π  I1 − 3 I 2 cos  φ + +   3 3 3  (4.18) σ3 = I1 2 2 4π  I1 − 3 I 2 cos  φ + +   3 3 3  (4.19) σ1 = © 2014 by Taylor & Francis Group, LLC Ammunition Design Practice 115 In Equations 4.17 through 4.19, the quantity 𝜙 is calculated through φ=  2I 3 − 9I I + 27 I  1 1 2 3  cos −1  1 3 3  2 ( I 2 − 3I ) 2  1 2   (4.20) Now we have all of the basic information necessary to discuss failure criteria. Limits of space prevent a more in-depth treatment of this topic. The reader is referred to the references at the end of this chapter for a more detailed treatment. Problem 1 For the state of stress below, find the principal stresses and the maximum shear stress.  20  [σ ] = 15  0 15 4 0 0  0  [MPa] −9  Answer: τmax = 19 [Mpa] 4.2 Failure Criteria When embarking on the design of a particular projectile component, we must initially determine certain characteristics of the material contemplated for the design: Will we use a metal or a plastic? Does it have a distinct yield point? Is it brittle or very ductile? Such determinations will govern which criteria we use when we calculate the stresses that will cause failure of the component. There are three commonly used criteria for yield or failure: von Mises, which is also known as the maximum distortion energy criterion; Tresca, which is known as the maximum shear stress criterion; and Coulomb, which uses a maximum normal stress criterion. Other materials may require unique failure criteria, for example, composites or nonisotropic metals may require Tsai–Wu or Tsai–Hill criteria. The von Mises or maximum distortion energy criterion is used typically when the component is to be made of metal. It assumes that the energy required to change the shape of the material is what causes yielding and that a hydrostatic state of stress will not result in failure. The materials for which it is used should have a distinct yield point. Our convention shall follow that of structural engineers in which we shall assume tensile stress to be positive. By this criterion, we assume that the distortion of the material will precipitate the failure. We shall order the stresses with 1 as largest to 3 being smallest and state the following: (σ 1 − σ 2 )2 + (σ 2 − σ 3 )2 + (σ 3 − σ 1 )2 = constant © 2014 by Taylor & Francis Group, LLC (4.21) Ballistics: Theory and Design of Guns and Ammunition 116 σ2 If stress state falls in this region, the component is OK +σY –σY +σY –σY σ1 This is a 2D representation of an ellipsoid which also includes the σ3 (out of the plane of the paper) direction FIGURE 4.3 von Mises failure surface. We set this constant equal to 2σ Y2 or 6K 2. Here σ Y is the yield stress in simple tension and K is the yield stress in pure shear. This implies that 1 2 σY = K2 3 (4.22) or K= 2  σY   σY     = 1.155  3 2   2  (4.23) σ Y is also known as the equivalent stress and either σ Y or K can be found experimentally. In σ 1–σ 2–σ 3 space, the criterion is represented by an ellipsoidal surface whose inner region symbolizes stress states that are safe (nondistorting). This is shown two dimensionally in Figure 4.3. The Tresca or maximum shear stress criterion is used when the material is known to have great ductility. It assumes the failure mechanism is by slippage along shear planes generated by the shear stress in the material. This assumption says that the material will not fail unless the shear stress it is experiencing is greater than that exhibited by a tensile test specimen of the same material at its failure point. Again we assume that tensile stress is positive and order the stresses with 1 the largest to 3 the smallest and state the following: (σ 1 − σ 3 ) = constant 2 (4.24) We set this constant equal to σ Y2 or K. Here σ Y is the yield stress in simple tension and K is the yield stress in pure shear. This implies that for a component not to exhibit failure (σ 1 − σ 3 ) < σ Y (4.25) σ Y = 2K (4.26) and © 2014 by Taylor & Francis Group, LLC Ammunition Design Practice 117 σ2 If stress state falls in this region, the component is OK +σY –σY σ1 +σY –σY This is a two-dimensional representation of a polyhedron which also includes the σ3 (out of the plane of the paper) direction FIGURE 4.4 Tresca failure surface. as well as (σ 1 − σ 2 ) < σ Y , (σ 1 − σ 3 ) < σ Y , and (σ 2 − σ 3 ) < σ Y (4.27) Once again either σ Y or K can be found experimentally. In σ 1–σ 2–σ 3 space, this is represented by a polyhedral surface whose inner region includes all stress states that are safe (nonfailing). This is shown as a 2D sketch in Figure 4.4. The Tresca criterion is slightly more conservative if used for metals than von Mises. The Tresca polyhedron is contained within (circumscribed by) the ellipsoid of von Mises. The third failure criterion we will examine is the Coulomb or maximum normal stress criterion. Here, we assume that the normal stress in the material will precipitate the failure. Tensile stress is again assumed to be in the positive direction and stresses from 1–3 are again in order of decreasing magnitude. In this criterion, we require that for a material that does not exhibit failure σ 1,σ 2 ,σ 3 < σ U (4.28) That is, all of the principal stresses must be less than the ultimate stress, σ U, in the material in that particular direction. Recall that we use this for brittle materials where there is no yield point or yielding behavior. The failure surface is a rectangular polyhedron whose edges are the ultimate stresses in each principal direction. Stress levels within the polyhedron will not cause failure. A two-dimensional representation is depicted in Figure 4.5. Even though this Figure is shown as a square, in many materials the compressive strength is much greater than the tensile strength, resulting in different limits and thereby changing the appearance (and sometimes resulting in a name change as well to a Mohr–Coulomb criteria) of the failure surface. In this instance, the failure surface would look like Figure 4.6. In the Mohr–Coulomb failure criteria, a greater compressive normal stress allows the material to carry more load. This is caused by the locking of slip planes akin to the sliding friction of a block causing greater resistance when the block gets heavier (i.e., an increase in normal stress on the slip plane). When this is applicable, our criteria result in an equation for the failure surface as follows: max[|τ |−λ ⋅ (σ )] = σ E © 2014 by Taylor & Francis Group, LLC (4.29) Ballistics: Theory and Design of Guns and Ammunition 118 If stress state falls in this region, the component is OK σ2 σUT σUC σUT Here σUC is the ultimate stress in compression and σUT is the ultimate stress in tension σ1 σUC FIGURE 4.5 Coulomb failure surface. σ2 If stress state falls in this region, the component is OK This is a two-dimensional representation of a polyhedron which also includes the σ3 (out of the plane of the paper) direction σYT σYC σYT σ1 Here σYC is the yield stress in compression and σYT is the yield stress in tension σ1 + σ2 = 0 σYC FIGURE 4.6 Mohr–Coulomb failure surface. This equation results in a greater stress to failure due to the internal friction coefficient, λ. Since compressive strength is negative and λ is a positive quantity, the equivalent failure stress, σ E, is greater with greater normal stress, σ. Occasionally, it will be essential that we combine two or more of these criteria due to a change in material behavior. We shall describe this in due course. Problem 2 A component has principal stress values of 20,000, 56,000, and −220,000 psi (note that negative means compressive stress), if the yield strength in a simple tension test of the material was found to be 180,000 psi, will the part survive based on the von Mises failure criteria? Answer: No the part will fail. © 2014 by Taylor & Francis Group, LLC Ammunition Design Practice 119 4.3 Ammunition Types Just as weapons are categorized by their usage as guns (low angle, line of sight, direct fire), howitzers (high angle, beyond line of sight, indirect fire), or mortars (very high angle, short range, indirect fire), the munitions for them are also categorized, not by use, but by their construction or assembly methods. Ammunition can be fixed, separable, or separate loaded. Fixed ammunition, usually called a cartridge, consists of a container for the propellant charge, called the cartridge case, that is firmly attached to the projectile by crimping or cement and that remains in the weapon after firing and is ejected near it or is consumed during firing, and the projectile that flies downrange to the target. The charge, priming, and ignition system are assembled inside the case and are not alterable. This type of ammunition is characteristically used in tank, antiaircraft, aircraft weapons, and in most small arms (rifles and pistols). Separable ammunition (also called semi-fixed ammunition) also consists of the cartridge case and projectile, but the case is not attached firmly to the projectile and can be removed in the field to adjust the charge, which can be changed incrementally. This type of ammunition was used in older howitzers and is still used in shotguns. Separate-loaded ammunition (sometimes called separated ammunition) consists of the projectile, which is loaded first into the weapon, the propellant charge loaded next, and finally the primer and igniter loaded last. The charge, which is supplied to the weapon site, is in bagged increments and is altered, along with the quadrant elevation of the weapon, to vary the range. The primer is usually loaded into the weapon’s breechblock. The block is self-sealing and assumes this function, which in fixed ammunition is done by the cartridge case. Ammunition of this type is used in howitzers and large naval guns. Mortar ammunition is essentially of the separated type. The charge is incremental to help vary the range by altering the muzzle velocity. The charge increments are held in place on the projectile body by clips or holders. Increments may be added or deleted in the field by the gunner. Priming is done through an integral attachment to the projectile (a boom). Primer initiation is by a firing pin in the weapon that strikes an initiator in the boom at the termination of the fall of the projectile as it is dropped down the tube from the muzzle end. Trigger firing is also possible in some weapon designs. The practical design of fixed ammunition cartridges, which is what we will mainly dwell on, encompasses the design of the propellant charge ignition system, the construction of the main body of the propellant charge, the design of the projectile body itself, including its shape and mass distribution, and its obturation and stabilization components. The design must also incorporate into the projectile the ancillary systems necessary for its intended functioning, for example, fuzes, expulsion charges, explosive trains, and in modern projectiles, guidance, and control. 4.4 Propellant Ignition Energetic devices that are combined in a specific manner into the ignition train accomplish the initiation of combustion of the propellant charge. The first of these elements is the highly sensitive, detonator cup filled with material such as lead azide, which itself receives an energy pulse from a trigger mechanism that delivers the pulse in the form of a mechanical (spring actuated) or electrical (hot wire or laser) impulse. The sensitive mix © 2014 by Taylor & Francis Group, LLC Ballistics: Theory and Design of Guns and Ammunition 120 is detonated by the impulse and flashes into the main, less sensitive ignition charge. This material is contained in the primer head. Secondary ignition takes place in the main primer body, where the ignition material known as the primer charge is stored. This material has traditionally been fine-grained black powder, which is known to have certain undesirable properties such as hygroscopicity. Attempts have been made to replace black powder, but it still remains the chief secondary ignition material. Two basic forms of primer charge are used in large caliber munitions: flat base-pad igniter charges are used with separate-loaded bagged propellant charges and in fixed, stick propellant charges; central core or bayonet-type primer bodies are used in most fixed, loose, granular propellant charges. The design goal of all ignition systems is to provide rapid but smooth ignition of the main propellant charge avoiding at all cost pressure surges or spikes. Such surges can crush individual grains or sticks causing large, uncontrolled increases in burning surfaces and uncontrolled burning of the main charge. Symptoms of such burning are negative delta pressure (−∆p) waves, that is, negative gradients of pressure along the length of the chamber. One cause of pressure surges are the so-called blind primers, where vent holes are missing along the length of the primer body tube. The pressure build-up in the tube can rupture it causing asymmetric ignition and a −∆p. Other caveats are to avoid overly sensitive detonator mixes and to provide gas flow space in the main propellant charge. Ignition and burning are surface phenomena and too tightly packed charges do not provide the necessary surfaces. 4.5 The Gun Chamber To the rear of the long cylindrical portion of the gun (the bore) is the chamber, shown in Figure 4.7. The tapers shown facilitate the removal rearward of the spent cartridge case that hugs the chamber wall. During the firing cycle, the case swells because of the internal pressure and firmly contacts the chamber wall sealing the gases from exiting rearward. When the pressure decays, a properly designed case comes away from the wall and the tapers insure that it does Tapers greatly exaggerated Bottom of groove Top of land Dc D Forcing cone Shoulder Rear face of tube FIGURE 4.7 Chamber geometry. © 2014 by Taylor & Francis Group, LLC Ammunition Design Practice 121 not stick in the chamber. When a fixed round of ammunition is loaded into the chamber, the rear face of the tube provides the stop and seat for the rim of the case. During the expansion of the pressurized case, the forcing cone of the chamber forms the seal for the hot gases by the extrusion and engraving of the rotating band in a rifled bore or the extrusion of the obturating band in a smooth bore. The ratio Dc/D is known as the chambrage, an important characteristic of the design. Large values of the chambrage tend to cause turbulent flow of the gases as they enter the bore. Such turbulence contributes to the erosion of the bore surfaces. The gun designer is caught in a curious bind: for a desired volume of propellant, a large chambrage provides a shorter cartridge length, frequently a highly desirable parameter in the tight confines of a turret, for example; on the other hand, large chambrage values subject the bore to more erosion. Some of this difficulty has been overcome by the use of erosion reducing coolants. It has been found that much of the erosive wear in high-performance guns and howitzers can be ameliorated by the introduction of a cool liquid, gaseous, or particulate layer between the hot propellant gases and the bore. Materials such as titanium dioxide, wax, talc, or silicone oil have proven efficacious. If these materials are assembled in the body of the propellant charge so that the gas flow keeps the coolant at the bore wall, a substantial decrease of erosion results. This is called laminar flow and is observed in low chambrage guns. Thus, a compromise may have to be made in the chambrage to reduce the turbulent flow. 4.6 Propellant Charge Construction In fixed cartridges, the most common practice is to fill a metallic cartridge case with perforated granular propellant grains around a bayonet-type primer that has already been inserted in the case. The grains commonly have seven perforations for progressive burning. In high-performance rounds, vibrating the case to help settle the grains maximizes the loading density of the charge. Tank munitions are often loaded with perforated stick propellant. The sticks are bundled and carefully laid up around the boom and fin components that intrude into the depth of the case. Supplementary granular propellant is occasionally added to the stick bundles to further boost the charge mass and increase the progressivity of burning. Rocket grain configurations with complex star and slit perforations have been tried as well as 19-perf grains to raise the burning rate, but these are difficult to make and are not standard. Howitzer (separate-loaded) charges are made up of bagged increments that are ignited by the last increment loaded, the base-pad igniter. A primer in the breechblock sets off the igniter. In these and in the fixed ammunition charges, coolants are strategically emplaced to promote erosion resistance. With bagged propelling charges, since there is no cartridge case present, it is extremely important that all of the material be combusted. Great care is taken in selecting materials—silk was used for many years in the Navy—to assure that there are no burning embers left in the weapon after it was fired. It is typical for a howitzer crewman to look down the bore and shout “bore clear” during firing operations. If burning materials are present and a fresh charge is inserted into the bore, the propellant may ignite and cause serious injury to the gun crew. This has been termed “cook-off.” There have been extensive efforts to take advantage of the convenience of stowage, low cost, and inherent safety of liquid bipropellants (LP). However, severe operational and performance problems have prevented their adoption. These problems have centered on combustion instability that manifests itself in destructive, unpredictable pressure peaks, particularly in bulk-loaded systems. Attempts to get around these so-called Taylor instabilities © 2014 by Taylor & Francis Group, LLC Ballistics: Theory and Design of Guns and Ammunition 122 have had some success with regenerative pressurized systems that atomize the pumped-in liquids, ignite this cloud, and avoid the pressure wave unpredictability of an ignited bulk of liquid. This concept, even though it has shown promise, still may not be able to overcome the poor low-temperature properties of the liquid propellants. They show marked increases in viscosity at low temperatures causing severe flow and pumping problems. Two other concepts of gun propulsion should be mentioned. These are the use of electromagnetically generated force to propel a projectile down a gun and the idea of using a low molecular weight gas to propel the projectile—the light gas gun. At the time of this writing neither concept has shown the ability to progress beyond the laboratory stage to a fieldable weapon, although light gas guns are in common use in laboratories to reach velocities with small projectiles approaching meteorite entry speeds. 4.7 Propellant Geometry The geometry of the propellant grain is one of the parameters available to the interior ballistician to tailor the pressure curve in the gun. Production of gas from a grain depends on the evolution of the total surface of the grain as the burning proceeds. If the surface area increases with time, the grain is considered progressive. If the total surface remains constant over time, the grain is neutral, and if the surface decreases with time the grain is considered regressive. The perforations in the grain affect the surface area and therefore the burning characteristics. In cylindrical grains, the number of perforations is usually one of the numbers in the sequence: 1, 7, 19, and 37. The largest number in use in the United States is 19, and this is rarely found because of the difficulty of manufacture. The various types of grains are shown in Figure 4.8. The web, D, that is the smallest thickness of propellant between any two surfaces is one of the major parameters in interior ballistic computations. D fD D Single perf. grain—neutral fD Seven perf. grain—progressive W D D << 1 W 4 Strip or flake—regressive Ball—regressive FIGURE 4.8 Typical propellant grain geometries. © 2014 by Taylor & Francis Group, LLC Ammunition Design Practice 123 4.8 Cartridge Case Design The design of a metallic cartridge case must fulfill four basic roles: the case must seal or obturate the gun breech so that gases do not stream backward out of the gun; it must serve as a protective container for the propellant charge; it must act as a structural member of the cartridge assembly to allow for vigorous handling during shipping, stowage, and loading into the chamber; and it must be easily extractable from the chamber after the round is fired. Metallic cases have been used for much more than a hundred years and the design practices are well established to fulfill these roles. Yet difficulties still arise in the extraction of the case after firing—it can stick in the chamber, rendering the weapon useless until it is removed. The case by itself cannot sustain the gun pressure and is intended to be supported by the chamber walls. Yet, the case must be designed with sufficient clearance to permit loading and ramming. The analysis of sticking that follows must be part of the design engineer’s overall task before a new weapon can be fielded. It is possible, through use of some relatively simple equations, to determine if a cartridge case will expand enough to stick in the chamber of the weapon after firing. Graphically, we can depict this as shown in Figure 4.9. In this figure, we see the effect when a case with a low yield strength is loaded to the same levels as a good case. The expansion and contraction of the gun tube itself must be taken into account when the cartridge case is designed. This condition can be approximated using a bilinear, kinematic hardening model where the stress–strain curves of the case material are modeled, as depicted in Figure 4.10. The first step in this procedure is to model the gun tube. In this case, we assume that the material is perfectly elastic—which will be the case for any properly designed tube—and we can determine the radial expansion through [2] utube = a′ (1 −ν )( p1a′2 − p2b 2 ) + (1 + ν )b 2 ( p1 − p2 ) Etube (b 2 − a′2 )  (4.30) In this equation (which has been tailored from a previous formula for a thick-walled cylinder because the point we are interested in is on the inside radius of the tube wall), Hoop stress in case wall Clearance between case and chamber before firing Elastic expansion of the chamber on firing Yp Normal case Yp Low yield case Expansion (strain) Low yield case interferes with chamber by this amount FIGURE 4.9 Stress–strain diagram of a normal case and one with low yield strength. © 2014 by Taylor & Francis Group, LLC Ballistics: Theory and Design of Guns and Ammunition 124 Hoop stress in case wall Clearance between case and chamber before firing Elastic expansion of the chamber on firing Tangent modulus Modulus of elasticity Modulus of elasticity Expansion (strain) Low yield case interferes with chamber by this amount FIGURE 4.10 Stress–strain diagram of a normal case and one with low yield strength modeled as bilinear kinematic hardening materials. a′ is the inner radius of the chamber, b is the outer radius of the gun tube, p1 is the internal pressure, p2 is the external pressure (usually conservatively taken as 0), ν is Poisson’s ratio for the tube material, and Etube is the modulus of elasticity. We now calculate the stress, strain, and displacement of the case through use of the thinwall cylinder equations [3] a 2 p1 Ecase h (4.31) σ θθ = ap1 h (4.32) εθθ = σ θθ Ecase (4.33) ucase = In these equations, ucase is the radial expansion of the case, σ θθ is the hoop stress in the case, εθθ is the hoop strain, a is the outside radius of the case, and h is the case wall thickness. Now, the gun tube will stop the case from expanding further once contact is made so the maximum expansion of the case will be as follows: rcase + ucasemax = rtube + utube = rcase + aεθθmax (4.34) Because we know the pressure and the tube dimensions and therefore the value of utube, we can calculate εθθmax. We can then use this value to calculate the stress in the case at the maximum expansion: εθθmax − ε Y = σ θθmax − σ Y Ecase-tangent (4.35) In this expression, the subscript Y indicates yield values and Ecase-tangent is the tangent modulus of the cartridge case material. Once we determine the stress at the maximum © 2014 by Taylor & Francis Group, LLC Ammunition Design Practice 125 expansion, we need to recall that a material which has yielded will retract along its original elastic modulus. Thus, we can write ε return = σ θθmax Ecase (4.36) Now the residual strain in the case is given by ε residual = εθθmax − ε return (4.37) We can then find the permanent radial displacement through uresidual = aε residual (4.38) If we now add uresidual to the original radius of the case, a, we can see that if uresidual + a ≥ a′, the case will stick (4.39) uresidual + a < a′, the case will not stick (4.40) or if Over the last 20 years, the metallic case of drawn brass, extruded steel, or spirally wrapped steel has been replaced in certain systems by a fully combustible or consumable case. These cases are manufactured of felted nitrocellulose and usually consist of a base and sidewall that are assembled with cement, filled with granular or stick propellant, and attached to the projectile with clamps and cement. Since the cases are consumed completely, they do not seal the breech. With these munitions, a self-sealing breech must be designed for the weapon. For guns with nonsealing breeches that are already fielded for use with conventional metallic cartridge cases, a case that has a metallic stub and a combustible sidewall has been devised to take advantage of the small volume of the ejected stub in the confines of a tank turret, e.g., and the overall reduction in cost and weight of the round. While systems with the combustible case have been fielded, the success of this development has not been complete. Occasional problems with incomplete combustion of a case that leaves smoldering residue capable of igniting the next loaded round (cook-off) have required scavenging systems for the chamber to be installed. The inherent structural weakness of nitrocellulose has also posed problems of case attachment and handling. Yet the obvious advantages of the combustible case have kept the concept in the weapon designer’s toolbox for possible use. Problem 3 A design for a 105 mm weapon is being considered. The chamber is stated to withstand the desired 35,000 psi and is essentially a steel cylinder of 4.5 in. ID and 7 in. OD (Etube = 30 × 106 psi, ν = 0.3). We have decided to use brass with an OD of 4.490 in. If we use a bilinear, kinematic hardening model where the brass has a modulus of elasticity of 15 × 106 psi, a local tangent modulus of 12.5 × 106 psi, a yield stress of 15,000 psi (yield occurs in this material at ε = 0.001), and an ultimate tensile strength of 45,000 psi, with the information given, what is the radial clearance between the case and the chamber after firing neglecting thermal effects? Answer: Approximately 0.004 in. radial clearance © 2014 by Taylor & Francis Group, LLC 126 Ballistics: Theory and Design of Guns and Ammunition 4.9 Projectile Design While propulsion systems are fairly straightforward in design because their intended use is simple, projectiles vary widely in use and as a consequence their designs are complex and demanding. The propulsion system must get the projectile through the launch environment with consistent muzzle velocities, but without undue stress to the gun or the projectile. The projectile, on the other hand, must withstand the forces of launch, be efficient, consistent and precise in its flight environment, and deliver its intended utility at the target. We will explore only projectile design for launch in this section, reserving design for flight and terminal effects until later. Projectiles may be classified into two general types: cargo carriers and pure kinetic energy deliverers. The cargo carriers include shells that deliver high explosives (HE), submunitions and mines, pyrotechnics, smart munitions, and other specialized lethal systems, e.g., shaped charges (HEAT) and explosively formed penetrators (EFP) shells. The kinetic energy delivery systems, used chiefly for the attack of armor, are monobloc steel shot (AP), saboted, long-rod, heavy metal penetrators (APFSDS), and older types of spinstabilized, saboted (APDS) projectiles. The stresses induced into a projectile during launch are chiefly due to the acceleration that the gases impart to it. The cargo carriers are shells whose stresses are due to relatively low accelerations and which, except for the tank cannon fired HEAT shell, achieve only moderate muzzle velocities. We will therefore explore the kinds of stresses and failures inherent in shell-like structures under load in Section 4.10. Kinetic energy munitions, on the other hand, are subject to extremely high accelerations and have high muzzle velocities. For these types, we will explore the driving mechanism stresses and other aspects of these designs. The gamut of topics in projectile design is almost unlimited. However, several suggest themselves because of their general applicability or timely interest. Shell design is a ubiquitous problem and will be explored in depth in Section 4.10. The use of buttress threads is so common in projectile and gun design that it warrants its own in Section 4.11. Sabot design is more specialized as are the problems of kinetic energy rods and their buttress driving grooves. These will be explored in Section 4.12. Modern projectiles employ a variety of electronic and electromechanical devices for fuzing, target detection, and guidance and control. This relatively new engineering discipline called “gun hardening” deals with designing these devices to survive the harsh environment of gun launch. 4.10 Shell Structural Analysis Most cargo-carrying projectiles, whether fin- or spin-stabilized, are designed with cargo bodies in the shape of an axisymmetric cylindrical shell. Because the loads on these cylinders are the result of spin and acceleration of the shells and their contents, the stresses encountered are highly variable along and through their walls. These stresses will be examined as will the consequences of failure criteria. © 2014 by Taylor & Francis Group, LLC Ammunition Design Practice 127 The symbols and definitions of the constants and variables of shell loading are tabulated next: A—Bore area of the gun a—Linear acceleration d—Diameter of bore (across lands), diameter of assumed shear circle in base of shell di—Inside diameter (ID) of projectile do—Outside diameter (OD) of projectile Fb—Maximum force on base of projectile and rotating band FT—Maximum tangential force on projectile wall FTR—Hoop tension (force) in wall of projectile resulting from rotation of the shell FT′ —Tangential force at section of shell f′—Setback force g—Acceleration due to gravity h—Total depth of filler from nose h′—Total depth of filler from nose end of cavity to section under consideration Izz —Polar moment of inertia I ′zz —Polar moment of inertia of metal parts forward of section when section is ahead of rotating band and aft of it when section is aft of the rotating band n—Twist of the rifling pb—Maximum propellant pressure ph—Filler pressure due to setback prot—Filler equivalent pressure due to rotation, includes wall inertia ri—Inside radius of projectile ro—Outside radius of projectile S—Compressive strength of the rotating band S1—Longitudinal stress S2—Tangential stress S3—Radial stress τ—Shear stress σ Y—Static yield stress in tension T—Torque applied to the projectile t—Base thickness, wall thickness V—Muzzle velocity w—Total projectile weight w′—Weight of metal parts forward of section under consideration w′f —Weight of filler forward of section under consideration α—Angular acceleration ρm—Density of projectile material ρf—Density of filler material ω—Angular velocity r b—Radius of band seat pband—Band pressure We distinguish between thin-walled and thick-walled cylinders in this analysis so that the designer may run quick, ballpark estimates of the stress levels encountered. In practice, FEA is usually conducted on the components, but as emphasized earlier, the designer should have a good idea of the bounds of the answer before beginning the FEA. © 2014 by Taylor & Francis Group, LLC Ballistics: Theory and Design of Guns and Ammunition 128 We begin with a review of basic mechanics of materials as applied to cylinders. If a cylinder is subjected to an axial load and does not buckle, the axial stress can be determined from S1 = − FAxial FAxial =− A π ro2 − ri1 ( ) (4.41) The stress–strain relationships for a cylinder is as follows: ε rr = 1 [σ rr −ν (σ θθ + σ zz )] E (4.42) εθθ = 1 [σ θθ −ν (σ rr + σ zz )] E (4.43) ε zz = 1 [σ zz −ν (σ θθ + σ rr )] E (4.44) where ν is Poisson’s ratio σrr is the radial stress σ θθ is the transverse (hoop) stress σzz is the axial stress E is Young’s modulus If a cylinder is subjected to a torsional load, it will twist. We typically assume that this deformation is small and plane sections remain plane. Thus, when we apply a torque, T, to a cylinder of length, L, with shear modulus, G, and polar moment of inertia, J, the structure will rotate through an angle ϕ (in radians): φ= TL JG (4.45) For a hollow cylinder, J= ( 1 π ro4 − ri4 2 ) (4.46) For a material that behaves according to Hooke’s law, G= E 2(1 + ν ) (4.47) Such a material under pure torsion will only exhibit shear stress according to τθ z = © 2014 by Taylor & Francis Group, LLC Tr J (4.48) Ammunition Design Practice 129 S2 r S1 t FIGURE 4.11 Thin-wall cylinder geometry. While the thick-wall cylinder analysis, which we describe next, is an exact solution, a quick way to assess the major stresses if the wall thickness is less than 10% of the cylinder radius is to assume that the stresses in the radial direction, S3, are negligible. Thus, we examine only the meridional or longitudinal and the circumferential or hoop stresses. We define S1 as the longitudinal stress, S2 as the hoop stress, and p as the pressure depicted in Figure 4.11. If the cylinder has closed ends, then internal pressure can cause a longitudinal stress S1 = σ zz = pr 2t (4.49) otherwise S1 = 0. Internal (or external) pressure always causes hoop stress S2 = σ θθ = pr t (4.50) In practical shell design, we always perform a thick-wall cylinder analysis assuming that the stresses in the radial direction are significant enough to be considered. Thus, we must examine longitudinal, hoop, and radial stresses. We again define S1 = σzz = longitudinal stress, S2 = σ θθ = hoop stress, S3 = σrr = radial stress, and p = pressure. This is depicted in Figure 4.12. The following solutions are known as the Lamé formulas and assume open ends, which implies S1 = 0 if no axial loads are present. If axial loads are present, they must be accounted for. Internal (or external) pressure always causes hoop stress. (Note that the subscripts “o” and “i” refer to the outer and inner surfaces, respectively.) S2 = σ θθ = (  2 ri2ro2 ( po − pi )  1 2 p r − p r − i i o o   r2 ro2 − ri2   ) S3 S2 ro r ri FIGURE 4.12 Thick-wall cylinder geometry. © 2014 by Taylor & Francis Group, LLC S1 (4.51) Ballistics: Theory and Design of Guns and Ammunition 130 with a maximum at r = ri. The radial stress can be calculated from S3 = σ rr = (  2 r 2r 2 ( p − pi )  1 p r − po ro2 + i o o2  2  i i r r − ri   2 o ) (4.52) with a maximum again at the inner surface r = ri, and equal to S3 = −pi. Initially, we will analyze the state of stress caused by the centrifugal loading induced by the rotation of a projectile in a rifled gun tube. In a spin-stabilized projectile, besides the longitudinal loads induced by the acceleration through the tube, the rotation of the projectile, which is dependent upon the axial velocity and the twist of the rifling in the tube, induces stresses in the walls. The twist of the rifling is usually measured in revolutions per caliber of travel (i.e., a twist of 1 in 20 means the projectile makes one revolution in 20 calibers of travel [n = 20]). The units of n are calibers per revolution. If we multiply n by the diameter, d, we get units of length per revolution:  length   caliber   length  n ×d = nd    revolution caliber     revolution   (4.53) Since there are 2π radians per revolution, the angular velocity a projectile has attained is defined as  radians   length  2π V  revolution   time   radians  = = [t −1 ] ω= =   length  time nd    revolu ution  (4.54) The centrifugal force directed radially outward on an element of material at radius r is Fc = mar = w 2 rω g (4.55) In the tangential direction, the inertial forces on an element of material can be determined from Ft = mat = w rα g (4.56) We can determine the centrifugal force on the cylinder wall caused by spinning the cylinder in the absence of other loads by integrating Equation 4.55 from the inner diameter to the outer diameter. To do this, we consider the differential element as depicted in Figure 4.13. From this diagram, we see that the mass of an infinitesimal annular ring of material is dm = © 2014 by Taylor & Francis Group, LLC dw = ρ dV = ρ l 2π rdr g (4.57) Ammunition Design Practice 131 l Angular velocity, ω Density ρ r dr ri ro Projectile model FIGURE 4.13 Differential thickness element geometry. Inserting Equation 4.57 into Equation 4.55 yields dFc = ar dm = ρ l 2π r 2ω 2dr (4.58) which, when integrated from the inner to the outer radius, gives ro FcWALL = 2πρ lω 2 ∫ r 2 dr = ri 2πρ lω 2 3 ro − ri3 3 ( ) (4.59) This is the radial force on the wall due to the inertia of the wall material only. If the projectile is filled with material, we need to account for this filler as well. Thus, if we integrate from the centerline to the inner radius of the projectile wall, we obtain ri ∫ dFcFILL = 2πρFILL lω 2 r 2dr = 0 2πρFILL lω 2 3 ri 3 ( ) (4.60) The total force acting on the projectile wall due to spin is then Fc = FcWALL + FcFILL = 2π lω 2  ρ ro3 − ri3 + ρFILL ri3   3  ( ) ( ) (4.61) For stress computations, we require an internal pressure; thus, we need to convert the centrifugal forces to an equivalent internal pressure. If we assume that our centrifugal forces are acting on the interior of the shell, pushing radially outward, the area for our equivalent pressure is Arad = 2π ril © 2014 by Taylor & Francis Group, LLC (4.62) Ballistics: Theory and Design of Guns and Ammunition 132 Thus, our equivalent pressure can be written as prot = ω2  Fc ρ ro3 − ri3 + ρFILL ri3  =  Arad 3ri  ( ) ( ) (4.63) In Equation 4.56, we determined the tangential force arising from the angular acceleration. If we perform a similar analysis to that which developed Equation 4.61, we will obtain an expression for the torque as follows: T = MWALL + MFILL = ( ) ( ) 1 πα l  ρ ro4 − ri4 + ρFILL ri4    2 (4.64) The derivation of this is left as an exercise for the interested reader and is included as a problem at the end of the chapter. The formulas for calculating the tangential and radial stresses at radial location, r, in a rotating cylinder where ro > 10(ro − ri) can be given as 2 2  3 + ν   2 2 ri ro 1 + 3ν 2  r  σ θθ = ρω 2    ri + ro + 2 − 3 +ν r  8   (4.65) 2 2  3 + ν   2 2 ri ro 2 σ rr = ρω 2    ri + ro + 2 − r  8 r    (4.66) We are reminded that the longitudinal stress (assuming the structure does not buckle) is simply the axial acceleration multiplied by the weight of all of the material forward of the location of interest divided by the shell cross-sectional area—we will discuss this presently. These formulas were developed for the centrifugal loading of a spinning projectile by forces that act during both in-gun setback and flight. The axial load on a projectile, however, is for the most part only present during acceleration in the tube, is a function of time, and occurs whether the projectile is spinning or not. Beyond this, there is also an applied torque due to the angular acceleration, which is applied through the rotating band or slip obturator. The setback load and (if spinning) the centrifugal and torsional loads must all be superimposed on the projectile to determine its state of stress. The axial force on the projectile during firing is given by F = ps A (4.67) where ps is the pressure acting on the base of the projectile defined by the Lagrange approximation   1 ps = pB  c  1+ 2w  © 2014 by Taylor & Francis Group, LLC      (4.68) Ammunition Design Practice 133 The D’Alembert force is the force due to acceleration that exactly equals this pressure force a= ps Ag w (4.69) At any axial position, the force on the cross-sectional area can be shown to be proportional to the weight of material forward of the section: f′ = w′ ps A w (4.70) To calculate the force (or really the pressure) in the filler material, we usually resort to a hydrostatic model ph = ρ ha = ρ h ps Ag w (4.71) where ρ is the density of the filler h is the filler head height ph is the hydrostatic pressure that is developed In a spin-stabilized projectile, the angular acceleration, α, is proportional to the linear acceleration, a, where α = Ka (4.72) Then, α =K ps Ag w (4.73) where K has units of length−1 and is dependent upon the twist, n (in calibers of travel per turn), and the bore diameter, d, thus K = tan θ = 2π nd (4.74) where θ is the angle between the circumferential twist distance and the axial distance traveled. From Equations 4.73 and 4.74, we get α= 2π ps Ag nd w (4.75) If we define a tangential force applied to the rotating band of the projectile as FT, then the torque on the projectile is T = FT © 2014 by Taylor & Francis Group, LLC d 2 (4.76) 134 Ballistics: Theory and Design of Guns and Ammunition We know that the torque is equal to the product of the polar moment of inertia of the projectile and its angular acceleration T = I zzα (4.77) Solving for the angular acceleration in terms of the tangential force, we get α= FT d I zz 2 (4.78) Inserting this into Equation 4.56 and solving for FT yields FT = π 2I zz ps nw (4.79) Since A =π d2 4 (4.80) The force that is applied by the rifling to the rotating band is transmitted through the structure to regions both forward and aft of the rotating band. These forces are proportional to the moment of inertia of the sections ahead of or behind the application of the torque load, I ′zz. We assume that this force acts over a mean diameter of the outer and inner wall surfaces of the shell and then we get FT′ = 16π I ′zz ps A n(do + di )2 w (4.81) Because the rotating band is intended to act as a gas seal (obturator) as well as the rotational driver, designs typically exhibit a diameter over the band that is slightly larger than the groove diameter of the weapon. The engraving action of the gun lands and the interference fit in the grooves causes a plastic flow of the band, resulting in a pressure on the band seat as well as a developed reaction in the gun wall. This pressure can be greater than the gas base pressure on the projectile. Measurements of this pressure have been obtained by strain gaging of the gun tube and computing the stress at the weapon’s inner diameter. The pressure required to cause this stress is called the interface pressure. It has been shown that cannelures or circumferential grooves cut into the band surface reduce this pressure substantially by allowing room for band material to flow rather than being loaded in a quasi-hydrostatic condition. This is depicted in Figure 4.14. The composition/material of the rotating band can have a dramatic effect upon the behavior of the projectile in the tube as well as tube wear. An excellent example of this relationship is contained in Ref. [4]. We have the forces on the projectile structure but now must translate these into stresses that allow us to determine how much design margin is present. Once determined, these stresses are then linked to well-established failure criteria to determine the failure point of the material. Since projectiles may be made of a variety of materials, © 2014 by Taylor & Francis Group, LLC Ammunition Design Practice 135 pband ps Cannelures FIGURE 4.14 Rotating band pressure. r z Point A j=1 j=4 j=2 j=3 FIGURE 4.15 Stress locations in an M1 high explosive (HE) projectile. specialized criteria may have to be used on each material. This full procedure is somewhat complicated and beyond the scope of this book, but we will attempt to describe the basics through an examination of a simple M1, high-explosive projectile structure depicted in Figure 4.15. Assume a thick-walled cylinder as shown for stress calculations where S1j —Longitudinal stress at the jth location S2j —Hoop stress at the jth location S3j —Radial stress at the jth location τ 11—Longitudinal shear at the base τ 2j —Torsional (shear) stress at the jth location It is helpful to recap here all of the loads on an element of projectile wall material at a generalized location (such as point A) in the diagram. This element of material is • Compressed in the axial direction due to the axial acceleration • Loaded in tension in the hoop direction because of the wall mass being pulled radially outward due to the spin • Loaded in tension in the hoop direction because of the filler material moving outward due to the setback load and the spin • Loaded in shear due to the rotating band accelerating the projectile in an angular direction • Loaded in shear due to the greater stress in the outer wall than on the inside wall © 2014 by Taylor & Francis Group, LLC Ballistics: Theory and Design of Guns and Ammunition 136 pband ps FIGURE 4.16 Load conditions for an M1 HE projectile. Note that when including mass forward of a particular section we must include all mass transmitting loads to the section, e.g., fuze, bushings, and cups. The pressures applied to our model of the M1 projectile are shown in Figure 4.16. Now let us examine specific locations of interest along the shell where experience tells us failures might occur. For convenience, these have been tabulated in Table 4.1 and tailored to each individual location with the symbol, source load, and type of stress noted. At location 1, these are the formulas used to calculate stresses due to the setback of filler on base, the moments caused thereby, and by gas pressure on base: S11 = − ph (4.82) S21 = 0 (4.83) S31 = ro2 ( ps − ph ) t2  3r 3 o S′31 = − ps   2 ro3 − ri3  ( ) (4.84)   r 3 + 2r 3 i  + ph  o   2 ro3 − ri3   ( )     (4.85) TABLE 4.1 Typical Stresses in a High Explosive (HE) Projectile and Their Sources Type of Stress Symbol Compressive load on base Radial stress on base at centerline S11 S31 Radial stress on base at centerline ′ S31 Hoop stress at rear of the band S22 Radial stress at ends of band and maximum ID Longitudinal stress at ends of band and maximum ID Hoop stress at forward end of band and maximum ID Shear stress through thickness t S32, S33, S34 τ11 Torsional shear in projectile wall τ22, τ23, τ24 © 2014 by Taylor & Francis Group, LLC S12, S13, S14 S23, S24 Source of Load Setback of filler Moments of filler setback and base pressures (flat base) Moments of filler setback and base pressures (round base) Setback of filler, rotation, and external pressure (band and gas) Rotation of projectile, filler setback, and filler rotation Setback of metal parts in wall (filler contribution usually neglected) Filler pressure and rotation of wall Moments of filler setback and base pressures (round base) Setback of filler, rotation, and external pressure (band and gas) Ammunition Design Practice 137 Equation 4.82 is the axial component stress. We can see that it is just driven by the reaction of the fill and shell to the axial acceleration. Since this is a centerline location, by definition there is no hoop stress, which is defined by Equation 4.83. Equation 4.84 specifies the radial stress assuming the base is flat faced. This comes about from the difference in the base pressure reacting against the internal forces and attempting to push the center of the base into the fill. Equation 4.85 is the radial stress equation assuming the base is a rounded bottom (i.e., with the concave portion enclosing the fill). We can see from this equation that the stresses are much lower as it carries the load more efficiently than a flat bottom shell. The drawback is that a base of this type requires a skirted boat tail which is more expensive to manufacture but saves considerable weight. Moving to location 2, these are the stresses due to setback of filler, filler rotation, wall rotation, and band pressure: S12 = − w′ + w′f  ps A  w  π ro2 − ri2  ( )   ( ph + prot ) ro2  +    ro2 − ri2     ( ) (4.86)  r2 + r2   r2  S22 = ( ph + prot )  o2 i2  − pband  2 b 2   ro − ri   rb − ri  (4.87) S32 = −( ph + prot ) (4.88) At this location, we see that the axial stress defined by Equation 4.86 has two parts. The first term on the RHS is the inertia of all the fill and shell material ahead of this location. The second term is the axial stress caused by the internal pressure of the fill expanding. In Equation 4.87, the first term on the RHS is the contribution of spin to the hoop stress and the second term is the restoring force caused by the gun tube pushing in on the rotating band. Equation 4.88 is simply the radial stress caused by the rotation and compression of the fill and wall. Further forward on the shell at location 3, the stresses due to setback of filler, filler rotation, wall rotation, and band pressure have identical formulas to location 2 but with, of course, different values of the variables due to the lower hydrostatic pressure component: S13 = − w′ + w′f  ps A  w  π ro2 − ri2  ( )   ( ph + prot ) ro2  +    ro2 − ri2     ( ) (4.89)  r2 + r2   r2  S23 = ( ph + prot )  o2 i2  − pband  2 b 2   ro − ri   rb − ri  (4.90) S33 = −( ph + prot ) (4.91) Finally at location 4, near the nose of the shell, the stresses due to setback of filler, filler rotation, and wall rotation are as follows: S14 = − © 2014 by Taylor & Francis Group, LLC w′ + w′f  ps A  w  π ro2 − ri2  ( )   ( ph + prot ) ro2  +    ro2 − ri2     ( ) (4.92) Ballistics: Theory and Design of Guns and Ammunition 138  r2 + r2  S24 = ( ph + prot )  o2 i2   ro − ri  (4.93) S34 = −( ph + prot ) (4.94) At each location, one must be certain to use the proper head height of filler and the proper inner and outer radii of the shell. We must also account for the shear stresses which are most severe at location 1. For simplicity, we will assume a flat base and calculate the shear stress due to wall torsion. Wherever these calculations are done on the shell, the proper Izz and the proper inner and outer diameters must be used: τ 11 = τ 22 , τ 23 , τ 24 = ( ps − ph )πri2 ( ps − ph )ri = 2πrt 2t i (4.95) ps A 64I ′zz FT′ = 3 π 2 ( do − di2 ) n(do + di ) (do − di ) w 4 (4.96) A typical loading of the shell using known weights, pressures, and acceleration is shown in Figure 4.17 and Table 4.2. Location of interest— on inside wall 6 in. Y CS00 Z X Section properties at or ahead of 6-in. location: m6 = 17.15 lbm Mass do 6 = 4.10 in. OD di6 = 2.95 in. ID h6 = 8.189 in. Head height Izz6 = 41.15 lbm-in.2 Moment of inertia Polar MOI J6 = 108.3 in.4 FIGURE 4.17 Location of interest on a 105 mm M1 projectile. TABLE 4.2 Typical Values for Use in an HE Projectile Design Component Weight (lbm) Fuze Body Rotating band Filler (TNT) Total © 2014 by Taylor & Francis Group, LLC 2.1 34.0 0.4 5.5 42.0 Loads Breech pressure (psi) Spin rate—maximum p (Hz) Base pressure (psi) Acceleration (g’s) Angular acceleration (rad/s2) 38,400 82.4 37,150 11,873 348,600 Ammunition Design Practice 139 The common practice currently used in projectile design is to dispense with the hand calculations and go right to a FEA. While this is usually very accurate and saves a good deal of time, there are instances when one would like to check the answers through a hand calculation. Let us examine one location on this 105 mm M1 HE projectile fired from an M2A2 cannon at 145°F. Projectile data: Shell material: HF-1 steel • • • • • • Density—0.283 lbm/in.3 Projectile OD—4.10 in. Projectile ID (average)—2.95 in. Projectile effective (including friction) mass (fuzed)—42 lb Projectile base intrusion into cartridge case—85.43 in.3 Izz —80.24 lbm-in.2 Fill material: TNT • • • • • Density—0.036 lbm/in.3 Total length of explosive column—13.44 in. Izz —5.17 lbm-in.2 Average fill cross-sectional area—6.49 in.2 Fill surface area—124 in.2 The M1 projectile fired from our cannon is depicted in Figure 4.17. The properties of the section ahead of the location of interest are provided in Figure 4.17. We shall determine the stress tensor at the location shown. We shall assume the projectile obturates perfectly and that there is no friction between the projectile and the tube. To begin, we should always draw a free-body diagram of an infinitesimal element at the point of interest. Let us look at the hoop direction first. We shall use Equation 4.51: σ θθ = (  2 r 2r 2 ( p − pi )  1 p r − po ro2 − i o o2  2  i i r r − ri   2 o ) (4.97) σθθ σθθ In this case, r = ri and po = 0 so we can write σ θθ = © 2014 by Taylor & Francis Group, LLC ( ( ) 1  pi ri2 + ro2   ro2 − ri2  ) (4.98) Ballistics: Theory and Design of Guns and Ammunition 140 The internal pressure is found through our equivalent pressure technique mentioned earlier: prot = ωp2max  ρ ro3 − ri3 + ρfill ri3   3ri  ( (4.99) 2 2 prot ) rev   rad  (82.4)2  (2π )2    s   rev  = in.   lbm-ft  (3)(1.475) [in.](12)   (32.2)  2  ft   lbf-s     lbm   lbm  × (0.283)  3  (2.05)3 − (1.475)3  [in.3 ] + (0.036)  3  (1.475)3 [in.3 ]  in.   in.     lbf  prot = 258  2   in.  For the hydrostatic component of the equivalent pressure, we know that ph = ρfill ap max h6 (4.100)  lbm   ft  (0.036)  3  (382,300)  2  (8.189) [in.] in.   s  ph =  lbm-ft  (32.2)  2  lbf-s   lbf  ph = 3500  2   in.  The equivalent internal pressure is then peq = prot + ph  lbf  peq = pi = 3758  2   in.  The hoop stress is then σ θθ = 2 2    lbf   4.10   2.95   2  ( ) + 3758    [in. ]   2   2 2   4.10   2.95    in.   2   2    2      [in. ]  − 2 2      1  lbf  σ θθ = 11, 830  2   in.  © 2014 by Taylor & Francis Group, LLC (4.101) Ammunition Design Practice 141 Now let us look at the axial stress. This is the stress at the point due to two things: the axial inertia of all the material ahead of the cut setting back and the effective internal pressure caused by the rotation of the projectile and the hydrostatic compression of the fill material: σ zz = (p r (r 2 i i 2 o − po ro2 −r 2 i ) )− FAxial π ro2 − ri2 ( ) (4.102) σzz σzz We shall use the radii given in the problem statement. We put negative sign in the aforementioned equation to denote compressive stress because only the axial component loads the inner wall in compression. The force acting on the section of interest due to setback is given by FAxial = m6 apmax FAxial  ft  (382,300)  2  (17.15) [lbm] s  = 203,600 [lbf ] =  lbm-ft  (32.2)  2  lbf-s  Using this result, we have 2 σ zz  lbf   2.95  2 (3758)  2    [in. ] (203,600) [lbf ] in. 2    − = 2 2 π  4.10   2.95   ( 4.10)2 − (2.95)2  [in.2 ] 2     [in. ] 4   −  2   2    lbf  σ zz = −27 ,940  2   in.  © 2014 by Taylor & Francis Group, LLC (4.103) Ballistics: Theory and Design of Guns and Ammunition 142 Many times we neglect the first term in the aforementioned equation for conservatism. In the radial direction, we only have our equivalent pressure pushing radially outward and our location of interest is on the ID, so σ rr = − peq (4.104)  lbf  σ rr = −3758  2   in.  σrr σrr The angular acceleration will generate a torque through the rotating band that results in a shear stress in the plane normal to the axis of the projectile. τzθ τzθ The torque on the projectile is also the opposite of the torque on the gun tube and comes directly from Equation 4.77: T6 = I zz6 α pmax (4.105) The moments of inertia were provided and we must use the angular acceleration calculated at peak pressure provided earlier. Now the torque comes about through 2  rad   1   lbf-s   1   ft  T6 = ( 41.15) [lbm-in.2 ](348,600)  2       s   32.2   lbm-ft   12   in.  T6 = 37 ,130 [lbf-in.] The in-plane shear stress is given by τ= Tr J Then we have τ zθ  2.95  (37 ,130) [lbf-in.]   [in.]  lbf   2  = = 506  2  4 (108.3) [in. ]  in.  © 2014 by Taylor & Francis Group, LLC (4.106) Ammunition Design Practice 143 The shear stress caused by the rotation is generated by the shell trying to spin up the explosive fill. The torque on the explosive fill is determined through Tfill = I zzfill α pmax (4.107) 2  rad   1   lbf-s   1   ft  Tfill = (5.17 ) [lbm-in.2 ] (348,600)  2       s   32.2   lbm-ft   12   in.  Tfill = 4644 [lbf-in.] This generates a force at the internal radius of Ffill = Ffill = Tfill ri (4.108) ( 4644) [lbf-in.] = 3162 [lbf ]  2.95    [in.]  2  Smearing this over the entire internal surface area gives us τ rθ = (3162) [lbf ]  lbf  = 25.5  2  2 (124) [in. ]  in.  The axial shear is approximated as a worst case by calculating the hydrostatic pressure at the bottom of the explosive column, transforming it into a force, and smearing that force over the entire internal cavity area. We know the entire explosive column height is h = 13.44 [in.] Then the peak hydrostatic pressure of the fill is ph = ρfill apmax h (4.109)  lbm   ft  (0.036)  3  (382,300)  2  (13.44) [in.] in.   s  ph =  lbm-ft  (32.2)  2  lbf-s   lbf  ph = 5744  2   in.  Calculating this pressure over the average cross-sectional area of the projectile, we obtain Fbase = ph Aavgfill  lbf  FAxial = (5744)  2   in.  (6.49) [in.2 ] = 37 ,280 [lbf ] © 2014 by Taylor & Francis Group, LLC (4.110) Ballistics: Theory and Design of Guns and Ammunition 144 Now this force smeared over the interior surface area will yield the stress τ rz = (37 ,280) [lbf ] − Faxial  lbf  =− = −300  2  Afill (124) [in.2 ]  in.  (4.111) We can now write our stress tensor σ rr σ = τ rθ τ rz τ rθ σ θθ τθ z τ rz   −3,758 τ θ z  =  25.5 σ zz   −300 25.5 11, 830 506 −300    lbf  506   2   in.  −27 ,940  It must be noted that these equations assumed that there were no other forces acting on the projectile. For instance, in some projectiles with poorly designed rotating bands, leaking of the propellant gases (known as blow-by) causes the exterior of the projectile to be pressurized. This load must be considered because it has been known to collapse projectiles in development. Another point is that, while it is common to check a projectile at peak acceleration, the spin rate at this location is not a maximum. Maximum spin occurs at the exit of the muzzle of the weapon where the velocity is the highest. It is always good practice to check a projectile for maximum spin with no axial acceleration to simulate this. Problem 4 A high-explosive projectile is to be designed for a 155 mm cannon using a 12 in. thick steel wall with TNT as the filler material. Assume the shell and filler are a cylinder 0.75 m in length. It is to be capable of surviving a worn-tube torsional impulse (angular acceleration) of 440,000 rad/s2. 1. Derive the expression to calculate the torque on the projectile that achieves this acceleration if the torque is applied at the OD of the shell. 2. Calculate the value of the torque assuming the density of steel is 0.283 lbm/in.3 and TNT is 0.060 lbm/in.3 Hint: Start from FT = maT Answer: ( ) 1. T = MWALL + MFILL = 12 πα l  ρ ro4 − ri4 + ρFILL ri4  , 2. T = 796,600 [lbf-in.]   Problem 5 To participate in a failure investigation of an explosive, someone asks you to look at their design of a cylinder that was supposed to hold the explosive during a 155 mm Howitzer launch. Assume the explosive sticks completely to the interior wall. The firing conditions at the time of the failure were as follows: © 2014 by Taylor & Francis Group, LLC Ammunition Design Practice 145 6.092 in. 1 in. 1 in. 10 in. 4 in. Axial acceleration = 10,000 g Angular acceleration = 300,000 rad/s2 Angular velocity = 100 Hz The projectile was as shown above: The wall is AISI 4140  lbf  E = 30 × 106  2   in.  ν = 0.29  lbm  ρ = 0.283  3   in.  The explosive is composition B  g  ρfill = 0.71  3   cm  Write the stress tensor for a point on the inside diameter, 4 in. from the base Answer: σ rr  σ = τ rθ τ rz τ rθ σ θθ τθ z © 2014 by Taylor & Francis Group, LLC τ rz   −2265 τ θ z  =  21 σ zz   −266 21 5564 −2864 −266    lbf  −2864   2   in.  −19,307  Ballistics: Theory and Design of Guns and Ammunition 146 Problem 6 A 155 mm projectile is fired from a tube with a 1 in 20 twist. Its muzzle velocity is 1000 m/s. What is the spin rate at the muzzle in Hz? Answer: 322.6 [Hz] Problem 7 It is requested that a brass slip ring be constructed for a spin test fixture to allow electrical signals to be passed (although real noisy) to some instrumentation. The design requirements are for the ring to have an ID of 4 in., a length of 2 in., and be capable of supporting itself during a 150 Hz spin test. How thick does the ring have to be? The properties of brass are as follows: yield strength of 15,000 psi and density of approximately 0.32 lbm/in.3 Answer: 1/4 in. thickness will work but it can be thinner Problem 8 A 155 mm projectile is to be designed with a rotating band designed to discard as the projectile leaves the muzzle of the weapon. The rotating band is fixed to the projectile (so it transmits the proper torque to the projectile) with splines that prevent rotational motion relative to the projectile while allowing the band to expand in the radial direction for proper discard. This function must occur at the highest as well as the lowest spin rates. The two extreme muzzle velocities are 250 and 800 m/s, respectively, with corresponding peak axial accelerations of 2000 and 15,500 g, respectively. The projectile mass is 98 lbm and the axial moment of inertia is 41 lbm in2. Geometry constraints require the band to have an engraved outside diameter of 6.2 in. and the outer diameter of the band seat (band inside diameter) is 5.5 in. The band is 2.5 in. long. The 48 rifling lands are 0.05 in. high and are 0.2 in. wide. Consider both a copper and a soft iron band with the properties provided next and determine whether the bands will (a) Withstand the shear at peak angular acceleration. (b) Break up upon muzzle exit. (c) If one (or both) designs fail to work properly, what can be done with the analysis and/or design to make it work? Is there anything we must be careful of? Assume the following: The weapon has a 1 in 20 twist. At peak acceleration, only shear on the rotating band need be considered. Ignore the increase in shear area caused by the rifling helix. Ignore the stress concentration developed by the engraving for discard calculations. Ignore the effect of the splines. For conservatism, on muzzle exit ignore the mass of material above the rifling marks (i.e., use an OD of 6.1 in. for the band). The properties for the copper and steel are as follows: Ultimate tensile strength Shear strength Density Poisson’s ratio Copper Soft Iron 35,000 psi 19,250 psi 0.316 lbm/in.3 0.28 40,000 psi 22,000 psi 0.263 lbm/in.3 0.34 Problem 9 An experimental 40 mm gun has an average chamber inner diameter 60 mm. The weapon is expected to develop a maximum breech pressure of 35,000 psi. If we would like the © 2014 by Taylor & Francis Group, LLC Ammunition Design Practice 147 weapon to withstand 10,000 cycles at this pressure and given the properties of the steel given later, determine the outside diameter of the chamber. Without proper design experience an interference fit can sometimes be catastrophic. If we were instead to design this chamber out of two tubes, each at half of this thickness but with the outer tube compressing the inner tube by 0.002 in. diametrally, what is the maximum pressure the design will accommodate and still function for the 10,000 cycles? Assume AISI 4340 steel with a yield strength (SY) of 100,000 psi. The endurance stress (S’n) for 4340 is 0.875SY for the amount of cycles desired. Assume the following factors from our cyclic loading discussion: CR = 0.93, CG = 0.95 and CS = 0.99. Assume the chamber is open ended. The modulus of elasticity and Poisson’s ratio are 30 × 106 psi and 0.3, respectively. Problem 10 You are to design a fragment throwing gun system for another organization to be used in fragment impact testing. The gun is to throw a 22 g, 0.500 in. diameter, cylindrical fragment at 2300 m/s. Your design must use a brass cartridge case to assist with obturation of the breech. Other assumptions and information are 1. You do not need to design the breech—assume it will hold the cartridge case in properly (in reality we can always add more threads to the design). 2. Even though there will be a slight taper on the chamber (which must be larger than the bore diameter for seating purposes), assume, for calculation purposes, that the chamber is cylindrical at its maximum diameter. 3. The tube is to be steel and assume that the yield strength is 60,000 psi (this accounts for the effect of cyclic loading). The modulus of elasticity is 29 × 106 psi. Poisson’s ratio is 0.29. 4. Assume the propellant is either cylindrical or single perforated (and state your assumption). 5. Choose from the following propellants: Propellant Linearized Burn Rate, β (in./s/psi) Solid Density, δ (lbm/in.3) Adiabatic Flame Temperature, T0 (°R) Propellant Force, λ (ft-lbf/lbm) Specific Heat Ratio, γ IMR M12 Bullseye Red dot Navy pyro 0.000132 0.000137 0.000316 0.000153 0.000135 0.0602 0.0600 0.0590 0.0593 0.0566 5103 5393 6804 5774 4477 327,000 362,000 425,000 375,000 321,000 1.2413 1.2326 1.2523 1.2400 1.2454 6. Assume the cartridge case is brass and use a bilinear kinematic hardening model where the brass has a modulus of elasticity of 15 × 106 psi, a local tangent modulus of 13 × 106 psi, and a yield stress of 16,000 psi (yield occurs in this material at ε = 0.002). 7. Weight is not a major concern; however, you should make the design light enough to be moved using reasonable test range equipment. © 2014 by Taylor & Francis Group, LLC 148 Ballistics: Theory and Design of Guns and Ammunition The design is to proceed as follows (not necessarily in the order given): A. Interior ballistics design a. Size the chamber length and diameter. b. Determine the amount of propellant needed based on your choice of the aforementioned propellants and propellant geometry (make sure it fits in the chamber). c. Determine a web thickness for the propellant. d. Determine the length of the gun. e. Determine V, pB and x for the projectile at peak pressure. f. Determine Vc, pBc and xc for the projectile at charge burnout. g. Determine the muzzle velocity of the projectile. B. Gun tube design a. Based on the calculations of part (A), develop a pressure–distance curve to use as criteria for your gun design b. Determine the outside diameter of the gun tube. To keep the design light as possible, use the design rules provided in the text and taper the tube toward the muzzle. If needed, over the chamber, you may shrink fit cylinders to build up a composite tube. c. Determine the weight of your gun and comment on if it is reasonable. C. Cartridge case design a. Determine a thickness and tolerance for your cartridge case. b. Determine the outside diameter and tolerance for the cartridge case. c. Decide on a tolerance for your chamber inside diameter. Note that for these calculations show that the case may be easily extracted at the limits of the tolerance. It is important that you write down all of your assumptions. It is also highly likely that as you proceed further along with your design you may come upon a situation that requires you to revisit an assumption you made earlier—this is to be expected and it is part of the design process. 4.11 Buttress Thread Design There are a variety of instances where a buttress thread form is the desired means of transmitting loads between mating components. In some instances, the thread form is not the usual continuous spiral associated with a normal thread but a series of discontinuous grooves that exhibit the cross-sectional form of the buttress. In this section, we will discuss a true thread with lead-ins and partial thread shapes, but we will assume that the basic analysis will apply to buttress grooves as well. Buttress threads are designed to maximize the load-carrying capability in one direction of a threaded joint. There are many variations on such threads, but on ammunition © 2014 by Taylor & Francis Group, LLC Ammunition Design Practice 149 Pitch of the thread Pressure flank 7° 45° Load carrying or shearing FIGURE 4.18 Depiction of a standard buttress thread. components we predominantly use threads with a pressure flank angle (described later) of 7° as shown in Figure 4.18. Thread callouts on drawings usually appear, e.g., 2.750-4UNC-2A LH Buttress The meanings of these callouts are as follows: • • • • First number is the major diameter of the thread (here it is in inches) Second number is how many threads per inch The letters are the thread form callout (UNC = Unified National Coarse) The last number is the class of fit of the thread related to clearances in the engagement (3 is the tightest fit, 1 the loosest) • The last letter determines whether the thread is male (A) or female (B) (mnemonic − A = Adam = male) • LH means left handed (there will be no callout if the threads are right-hand twist or if the thread is a groove and not a continuous spiral) Thread nomenclature of relevance is as follows: • Major diameter is the largest diameter of the thread form • Minor diameter is the smallest diameter of the thread form • Pitch diameter is the diameter where there is 1/2 metal and 1/2 air We use buttress threads for several reasons: most important is to improve the directional loading characteristics of the thread; also to allow for a more repeatable, controllable shear during an expulsion event, i.e., if we want the thread to intentionally and controllably fail allowing separation of the components; and to prevent thread slip in joints with fine threads or threads on thin shell walls. If thread slip occurs, the threads can either dilate or contract elastically and the joint can pop apart with little or no apparent damage to the threads. When we design for strength, we typically calculate the strength based on the shear area at the pitch diameter in the weaker material. This, of course, translates to half the length of engagement of the threads. This is acceptable because we usually use conservative properties and add a safety factor to account for material variations and tolerances. © 2014 by Taylor & Francis Group, LLC 150 Ballistics: Theory and Design of Guns and Ammunition Element at point A r Compressive stress from F F Neutral axis di σ= Mc I Tensile stress Mc from σ = I FIGURE 4.19 Depiction of a standard buttress thread. We must always base our calculations on the weaker material if the design is to be robust. When designing to actually fail the threads, however, we need to be more exact in our analysis and take everything such as actual material property variation and tolerancing into account or our answers will be wrong. We will proceed in this analysis in meticulous detail, initially, as a cantilevered beam subjected to compressive and tensile stresses caused by contact forces and bending moments. This technique was first developed during the U.S. Army’s sense and destroy armor (SADARM) program by Dan Pangburn of Aerojet Corporation [5] and has been used by the U.S. Army. We consider the thread form as a short, tapered, cantilever beam and assume that failure will occur as a result of a combination of stresses and that combined bending and compressive stress precipitate the failure. This is depicted in Figure 4.19. If we examine this figure, we see that the distributed force, F, causes our beam to bend in the classical sense with the loaded flank in tension and the unloaded flank in compression about the neutral axis. We have separated an element of material out from point A in the Figure. The freebody diagram of this element shows that the bending of the beam puts it in tension, while the loading on the pressure flank puts it in compression. It is this combined load that will cause failure of the material. If we were analyzing this in a finite element code, the bending and compression would cause combined stresses and the part would fail by one of the failure criteria that were discussed earlier. However, in this case, we will use the maximum shear criteria to check for failure at some radius in the thread and will also check the load at which failure occurs with the von Mises criteria at the thread roots, di on the male thread and do on the female thread. These are the diameters of the loading (i.e., the mating thread contact areas) as depicted in Figure 4.20. For simplicity, we shall call the male thread the “bolt” (subscript 1) and the female thread the “nut” (subscript 2). The loading is further described by Figure 4.21. In this figure, the radius, r, is the plane at which the threads will shear. If we assume the contact is frictionless, the average normal stress is simply the total axial force, F, divided by the projected area, A. We have assumed that the normal stress is constant over the contact area. This gives us a negative value because the stress is compressive. Figure 4.22 shows the configuration where the normal force has been termed F4 and the thread area is A4. Since an axial loading is what shears the threads, we need to project the © 2014 by Taylor & Francis Group, LLC Ammunition Design Practice 151 do Bolt (1) r F di The location where these stresses are the greatest is here along the contact surface FIGURE 4.20 Definition of load radii. Bolt (1) r = Shear radius do = Outer diameter di = Inner diameter r do F2 t2 F1 Nut (2) di t1 1 2 FIGURE 4.21 Loading diagram of buttress threads. F F4 A4 1 A FIGURE 4.22 Loading of a thread surface. © 2014 by Taylor & Francis Group, LLC Ballistics: Theory and Design of Guns and Ammunition 152 components of this force along the axis of the projectile (i.e., rotate through the angle, ϕ1). This allows us to express the stress as −F σN = 4 = A4 F F cos φ1 = − = σv A A cos φ1 − (4.112) where σ N and σv are the normal and axial stresses, respectively. By substituting the area, A, we get σv = − F π 2 2 do − di 4 ( ) (4.113) If we assume that failure takes place at a radius, r, yet to be determined, the bearing force on the external thread (bolt) that produces bending in the thread is  d  2  F1 = −πσ v  o  − r 2   2   (4.114) Similarly, the force that produces bending in the internal thread (nut) is 2  d   F2 = −πσ v  r 2 −  i    2    (4.115) Now the pitch diameter is defined as the location where the thickness of the thread is onehalf the thread pitch. Since thread failure occurs at an assumed radius, r, we need to define the thicknesses of both the male and female threads at this location. First, recall that the thread pitch is p and then define dpf as the internal (female) thread pitch diameter and dpm as the external (male) thread pitch diameter. Then t1 and t2 from our earlier diagram can be expressed as follows: t1 = dpm p  −r − 2  2   (tan φ1 + tan φ2 )  (4.116) t2 = p  dpf  − − r  (tan φ1 + tan φ2 ) 2  2  (4.117) Then the bending stress can be calculated from simple beam theory as t M 3M Mc 2 σ= = = 1 3 I rt 2 π (2π r )t 12 © 2014 by Taylor & Francis Group, LLC (4.118) Ammunition Design Practice 153 where c is the distance from the point of interest, r, to the neutral (bending) axis and I is the area moment of inertia of the cross section. The bending stress in the external (male) thread is then d  3 F1  o − r   2  σ1 = 2π rt12 (4.119) Similarly, we can show that the bending stress in the internal (female) thread is d   3 F2  r − i  2  σ2 = 2π rt22 (4.120) In considering the failure criteria, we shall assume that the maximum shear stress in the material must not exceed 0.6 times the material strength in a tensile test. We will use the yield strength as this material strength because at that point in failure the geometry of the part is changing. Experience has shown that once this begins to happen the part is in the process of failing anyway and will not recover. In a state of combined loading, the maximum shear stress can be found from 1 τ max = |σ max + σ min | 2 (4.121) This averaging can be shown to be τ max = σ −σN = 0.6Y 2 (4.122) Here we are reminded that σ N and σv are compressive therefore negative numbers and Y is the yield stress in tension. The equivalent stress at failure in the male thread is then Y1 = σ1 −σ v 1.2 (4.123) Y2 = σ2 −σv 1.2 (4.124) and in the female thread it is In these equations, Y1 and Y2 are the yield stress in the male and female threads, respectively. We will now combine Equations 4.123 and 4.119 as well as Equations 4.124 and 4.120 to eliminate σ 1 and σ 2, respectively. This yields 1 do − r σ Y1 = 1.25F1 2 2 − v π rt1 1.2 © 2014 by Taylor & Francis Group, LLC (4.125) Ballistics: Theory and Design of Guns and Ammunition 154 and Y2 = 1.25F2 1 di 2 − σv 2 1.2 π rt2 r− (4.126) We now combine Equations 4.125 and 4.116 as well as Equations 4.126 and 4.117 to eliminate the thicknesses, t1 and t2, respectively. This yields 1 do − r σv 2 Y1 = 1.25F1 2 − 1 .2 1  1   π r  p −  r − dpm  (tan φ1 + tan φ2 ) 2   2  (4.127) and 1 di σv 2 Y2 = 1.25F2 2 − .2 1 1  1  π r  p −  dpf − r  (tan φ1 + tan φ2 ) 2  2  r− (4.128) We will now insert Equation 4.114 into Equation 4.127 and Equation 4.115 into Equation 4.128 to eliminate F1 and F2, respectively. This yields ( Y1 = −0.3125σ v d − 4r 2 o 2 ) 1 do − r σv 2 2 − 1.2 1  1   r  p −  r − dpm  (tan φ1 + tan φ2 ) 2 2     (4.129) and 1 di σv 2 Y2 = −0.3125σ v ( 4r − d ) 2 − 1.2 1  1  r  p −  dpf − r  (tan φ1 + tan φ2 ) 2  2  2 r− 2 i (4.130) Now we must solve Equations 4.129 and 4.130 in terms of σv. The first of these is σv = −Y1 G3 + G2 + G1 + G0 + 1 1.2 (4.131) where G3 = 0.15625do3 r(0.5 p − r tan φ1 − r tan φ2 + 0.5dpm tan φ1 + 0.5dpm tan φ2 )2 © 2014 by Taylor & Francis Group, LLC (4.132) Ammunition Design Practice 155 G2 = −0.3125do2 (0.5 p − r tan φ1 − r tan φ2 + 0.5dpm tan φ1 + 0.5dpm tan φ2 )2 (4.133) G1 = −0.625rdo (0.5 p − r tan φ1 − r tan φ2 + 0.5dpm tan φ1 + 0.5dpm tan φ2 )2 (4.134) G0 = 1.25r 2 (0.5 p − r tan φ1 − r tan φ2 + 0.5dpm tan φ1 + 0.5dpm tan φ2 )2 (4.135) The second equation is σv = −Y2 1 H 3 + H 2 + H1 + H 0 + 1.2 (4.136) where H3 = 0.15625di3 r(0.5 p + r tan φ1 + r tan φ2 − 0.5dpf tan φ1 − 0.5dpf tan φ2 )2 (4.137) H2 = −0.3125di2 (0.5 p + r tan φ1 + r tan φ2 − 0.5dpf tan φ1 − 0.5dpf tan φ2 )2 (4.138) H1 = −0.625rdi (0.5 p + r tan φ1 + r tan φ2 − 0.5dpf tan φ1 − 0.5dpf tan φ2 )2 (4.139) H0 = 1.25r 2 (0.5 p + r tan φ1 + r tan φ2 − 0.5dpf tan φ1 − 0.5dpf tan φ2 )2 (4.140) We now solve Equation 4.113 for F and we get F= π σ v do2 − di2 4 ( ) (4.141) Substitution of Equation 4.131 for σv yields (for a full thread on the bolt) F= π 2 −Y1 do − di2 1 4 G3 + G2 + G1 + G0 + 1.2 ( ) (4.142) We perform a similar operation with Equation 4.136 giving us (for a full thread on the nut) F= © 2014 by Taylor & Francis Group, LLC π 2 −Y2 do − di2 1 4 H 3 + H 2 + H1 + H 0 + 1.2 ( ) (4.143) 156 Ballistics: Theory and Design of Guns and Ammunition Equations 4.142 and 4.143 now contain only two unknowns, r and F. The procedure now involves solving both Equations 4.142 and 4.143 and plotting the force, F versus r. The lowest value in either equation is then the force (and location) at which the joint will fail. It is recommended that these solutions be performed with the aid of a computerized numerical calculation program such as MathCAD. Partial threads can have a significant effect on the failure strength of a joint. If the joint were designed to survive, it is generally best to ignore the additional strength afforded by partial threads and base the design margin on the calculation method earlier. When a joint is designed to fail, however, the lead in and run out must be accounted for unless sufficient margin is available in the expulsion system such that two additional threads may be added to the calculation, yet still allow the joint to be overcome with ease. 4.12 Sabot Design Sabots (French for wooden shoe) are used in both rifled and smoothbore guns to allow a standard weapon to fire a high density, streamlined sub-projectile whose diameter is much smaller than the bore, at a velocity higher than would normally be possible if the gun were sized to the sub-projectile’s diameter. Discarding sabots have been in general use since the Second World War and are still popular (in fact, an artist’s rendition of a discarding sabot is illustrated on the cover of this book). They are called “discarding sabots” since they are shed from the sub-projectile at the muzzle allowing it to fly unencumbered to the target. As stated previously, velocity is proportional to the square root of the pressure achieved in the tube, the area of the bore, the length of travel, and inversely proportional to the square root of the projectile weight. In mathematical terms, V~ pAL wp (4.144) We can see that if the area over which the pressure is applied is much greater than the area presented at the rear of the sub-projectile, a larger force would be applied to accelerate it than if it were fired at the same pressure from a bore of its own diameter. Furthermore, decreasing the launch weight of the as-fired assembly also increases the velocity. Therefore, we must design as light a sabot as feasible so that we can maintain a very dense, small diameter sub-projectile (usually an armor penetrator). The combination of the full bore area, a dense, streamlined sub-projectile, and a lightweight sabot has the overall effect of generating unusually high velocities, a characteristic essential for kinetic energy armor penetration. There are many requirements for a successful sabot: • It must seal the propellant gases behind the projectile (obturate) • It must support the sub-projectile during travel in the bore to provide stable motion (called providing a suitable wheelbase) • It must transfer the pressure load from the propellant gases to the sub-projectile • It must completely discard at the muzzle of the weapon without interfering with the flight of the sub-projectile © 2014 by Taylor & Francis Group, LLC Ammunition Design Practice Rotating band 157 Sabot Sub-projectile FIGURE 4.23 Simplified diagram of an armor-piercing, discarding-sabot (APDS) projectile. • The discarded sabot parts must also fall reliably within a danger area in front of the weapon so as not to injure troops nearby • It must be minimally parasitic, i.e., it must be as light as possible and remove as little energy from the sub-projectile as possible These are formidable requirements that necessitate great ingenuity on the part of the designers. The problem has been solved in a variety of ways. In the 1950s, designers, chiefly British, used cup- or pot-type sabots to launch armor-penetrating, discarding-sabot (APDS) subprojectiles (Figure 4.23). The guns from which these munitions were fired were rifled to launch conventional full caliber, spin-stabilized rounds and so the sub-projectiles of the APDS rounds were spin-stabilized too. Such armor-defeating munitions were highly effective against the tank armor of the times, and pot-type, saboted, kinetic energy penetrators were adopted in tank cannon around the world. Tank armor changed in the 1960s and became more difficult to penetrate with the tungsten carbide cores of the sub-projectiles in use. Initially, incremental changes were made in the material of the core (sintered tungsten was used instead of sintered tungsten carbide), but it was eventually realized that longer, smaller diameter, high-density penetrators were the answer. There are physical limits to the degree of sub-calibering practical in spinstabilized projectiles: the spin required for flight stability increases as the square of the ratio of bore to sub-projectile for conventionally shaped projectiles and it becomes nearly impossible to spin-stabilize very long projectiles. Rifling twists were increased to attempt to accommodate the APDS rounds; in one case, a 1:12 twist was tried when the normal twist would have been 1:40. In the end, APDS designs were abandoned in favor of very long, finstabilized penetrators (APFSDS) that used a radically different type of sabot (Figure 4.24). Obturator Fins Sabot Sub-projectile FIGURE 4.24 Simplified diagram of an armor-piercing, fin-stabilized, discarding-sabot (APFSDS) projectile. © 2014 by Taylor & Francis Group, LLC Ballistics: Theory and Design of Guns and Ammunition 158 laft lsabot lfwd ps ps a T—Shear traction FIGURE 4.25 Free-body diagram for rings and rods. The guns too were changed to smoothbores; although to preserve older weapons in use, designers learned how to make fin-stabilized munitions firable in rifled guns as well. The basic type of sabot used with long-rod, fin-stabilized penetrators is the ring with its subvarieties: base pull, double ramp, and saddle sabots. Whereas pot sabots were essentially discarded rearward as a unit, ring sabots are segmented into three or more sections and discard radially outward at the muzzle to clear the fins that are larger in diameter than the rod. The finned sub-projectile is frequently imparted with a slow spin to average out unavoidable manufacturing asymmetries during flight that could cause trajectory drift. This type of munition is now in the arsenals of all nations. The design of the ring sabot begins with the stress analysis of the shear traction between the sabot inner diameter and the penetrator outer diameter. This analysis is crucial for determining the mass of the ring and thus the parasitic weight of the sabot. We will follow the work of Drysdale [6] throughout this development. The essential parameters of the computation are shown in Figure 4.25. From this free-body diagram, we can infer that T = ps ( A − Ap ) − msabot a (4.145) A reasonable estimate for the masses where the symbols are as follows: msabot = 1 msub-projectile 2 (4.146) where T is the total shear traction force A is the bore area Ap is the area of the penetrator cross section msabot is the mass of the sabot msub-projectile is the mass of the sub-projectile a is the projectile acceleration ps is the pressure on the base of the shot (note that the net pressure on the fins is zero) σ 1 is the axial stress on the penetrator Because the sabot needs to be as light as possible, the material is usually much weaker than the penetrator; thus, the sabot length depends mostly on the sabot material. If the © 2014 by Taylor & Francis Group, LLC Ammunition Design Practice 159 penetrator were weaker for some reason, the sabot length would depend upon that material. Thus, we can write for the surface traction Tallow = π dplsabotτ allow 2 (4.147) where dp is the diameter of the penetrator or sub-projectile Tallow is the allowable traction force τallow is the maximum shear stress allowed in the weaker material The shear traction is usually transmitted through matching grooves or threads. Analysis of these surfaces can be rather complicated but is similar to standard or buttress thread design practice. Given no actual data on the allowable shear stress in the material, we can use the following formulas based on the Tresca or the von Mises yield criteria: σe 2 (4.148) 1.155σ e = 0.577σ e 2 (4.149) τ allow = By the Tresca criteria or τ allow = by the von Mises criteria. In both of these expressions, σe is the equivalent stress as discussed in Section 4.2. Thus, the allowable surface traction can be stated as Tallow = Kπ dplsabotσ e (4.150) where K is either 0.25 or 0.2885 dependent upon the failure criteria. If we substitute Equation 4.150 into Equation 4.145, we can solve for the proper sabot length Kπ dplsabotσ e = ps ( A − Ap ) − msabot a (4.151) and lsabot =  ps  A m a − 1  Ap − sabot  Kπ dpσ e  Ap Kπ dpσ e  (4.152) π 2 dp 4 (4.153)  m a ps dp  A − 1  − sabot  4Kσ e  Ap  Kπ dpσ e (4.154) But Ap = Then, lsabot = © 2014 by Taylor & Francis Group, LLC Ballistics: Theory and Design of Guns and Ammunition 160 Now, by our earlier assumption (Equation 4.146) ps A = ma = (msabot + msub-projectile )a = 3msabot a (4.155) Then, lsabot =  ps dp  A ps a − 1  −  4Kσ e  Ap  3Kπ dpσ e (4.156) Multiplying and dividing the second RHS term by Ap and simplifying, we get lsabot =  ps dp  A ps dp A − 1  −  Kσ e Ap 4Kσ e  Ap 12  (4.157) More generally, if the mass of the sabot is not half of the sub-projectile mass, then we must use Equation 4.154 to determine the proper length. The shape of ring sabots evolved over time from quite heavy designs to highly efficient ones. Early sabots were saddle shaped (Figure 4.26). These had points of high shear stress concentrations near the ends. These sabots had an excellent wheel base (the distance between the forward and aft bourrelets), which prevented balloting in the tube and provided good accuracy. The parasitic weight, however, was high and sufficiently high muzzle velocities were not attained. Single- and double-ramp sabots have come into use because of the favorable weight reduction that can be obtained with this design. They utilize gun pressure to help clamp the sabot to the penetrator and have the added advantage of maintaining an almost constant shear stress between the sabot and the penetrator. The double-ramp sabot is shown in Figure 4.27. Detailed studies have shown that a higher order (nonlinear) curved ramp yields a constant shear stress under load. The method of solution for finding the best shape of the sabot taper depends on a free-body analysis of the sabot and the penetrator. Figures 4.28 and 4.29 show differential elements of the sub-projectile and the sabot, respectively. If we examine Figure 4.28, we see that the axial forces consist of the net internal stress, (dσzp/dz)∆z; the inertial resistance to acceleration, ρpVpa; and the shear stress imparted by Saddle ps Penetrator σ1 τ Shear stress Axial distance FIGURE 4.26 Shear stress variation in a saddle-type sabot. © 2014 by Taylor & Francis Group, LLC Ammunition Design Practice 161 Double-ramp sabot ps Penetrator σ1 τ Shear stress Axial distance FIGURE 4.27 Shear stress variation in a double ramp-type sabot. r Δz τ RP σzp + σzp dσzp dz Δz FIGURE 4.28 Differential element in a sub-projectile showing forces acting. ps r RS(z) σsp + σzs dσzs dz Δz τ Δz FIGURE 4.29 Differential element in a sabot showing forces acting. the sabot, τ. Similarly, on the sabot, we have the net internal stress, (dσzs/dz)∆z; the inertial resistance to acceleration, ρsVs a; the shear stress imparted by the sub-projectile, τ; and the component of pressure in the axial (z) direction. We proceed by initially finding an expression for the volume of the sabot free body. Details of this derivation are found in Ref. [6]. The incremental volume of the sabot can be shown as follows: Vs = π  Rs2 ( z) − Rp2  ∆z © 2014 by Taylor & Francis Group, LLC (4.158) Ballistics: Theory and Design of Guns and Ammunition 162 We then sum the forces on the sabot in the axial direction psπ  Rs2 ( z + ∆z) − Rp2  ∆z − σ zsπ  Rs2 ( z) − Rp2  dσ zs   ∆z  π  Rs2 ( z + ∆ z)Rp2  − ρs Vs a − 2π Rpτ∆z = 0 +  σ zs + dz   (4.159) After collection of terms and simplification, we get ( ps + σ zs ) dRs2  dσ zs  + − ρs a   Rs2 ( z) − Rp2  − 2Rpτ = 0 dz  dz  (4.160) Note here that Rs and σzs are functions of z. Next we find σzp assuming it is linear in z through the expression σ zp = ( ) F 1 1 ρ pπ Rp2 a − 2π Rpτ ∆z + σ 1 ( ρ p Vp a − 2π Rpτ∆z) + σ 1 = = A π Rp2 π Rp2 (4.161) or  2τ  σ zp =  ρ p a −  ∆z + σ 1 Rp   (4.162) where σ 1 is the axial stress in the penetrator as depicted earlier. Now we need to relate σzp to σzs by applying the assumption of strain compatibility, i.e., the strain in the sabot equals the strain in the penetrator. We then use the appropriate elastic moduli and Poisson’s ratio in Hooke’s law to relate the penetrator stresses to those in the sabot ε zs = 1 1 [σ zs −ν s (σ rs + σ θ s )] = ε zp = E σ zp −ν p (σ rp + σ θ p ) Es p (4.163) Es σ zp −ν p (σ rp + σ θ p ) + ν s (σ rs + σ θ s ) Ep  (4.164) Thus, σ zs = If we ignore the bimetallic nature of the components and assume that σ rp + σ θ p = σ rs + σ θ s = −2 ps (4.165) then Equation 4.164 becomes σ zs = © 2014 by Taylor & Francis Group, LLC Es (σ zp + 2ν p ps ) − 2ν s ps Ep (4.166) Ammunition Design Practice 163 r Sabot profile Conical approximation Penetrator OD Reduction of thickness, R(z) by Increasing Es /Ep Decreasing τ Increasing σθ Increasing ps z FIGURE 4.30 Sabot radial profile. (Based on analysis from Drysdale, W.H., Design of kinetic energy projectiles for structural integrity, Technical Report ARBRL-TR-02365, U.S. Army Ballistic Research Laboratory, Aberdeen, MD, September 1981.) These assumptions allow integration of the differential equation for R(z) producing the profile in Figure 4.30 (solid curve). Two of the basic types of sabots are shown in Figures 4.26 and 4.27. The double ramp also incorporates a front air scoop to facilitate discard in the air stream as well as providing additional bourrelets riding surface in the tube. A great deal of work on the effect of sabot design parameters has been accomplished at the U.S. Army Research Laboratory (formerly the Ballistics Research Laboratory) and Picatinny Arsenal. A treatment of the effect of sabot stiffness on how clean a projectile launch is can be found in Ref. [7]. Problem 11 You are to design a rifled 20 mm gun system for an antivehicle application. The gun is to throw a 0.25 lbm, 0.5 in. diameter, cylindrical sub-projectile at 4500 ft/s. A sabot will be placed on the outside of the sub-projectile. The sabot can be any material you like and can find properties for. The sabot shall be made of three pieces, which are at least 3 in. in length. Your design must use a brass cartridge case to assist with obturation of the breech. Other assumptions and information are 1. You do not need to design the breech—assume it will hold the cartridge case in properly (in reality we can always add more threads to the design). 2. Even though there will be a slight taper on the chamber (which must be larger than the bore diameter for seating purposes), assume, for calculation purposes, that the chamber is cylindrical at its maximum diameter. 3. The tube is to be steel and assume that the yield strength is 60,000 psi (this accounts for the effect of cyclic loading). The modulus of elasticity is 29 × 106 psi. Poisson’s ratio is 0.29. 4. Assume the propellant is either cylindrical or single perforated (and state your assumption) © 2014 by Taylor & Francis Group, LLC Ballistics: Theory and Design of Guns and Ammunition 164 5. Choose from the following propellants: Propellant Linearized Burn Rate, β (in./s/psi) Solid Density, δ (lbm/in.3) Adiabatic Flame Temperature, T0 (°R) Propellant Force, λ (ft-lbf/lbm) Specific Heat Ratio, γ IMR M12 Bullseye Red Dot Navy Pyro 0.000132 0.000137 0.000316 0.000153 0.000135 0.0602 0.0600 0.0590 0.0593 0.0566 5103 5393 6804 5774 4477 327,000 362,000 425,000 375,000 321,000 1.2413 1.2326 1.2523 1.2400 1.2454 6. Assume the cartridge case is brass and use a bilinear kinematic hardening model where the brass has a modulus of elasticity of 15 × 106 psi, a local tangent modulus of 13 × 106 psi, and a yield stress of 16,000 psi (yield occurs in this material at ε = 0.002) 7. The weapon shall be as light as possible. The design is to proceed as follows (not necessarily in the order given): a. Interior ballistics design i. Size the chamber length and diameter. ii. Determine the amount of propellant needed based on your choice of the aforementioned propellants and propellant geometry (make sure it fits in the chamber). iii. Determine a web thickness for the propellant. iv. Determine the length of the gun. v. Determine V, pB, and x for the projectile at peak pressure. vi. Determine Vc, pBc, and xc for the projectile at charge burnout. vii. Determine the muzzle velocity of the projectile. b. Gun tube design i. Based on the calculations of part (a), develop a pressure–distance curve to use as criteria for your gun design. ii. Determine the outside diameter of the gun tube. To keep the design light as possible, use the design rules provided in the text and taper the tube toward the muzzle. If needed, over the chamber, you may shrink fit cylinders to build up a composite tube. iii. Determine the weight of your gun and comment on if it is reasonable. c. Cartridge case design i. Determine a thickness and tolerance for your cartridge case. ii. Determine the outside diameter and tolerance for the cartridge case. iii. Decide on a tolerance for your chamber inside diameter. Note that for these calculations you must show that the case may be easily extracted at the limits of the tolerance. © 2014 by Taylor & Francis Group, LLC Ammunition Design Practice 165 d. Sabot design i. Based on a twist rate of 1 in 30, calculate the thickness of a restraining band, 1” wide located over the petal C.G., required to reliably break at shot exit. This band can be considered as a thin cylinder evenly loaded by the three sabot petals (i.e., no moment). “Reliable” breakage can be considered as when the hoop stress is 1.25 times the allowable hoop stress. Neglect the dynamic pressure cause by the forward velocity of the projectile. It is important that you write down all of your assumptions. It is also highly likely that as you proceed further along with your design you may come upon a situation that requires you to revisit an assumption you made earlier—this is to be expected and it is part of the design process. References 1. Budynas, R.G., Advanced Strength and Applied Stress Analysis, 2nd edn, McGraw-Hill, New York, 1999. 2. Boresi, A.P., Schmidt, R.J., and Sidebottom, O.M., Advanced Mechanics of Materials, 5th edn., John Wiley & Sons, New York, 1993. 3. Beer, F.P., Johnston, E.R., and DeWolf, J.T., Mechanics of Materials, 4th edn., McGraw-Hill, New York, 2006. 4. Montgomery, R.S., Interaction of copper containing rotating band metal with gun bores at the environment present in a gun tube, Report AD-780-759, Watervliet Arsenal, New York, June 1974. 5. Pangburn, D., Personal communications with author December 1995 to March 2004. 6. Drysdale, W.H., Design of kinetic energy projectiles for structural integrity, Technical Report ARBRL-TR-02365, U.S. Army Ballistic Research Laboratory, Aberdeen, MD, September 1981. 7. Plostins, P., Clemins, I., Bornstein, J., and Diebler, J.E., The effect of sabot front borerider stiffness on the launch dynamics of fin-stabilized kinetic energy ammunition, BRL-TR-3047, U.S. Army Ballistic Research Laboratory, Aberdeen, MD, October 1989. Further Readings Barber, J.R., Intermediate Mechanics of Materials, McGraw-Hill, New York, 2001. Ugural, A.C. and Fenster, S.K., Advanced Strength and Applied Elasticity, 3rd edn., Prentice Hall, Upper Saddle River, NJ, 1995. © 2014 by Taylor & Francis Group, LLC © 2014 by Taylor & Francis Group, LLC Part II Exterior Ballistics © 2014 by Taylor & Francis Group, LLC © 2014 by Taylor & Francis Group, LLC 6 Introductory Concepts When the projectile has left the environment of the gun, including the effects of the exiting propellant gases that momentarily surround it, it enters the realm of the exterior ballistician. Here, it is subject to the force of the pressure of the atmosphere that it is flying through, the force induced by its spin, and the force due to the acceleration of gravity. The projectile in flight is no longer constrained in lateral motions by the walls of the gun and, as a free body, can develop motions that are complex and occasionally inimical to the intent of its user and embarrassingly, to its designer. The study of these motions and the progress of the projectile to its target are the subject of this part of the book. We will begin with the simplest case, consideration of the projectile as a point mass flying in a vacuum with only the force due to gravity acting on it. Then we will proceed to introduce the force due to the pressure of the air, but still considering the projectile as a mass concentrated at a point. Finally, we will consider the projectile as a three-dimensional body acted upon by the air, its spin, and gravity. In the final sections of this part of the text, we shall examine the complex motions arising from the coupling of projectile dynamics and aero-mechanical forces. Our object will be to examine the conditions necessary for a precise, predictable, satisfactory trajectory enabling the projectile to fulfill its terminal ballistic utility. Since this text is intended to have a broad scope, some of the material is not derived in detail. The reader is encouraged to seek the more detailed treatments in the references noted throughout each section. Many of the principles and terms concerned with fluid mechanics required for the understanding of interior ballistics were introduced in Section 2.7. These principles will be extended in this section with a view toward an exterior ballistician—commonly called an aero-ballistician. We shall first examine the elements of a trajectory as depicted in Figure 6.1. These terms are commonly used in the military by gunners and researchers alike. Although most of the symbols and terms in this figure are self-explanatory, some require comment. First is the so-called map range. This is the range to the target that the gunner would see if he or she were to plan firing using a map. The base of the trajectory is quite important and is defined as being level in a plane with the firing point. Gunners of large caliber weapons and mortars take great care in assuring that the sights on the weapon are leveled in the direction depicted as well as the plane out of the paper. Since larger ordnance fires over extensive ranges, it is common to assume that the origin of the trajectory is coincident with the ground beneath the artillery piece. The line of site and angle of site (yes, they are spelled that way in much of the literature) are what the gunner uses to aim at the target. As you can see, they only assist in determination of the pointing of the weapon and the relative height of the target. An important feature of this diagram is the line of departure. You have probably noticed that it is not collinear with the elevation of the weapon (i.e., where the bore is pointed). The reality is that a projectile almost never leaves the bore of a gun aligned with the bore—we shall discuss this in detail later. For now, we will simply state that this is due to 193 © 2014 by Taylor & Francis Group, LLC Ballistics: Theory and Design of Guns and Ammunition 194 Vertical jump Quadrant elevation of departure Origin e f eo pa de n Li Quadrant elevation r rtu n Trajectory atio lev fe eo Lin Maximum ordinate Line of fall Angle of lift Base of trajectory Level point Angle of site Line Angle of elevation of sit e Point of impact Map range FIGURE 6.1 Elements of a trajectory. the dynamics of the projectile and gun as well as aerodynamic effects. It should be noted that Figure 6.1 is drawn as two dimensional. The out-of-plane angular position of the projectile at muzzle exit is known as lateral or azimuthal jump. This will combine vectorially with the vertical jump that is depicted to give a resultant jump vector. The angle of lift and line of fall are defined for the level point; however, it is common to see these used at the target even though officially these quantities at the target are called angle of impact and line of impact (sometimes shot line). The aerodynamics and ballistics literature are quite diverse and terminology is far from consistent. This has particular significance in the coordinate systems used to define the equations of motion. In this text, we shall use the coordinate system of Ref. [1] as depicted in Figure 6.2. The primary difference between this scheme and those of, say, Refs. [2–5], is that the y-axis is deemed to be positive pointing up, with the z-axis as positive to the right as opposed to the z-axis down and y-axis to the right. This makes sense to the authors with up being a more intuitive positive direction. The only issues (and some people consider them significant) with this scheme are that first, the nice right-handed naming convention y Positive yaw x Positive roll z Positive pitch FIGURE 6.2 Definition of projectile coordinates. © 2014 by Taylor & Francis Group, LLC Introductory Concepts 195 x B α V sin αt αt O A β C V cos β V cos β sin α x V sin β αt V V Trajectory FIGURE 6.3 Generalized yaw of a projectile. of the aerodynamic coefficients is disturbed (as we shall see later x–y–z corresponds to l–n–m not l–m–n as one would normally like); second, what would normally be a positive rotation in the y-direction (i.e., nose left) is defined as negative—we shall handle this when we define the associated equations. We shall now define some terminology and, more importantly, the forces, moments, and associated coefficients that are used throughout this part of the text. It is important for the reader to recognize that these force, moment, and coefficient definitions are by no means an all-inclusive collection. Occurrences of additional forces or moments at times require additional definitions—e.g., control deflections. We shall adhere to the broad scope of this text by including only what is necessary for a basic understanding of ballistics. We mentioned the yaw and pitch of the projectile earlier in this section. The projectile geometry in an arbitrary state of yaw is depicted in Figure 6.3. This illustration shows the projectile yawed and pitched to some angle, αt, relative to the velocity vector. The illustration also shows the trajectory, which is defined as the curve traced out by the velocity vector. Thus, the velocity vector is everywhere tangent to the trajectory curve. The inset shows the decomposition of the angle between the projectile axis of symmetry, x (OB), and the velocity vector, V (OA). We first measure the sideslip angle, β (∠AOC), and then measure the yaw angle, α (∠COB), from the axis of symmetry, x, to side OC = V cos β. The side BC of right triangle OBC then has a value of V cos β sin α. The resulting angle ∠AOB is defined as the total yaw angle, αt, and in triangle AOB where side AB = V sin α t . It should be noted that triangle ABC with sides V sin αt, V cos β sin α, and V sin β is not a right triangle. Most projectiles have at least trigonal symmetry. This is symmetry about three planes through the projectile long axis, 120° apart. Because of symmetry, it is common to vectorially combine the yaw and pitch of the projectile into one term, which we simply call total yaw, αt. All of our coefficients will be based on this total yaw. Later, when we discuss advanced topics, it will be necessary to once again separate them. An examination of Figure 6.3 shows that we can relate the total yaw to α and β through sin α t = sin 2 β + cos 2 β sin 2 α (6.1) The drag on a projectile is the force exerted on it by the medium through which it is moving, usually air. Since the drag is generated by the motion of the projectile through the air, it is naturally directed opposite to the velocity vector as illustrated in Figure 6.4. There are, in general, two types of drag: pressure drag and skin friction drag. This is because nature can only act on the surface area of the projectile in two ways: normal to the © 2014 by Taylor & Francis Group, LLC Ballistics: Theory and Design of Guns and Ammunition 196 x Drag force αt V Trajectory FIGURE 6.4 Drag of a projectile. surface and along it. A third type of drag called wave drag is a form of pressure drag generated by a shock wave formed when the local velocity along the surface of the projectile reaches Mach 1. We will discuss drag in further detail later, but in all cases it is convenient to lump the effect of the drag into one coefficient called the drag coefficient. The drag force is defined in terms of this drag coefficient as 1 1 Drag force = FD = − ρ SCD VV = ρV 2SCD 2 2 (6.2) Equation 6.2 shows two forms of the defining expression for drag force, vector, and scalar. We shall define all of our forces and moments in this way because; although we will initially examine the scalar forms, it will be necessary later to use the vector forms. For now, knowing that the drag force is opposite to the velocity vector and that its scalar magnitude, as depicted on the far RHS in Equation 6.2, is sufficient. Like many of the coefficients we shall discuss, the drag coefficient can be a complicated function of the yaw angle. In a more general form, we can write the drag coefficient as the sum of a linear part and a yaw-dependent term: CD = CD 0 + CD δ 2 δ 2 (6.3) δ = sin α t (6.4) where δ is the total yaw defined as The first term on the RHS is the linear part of the drag coefficient, known as the zeroyaw drag coefficient, while the second term is known as the yaw drag coefficient. The reason that there is no intermediate term is that, for a symmetric body, the drag has to be the same whether the body is angled at, say +5° or −5°. This is discussed more elegantly in Ref. [2]. We shall see later that the drag coefficient varies with Mach number in a complex manner. Dynamic pressure is a quantity defined as 1/2ρV2, where ρ is the density of a fluid that an object is immersed in and V is the velocity of the fluid relative to the object. It is simply the physical reaction of the fluid when trying to force an object through it and occurs so often that it has been given its own name. This dynamic pressure is multiplied by a reference area, S. It is always important to know what reference area is used in the definition of the coefficients. In every case we shall examine, this reference area is based on the projectile circular cross section. Also, as we shall soon see, moments require a length scale as well. In all of these instances, we shall use the projectile diameter as the reference length. © 2014 by Taylor & Francis Group, LLC Introductory Concepts 197 x, p Mlp αt V Trajectory FIGURE 6.5 Spin damping of a projectile. When a projectile spins in a medium, the viscous interaction of the medium and the projectile surface is such that the projectile will spin down throughout the flight. This phenomenon is accounted for by a moment applied to the projectile called the spin-damping moment. It is defined as Spin-damping moment = M1p = 1  pd  ρV 2Sd   C1p 2 V  (6.5) This moment is directed opposite to the spin vector, p, of the projectile as depicted in Figure 6.5, and the tendency is for the projectile to spin down, thus there is no negative sign in Equation 6.5 because the vector handles the decay. One needs to note that the figure is drawn for a right-hand twist. If a left-hand twist were involved, the spin vector, p, and the spin-damping moment vector would be reversed. Some projectiles have fins or jets that impart a roll torque to the projectile such that the spin rate increases. This rolling moment is depicted in Figure 6.6 and defined through Rolling moment = M1δ = 1 ρV 2Sdδ FC1δ 2 (6.6) In this expression, δ F is a cant angle provided to the fins to generate the lift required to sustain rotation. Lift is defined as the aerodynamic force that acts orthogonal to the velocity vector. This is depicted in Figure 6.7. The lift force can be defined in both scalar and vector notations as Lift force = FL = 1 1 ρ SCLα [V × (x × V )] = ρV 2SCLα δ 2 2 (6.7) x M1δ αt V Trajectory FIGURE 6.6 Roll moment of a projectile. © 2014 by Taylor & Francis Group, LLC Ballistics: Theory and Design of Guns and Ammunition 198 x αt FL V Trajectory FIGURE 6.7 Lift vector of a projectile. The lift force coefficient can be written in its nonlinear form as CLα = CLα 0 + CLα 2 δ 2 (6.8) With a symmetric projectile, we must note that if there is no angle of attack (i.e., δ = 0) there is no lift. This is obvious even for the linear case since δ appears in Equation 6.7. Some authors prefer to work in coordinates other than those we are utilizing here. In those cases, expressions such as Cx and CN are used for drag and lift, respectively. In these cases, it is important that proper transformations are used to change the coefficients. An example of this is provided in Ref. [1]. At this point, we must discuss two quantities known as center of pressure (CP) and center of gravity (CG) (sometimes called center of mass). The CG is the location on the projectile where all of the mass can be concentrated so that for an analysis the gravitational vector will operate at this point. The CP is the point through which a vector can be drawn, i.e., the resultant of all infinitesimal pressure forces acting on the projectile. For most projectiles that are spin-stabilized, the CP is ahead of the CG and the reverse is true with finor drag-stabilized projectiles. Figure 6.8 is an illustration of this. The separation of the CP and CG gives rise to an overturning moment in all projectiles (Figure 6.9). As we shall see later, this moment is destabilizing for spin-stabilized projectiles (which is why they must be spun) and stabilizing for fin-stabilized projectiles. The overturning moment (sometimes called the pitching moment) is defined as Overturning moment = Mα = 1 1 ρ SdVCMα (V × x) = ρV 2SdCMα δ 2 2 CP CG (center of mass) FIGURE 6.8 Center of gravity (CG) and center of pressure (CP) illustrated. © 2014 by Taylor & Francis Group, LLC (6.9) Introductory Concepts 199 x αt Mα V Trajectory FIGURE 6.9 Overturning moment vector of a projectile. We can see from Equation 6.9 that this moment is a function of the angle of attack and because of the cross-product, a positive overturning moment (nose up) is oriented along the positive z-axis. The overturning moment coefficient can be written in a nonlinear form similar to the lift and drag forces as CMα = CMα 0 + CMα 2 δ 2 (6.10) When a body of circular cross section is immersed in a flow-field perpendicular to its axis and is spun about its axis, a force known as the Magnus force is developed [6]. This force comes about because on one side of the body the free stream velocity of the flow is added to the velocity of the surface, while on the other side the free stream velocity is reduced by the surface velocity. On the basis of Bernoulli’s equation (Equation 6.11), we see that along the body surface streamline the pressure must be higher on the side with the lower velocity [7]: p 1 2 + V + z = constant ρ 2 (6.11) This results in a side force on the body as illustrated in Figure 6.10. This might not seem like a big deal because a projectile almost never flies sideways, but if we consider a projectile in a crosswind, or, more importantly, one that is yawed, we see that this side component can contribute somewhat to the aerodynamic loading. For all Angular velocity, ω Upper surface velocity = rω – V∞ Free stream velocity, V∞ Body radius, r Lower surface velocity = rω + V∞ FIGURE 6.10 Magnus effect on a projectile. © 2014 by Taylor & Francis Group, LLC Body will move in this direction—Magnus force direction Ballistics: Theory and Design of Guns and Ammunition 200 x αt FNpα V MMpα Trajectory FIGURE 6.11 Magnus force and moment on a projectile. practical purposes, however, if a projectile is not yawed in flight, then there is no Magnus force. The Magnus force is defined for our purposes as Magnus force = FNpα = 1 1  ωd  2  ωd  ρ SV   CN pα δ  CN pα (V × x) = ρV S  2 2 V  V    (6.12) The Magnus force coefficient can be written in a nonlinear form in the same manner as Equation 6.10, which we will not repeat (Figure 6.11). In many cases, the Magnus force is small and is usually neglected with respect to the other forces acting on the projectile. In contrast, the moment developed because of this force is considerable. We define the Magnus moment as Magnus moment = M Mpα = 1 1  ωd   ωd  2 ρVSd   CM pα [x × (V × x)] = ρV Sd   CM pα δ 2 2 V    V  (6.13) The Magnus moment contributes significantly to the stability of the projectile and will be discussed in detail later. The Magnus moment coefficient can be written as a nonlinear term in the same way as all our other coefficients. The CP for the lift force and the CP for the Magnus force are usually not the same, thus the moments will act through differing moment arms. The reason for this is the different physics that give rise to the different phenomena. These change during flight as well since a projectile’s yaw changes as it moves downrange. Pitch damping is the tendency of a projectile to cease its pitching motion due to air resistance. It is usually more difficult to visualize for someone new to the field. It is relatively simple to think about a right circular cylinder mounted in a fixture with its spin axis held by a frictionless bearing on each end. If we spin the projectile, it will slow down because of the sticking of the fluid to the surface and the resultant viscous action (remember the bearings are magically frictionless). If we mount the projectile such that the bearing is transverse to the long axis and spin it, we will still have the viscous action slowing the projectile down; however, this will be overwhelmed by the pressure forces that retard the motion and the projectile will spin down much faster. This combination of forces is called pitch damping. For projectiles, we can define the pitch damping force as Pitch damping forces = F Nq +αɺ = © 2014 by Taylor & Francis Group, LLC 1 1  dx dl   dx  −  ρVSd   CNq + ρVSdCNαɺ  2 2  dt dt   dt  (6.14) Introductory Concepts 201 Motion with q = sin ωt α=0 Motion with q=0 α = sin ωt FIGURE 6.12 . Pictorial description of q and α. or in scalar terms, Pitch damping forces = FNq +αɺ =  q d   1  αɺ d  ρV 2S  t  CNq +  t  CNαɺ  2  V   V   (6.15) In Equation 6.16, we have defined the total pitching motion, qt, and the total rate of change of angle of attack, α̇t, as qt = q2 + r 2 and αɺ t = dα t dt (6.16) We note here that this pitch damping comes about through two motions. The first motion is brought about through the pitching rate q, while the second is developed because of the resistance to the changing angle of attack. This is described in eloquent detail in Ref. [5]. The simplest way of depicting this is to assume a sinusoidal motion of a projectile along its flight path. With this assumption, Figure 6.12 shows what motions would result if q only . was present and contrasts this with motion if α only were present. . It is generally difficult to separate q and α in experimental flight data. For this reason, the two coefficients are almost always written as a sum and recorded in the literature as such. With assumptions on the yawing motion of the projectile and the practice of combining coefficients, as described previously in Equations 6.14 and 6.15, they can be combined as detailed in Ref. [1] into Pitch damping force = FNq +αɺ = = 1 dx ρVSd(CNq + CNαɺ ) 2 dt 1 qd ρV 2Sd  t  (CNq + CNαɺ ) 2  V  (6.17) The pitch damping force is, like the Magnus force, generally neglected because it is small with respect to the other forces such as lift and drag. The moment caused by this pitch damping is frequently significant (Figure 6.13). It can be described mathematically as follows: Pitch damping moment = M Mq +αɺ =  dx   dl   1 dx  1  2 ρVSd 2  x ×  −  x ×   CMq + ρVSd C M αɺ  x × dt   dt   2 dt  2   © 2014 by Taylor & Francis Group, LLC (6.18) Ballistics: Theory and Design of Guns and Ammunition 202 x FNq+α qt αt V, l Trajectory MMq+α FIGURE 6.13 Pitch damping force and moment on a projectile. In scalar form, we can write Pitch damping moment = M Mq +αɺ =  q d   1  αɺ d  ρV 2Sd  t  C Mq +  t  CMαɺ  2 V V       (6.19) 1 dx   ρVSd 2 (CMq + CMαɺ )  x ×  2 dt   (6.20) 1 qd ρV 2Sd  t  [CMq + CMαɺ ] 2  V  (6.21) These can be simplified as per Ref. [1] into Pitch damping moment = M Mq+αɺ = Pitch damping moment = M Mq +αɺ = At certain times and in some special cases, there are other combinations of forces and moments and therefore additional coefficients require attention. We will not go any further here as this text is meant to be most general. We now have the basic terms defined that we shall use in our study of exterior ballistics. Problem 1 In a test range, a 0.50 caliber M33 ball projectile has a velocity vector which is at an angle of 10° to the horizontal (assume zero azimuth) with a velocity of 3013 ft/s. The initial pitch and yaw angles are 1.030° and 1.263°, respectively. The initial rotational rate of change of the axial unit vector (dx/dt) is provided later. If the projectile has the coefficients below at this particular instant, draw the situation and determine the following: a. b. c. d. e. f. Velocity vector (ft/s) Projectile axial unit vector (x) Drag force vector (lbf) Spin damping moment vector (lbf-in) Overturning moment vector (lbf-in) Magnus moment vector (lbf-in) Assume the weapon has a right-hand twist. © 2014 by Taylor & Francis Group, LLC Introductory Concepts 203 Projectile information: CD = 0.2938 (CMq + CMαɺ ) = −5.5 I P = 7.85 [g cm 2 ] CMα = 2.88 (CNq + CNαɺ ) = 0.004 I T = 74.5 [g cm 2 ] CLα = 2.69 CMpα = 0.05 m = 42.02 [g]  lbm  ρ = 0.0751  3   ft   rad  ω = 15, 404   s  CNpα = −0.01 Clp = −0.003 dx  rad  = {0.405e1 − 1.963e2 − 0.981e3 }  dt  s  Please supply all answers in an inertial coordinate system labeled 1, 2, 3 with 1 being along the downrange direction and 3 being to the right side. Treat all missing coefficients as equal to zero. It is very important that you DRAW the situation. This will have a great deal of influence in obtaining the correct answer References 1. McCoy, R.L., Modern Exterior Ballistics, Schiffer Military History, Atglen, PA, 1999. 2. Murphy, C.H., Free flight motion of symmetric missiles, Ballistics Research Laboratory Report No. 1216, Aberdeen Proving Ground, MD, 1963. 3. McShane, E.J., Kelley, J.L., and Reno, F.V., Exterior Ballistics, University of Denver Press, Denver, CO, 1953. 4. Nicolaides, J.D., On the free flight motion of missiles having slight configurational asymmetries, Ballistics Research Laboratory Report No. 858, Aberdeen Proving Ground, MD, 1953. 5. Nielsen, J.N., Missile Aerodynamics, AIAA reprint, American Institute of Aeronautics and Astronautics, Reston, VA, 1988. 6. White, F.M., Fluid Mechanics, 5th edn., McGraw-Hill, New York, 2003. 7. Fox, R.W. and McDonald, A.T., Introduction to Fluid Mechanics, 4th edn., John Wiley & Sons, New York, 1992. Further Reading Bull, G.V. and Murphy, C.H., Paris Kanonen—The Paris Guns (Wilhelmgeschutze) and Project HARP, Verlag E.S. Mittler & Sohn GmbH, Herford und Bonn, Herford, Germany, 1988. © 2014 by Taylor & Francis Group, LLC © 2014 by Taylor & Francis Group, LLC 7 Dynamics Review Throughout the study of exterior ballistics, dynamics play a great role in the flight of the projectile. The Coriolis effect in long-range trajectories or the drag changes due to the precessional and nutational motion of the projectile are just two examples of the effect of projectile body dynamics on flight. We will find that at least a cursory review of dynamics is essential to the understanding of projectile motion. Analyzing dynamics of projectile flight is best approached through the use of vectors and we will begin our review with their study. A vector is defined as a quantity having a magnitude and a direction. Two vectors are considered equal if both their magnitude and direction are identical. However, this does not mean that they have to originate at the same point, i.e., a translation has no effect on whether vectors are equal. A scalar is simply a numerical quantity (a magnitude). When a scalar and a vector are multiplied (in any order) they form a vector. Thus, we can define any vector as a scalar magnitude multiplied by a vector of unit length (a unit vector) in the proper direction (Figure 7.1). A = Ae A (7.1) A vector can be written as the sum of its scalar magnitude in each individual coordinate direction times a unit vector in that particular direction. A = Axi + Ay j + Az k (7.2) The magnitude of the vector is defined as A =|A|= Ax2 + Ay2 + Az2 (7.3) Vectors may be added together in any order by summing the individual components in each direction. This is the commutative property: A + B = B + A = ( Ax + Bx )i + ( Ay + By )j + ( Az + Bz )k (7.4) The following is also true when adding more than one vector together: (A + B) + C = A + (B + C) (7.5) Equation 7.5 represents the associative property of vectors. In all of the above expressions, note that i, j, and k are the unit vectors in the x, y, and z coordinate directions, respectively. Multiplication of vectors can occur in two different ways—each applicable to particular situations. Consider two vectors A and B shown in Figure 7.2, we define the scalar product or dot product as A ⋅ B = A ⋅ B ⋅ cosθ (7.6) 205 © 2014 by Taylor & Francis Group, LLC Ballistics: Theory and Design of Guns and Ammunition 206 1 A A eA FIGURE 7.1 Vector and associated unit vector. B θ A FIGURE 7.2 Vector pair illustrated. Both the commutative and associative laws of multiplication apply to the dot product. A ⋅B = B⋅A (7.7) (A + B) ⋅ C = A ⋅ C + B ⋅ C (7.8) The dot product of two vectors is given by A ⋅ B = ( Axi + Ay j + Az k) ⋅ (Bxi + By j + Bz k) (7.9) This equation when expanded is A ⋅ B = Ax Bxi ⋅ i + Ax Byi ⋅ j + Ax Bzi ⋅ k + Ay Bx j ⋅ i + Ay By j ⋅ j + Ay Bz j ⋅ k + AzBx k ⋅ i + AzBy k ⋅ j + AzBz k ⋅ k (7.10) However, since the unit vectors are orthogonal, and the dot product of two orthogonal vectors is identically zero while the dot product of parallel vectors is unity as follows from i ⋅ i = j ⋅ j = k ⋅ k = 1 ⋅ 1 ⋅ cos(0 ) = 1 i ⋅ j = j ⋅ i = j ⋅ k = k ⋅ j = i ⋅ k = k ⋅ i = 1 ⋅ 1 ⋅ cos(90 ) = 0 Therefore, we can write A ⋅ B = Ax Bx + Ay By + AzBz © 2014 by Taylor & Francis Group, LLC (7.11) Dynamics Review 207 B en θ A FIGURE 7.3 Vector cross product normal unit vector. The second type of vector multiplication is the vector or cross product, which is defined as A × B = A ⋅ B ⋅ sinθ en (7.12) Here en is a unit vector normal to the plane made by vectors A and B. This is depicted in Figure 7.3. The cross product does not obey the commutative property because A × B = −B × A (7.13) The distributive property, however, does apply to the cross product. Thus, (A + B) × C = A × C + B × C (7.14) The cross product of two vectors is given by A × B = ( Axi + Ay j + Az k) × (Bxi + By j + Bz k) (7.15) When expanded, Equation 7.15 can be written as A × B = Ax Bxi × i + Ax Byi × j + Ax Bzi × k + Ay Bx j × i + Ay By j × j + Ay Bz j × k + AzBx k × i + AzBy k × j + AzBz k × k (7.16) Since the unit vectors are orthogonal, i × i = j × j = k × k = 1 ⋅ 1 ⋅ sin(0 )en = 0 and i × j = 1 ⋅ 1 ⋅ sin(90 ) = en But, since we have a right-handed coordinate system, by the right-hand rule, the normal to i and j is the unit vector k, thus i × j = k. We can also invoke Equation 7.13 to get j × i = −i × j = −k. We can carry this logic further to show that j × k = i or k × j = −i and i × k = −j or k × i = j. Thus, we can rewrite Equation 7.16 as A × B = ( Ay Bz − AzBy )i + ( AzBx − AxBz )j + ( AxBy − Ay Bx )k (7.17) This Equation 7.17 is the following determinant expanded by its minors: i A × B = Ax Bx © 2014 by Taylor & Francis Group, LLC j Ay By k Az Bz (7.18) Ballistics: Theory and Design of Guns and Ammunition 208 A + ΔA ΔA A FIGURE 7.4 Vector sum illustrated. We will proceed next to the calculus of vectors. Let us consider a vector, A, dependent upon a scalar variable, u, as shown in Figure 7.4. Then A + ∆A corresponds to u + ∆u and we can write for its derivative dA ∆A = lim du ∆u →0 ∆u (7.19) d(A + B) dA dB = + du du du (7.20) Differentiation is distributive so that The chain rule also applies for scalar and vector products so that dg d dA ( gA) = A+ g du du du (7.21) d dA dB (A ⋅ B) = ⋅B + A ⋅ du du du (7.22) d dA dB (A × B) = ×B + A× du du du (7.23) Consider a vector A dependent upon time, t. If we take its derivative with respect to time, we get dAy dA dAx dAz di dj dk i+ j+ k + Ax = + Ay + Az dt dt dt dt dt dt dt (7.24) If the coordinate system is inertial (i.e., it does not move), we can write dAy dA dAx dAz i+ j+ k = dt dt dt dt (7.25) If the coordinate system is moving (like on a rotating earth), the rate of change terms for the unit vectors cannot be neglected. This gives rise to what we call “Coriolis terms” as we shall discuss later. © 2014 by Taylor & Francis Group, LLC Dynamics Review 209 y P r O x z FIGURE 7.5 Position vector. We will now examine the kinematics of a particle. Kinematics is the study of the motion of particles and rigid bodies without regard to the forces which generate the motion. Particle kinematics assumes that a point can represent the body. The rotations of the particle itself are neglected making this a three degree of freedom (DOF) model. If we have the inertial reference frames x, y, and z, the position of a particle, P, is defined by a position vector, r, drawn from the origin to the particle as is shown in Figure 7.5. If the particle, P, moves along a trajectory, T, its instantaneous velocity is always in a direction tangent to the trajectory and its magnitude is the speed at which it moves along the curve. Thus, the tip of this vector, r, traces out the trajectory (Figure 7.6) and the velocity, v, is defined as the time rate of change of r, written as v= dr dt (7.26) If we were to take the velocity vector at every instant of time and fix its tail to the origin of an inertial coordinate system, then the curve traced out by its tip would be called a hodograph (Figure 7.7) and the velocity of the tip would be the time rate of change of velocity or the acceleration, a. Thus, we can write a= dv d 2r = dt dt 2 (7.27) y v T P O r x z FIGURE 7.6 Trajectory curve. © 2014 by Taylor & Francis Group, LLC Ballistics: Theory and Design of Guns and Ammunition 210 y a v O x z FIGURE 7.7 Hodograph. er eθ r θ O FIGURE 7.8 Differentiation of a vector through use of tangential and radial unit vectors. Now, if we examine the particle as moving in two dimensions only, we can break its motion up into two components, one parallel to and one perpendicular to the position vector, r (Figure 7.8). The position vector written in this coordinate system is given by r = re r (7.28) So from our definition for the velocity in Equation 7.26, we get v= dr dr de = er + r r dt dt dt (7.29) Since er is a unit vector (its magnitude is a constant = 1) the only thing that changes with time is its direction. This introduces the concept of curvilinear motion with radial coordinates, (r, θ). The direction is defined by the angle, θ. For a small change in the angle, θ, we can write der ∆e = lim r dt ∆t →0 ∆t (7.30) However, we can observe that for small angles ∆er =|er |sin(∆θ ) = (1)sin(∆θ ) ≈ ∆θ © 2014 by Taylor & Francis Group, LLC (7.31) Dynamics Review 211 Δe r er er + Δer eθ Δθ FIGURE 7.9 Rotation of the radial unit vector. We also see from Figure 7.9 that ∆er acts in the eθ direction thus, ∆er ≈ ∆θ eθ (7.32) Then, returning to Equation 7.30 we can write ∆θ dθ der = eθ lim = eθ t ∆ 0 → ∆t dt dt (7.33) Now, we can insert Equation 7.33 into Equation 7.29 to get the desired relation for the velocity. v= dθ dr dr = er + r eθ dt dt dt (7.34) The first term on the RHS of Equation 7.34 is the radial velocity, the second term is the tangential velocity sometimes denoted as vr and vθ, respectively. The magnitude of the velocity is given by 2  dr   dθ  v =|v|= vr2 + vθ2 =   +  r   dt   dt  2 (7.35) To obtain the acceleration in curvilinear coordinates, we need to take the time derivative of Equation 7.34 as follows: a= d 2θ dθ deθ dv d  dr dθ  d 2r dr der dr dθ + =  er + r eθ  = 2 er + eθ + r 2 eθ + r dt dt dt dt dt  dt dt  dt dt dt dt dt (7.36) We have already solved for the derivative of er with respect to time; now, in a similar manner, we will find the derivative of the tangential component. Again, since eθ is a unit vector, the only thing that changes with time is its direction. This direction is again defined by the angle, θ, so for a small change in the angle, θ, we can write deθ ∆e = lim θ ∆ t → 0 ∆t dt © 2014 by Taylor & Francis Group, LLC (7.37) Ballistics: Theory and Design of Guns and Ammunition 212 Δeθ eθ er eθ + Δeθ Δθ FIGURE 7.10 Rotation of the tangential unit vector. But we see again that for small angles ∆eθ =|eθ|sin(∆θ ) = (1)sin(∆θ ) ≈ ∆θ (7.38) which acts in the negative er direction as depicted in Figure 7.10. If ∆eθ ≈ −∆θ er (7.39) ∆θ deθ dθ er = −er lim =− ∆t → 0 ∆t dt dt (7.40) then we can write, Insertion of Equations 7.33 and 7.40 into Equation 7.36 yields 2 a= d2r dr dθ dr dθ d 2θ  dθ  r e + e + e + eθ − r  θ θ r  er 2 2 dt dt dt dt dt dt  dt  (7.41) Rearranging and combining like terms gives us 2  d 2r  d 2θ dr dθ   dθ   + a =  2 −r e  r r 2 + 2  eθ  dt dt   dt    dt  dt Each of these terms has a specific name and meaning in the dynamics of a body d2r = Radial acceleration dt 2 2  dθ  r  = Centripetal acceleration  dt  d 2θ = Angular acceleration d 2t 2 © 2014 by Taylor & Francis Group, LLC dr dθ = Coriolis acceleration dt dt (7.42) Dynamics Review 213 y΄ y B rB/A rB O x΄ Body-fixed coordinate A rA system (translates with the body and the body does not rotate) x Inertial coordinate system (fixed in space) rA = Position vector of point A rB = Position vector of point B rB/A = Relative position vector of point B with respect to point A vA = Velocity of point A vB = Velocity of point B vB/A = Relative velocity of point B with respect to point A aA = Acceleration of point A aB = Acceleration of point B aB/A = Acceleration of point B with respect to point A FIGURE 7.11 Definition of vectors associated with rigid body translational motion. To move on in our study, we need to examine the planar kinematics of a rigid body. First, we will examine a pure translation where we have a body-fixed coordinate system moving relative to our inertial coordinate system. Here we note from Figure 7.11 that by vector addition, we obtain rB = rA + rB/A (7.43) To determine the velocity of point B, which is under a pure translation, we have to differentiate Equation 7.43 to get vB = drB drA drB/A = + dt dt dt (7.44) We know, however, that since this is a pure translation (no rotation) drB/A/dt = 0 and drA/dt = vA. Thus, for a pure translation, vB = v A (7.45) If we differentiate Equation 7.45, we get the acceleration of a point during a pure translation as aB = dv B dv A = = aA dt dt (7.46) We will now examine the rotation of a body fixed in space. Let us define the angular velocity, ω, as the time rate of change of angular position, θ, thus ω= dθ dt (7.47) Let us further define the angular acceleration, α, as the time rate of change of angular velocity, or α= dω d 2θ = 2 dt dt (7.48) The angular velocity, ω, and the angular acceleration, α, are depicted in Figure 7.12. We will now look at the rotation in terms of the vector kinematic equations for point, P. We first © 2014 by Taylor & Francis Group, LLC Ballistics: Theory and Design of Guns and Ammunition 214 α, ω, dθ ω α O O r P dθ r P θ dθ θ FIGURE 7.12 Example of rigid body rotation. On the left is the body rotating in space. On the right is a view of this same body looking down the axis toward O. ω ω O r v O P r P rP E FIGURE 7.13 Rigid body rotation. On the left is the body rotating at angular velocity, ω. On the right is a view of this same body looking down the axis toward O. examine the velocity whose direction we specify by the right-hand rule. Then let us define the position of point, P, by the position vector, r, as shown in Figure 7.13. Now we can write the velocity, v, in terms of radial and circumferential components as we discussed earlier. Thus, from Equation 7.34 we have v= dθ dr dr = er + r eθ dt dt dt (7.49) But, since this is a rigid body, dr/dt = 0, so we get v=r dθ eθ = rωeθ dt (7.50) If we were instead to draw the position vector from a more general location such as point E in Figure 7.13, we see that r = rP sin φ (7.51) If we substitute Equation 7.51 into Equation 7.50, we see a form we have derived earlier. v = rP ω sin φ eθ (7.52) This can be written in vector form if we invoke Equation 7.12. Thus, we have v = ω × rP © 2014 by Taylor & Francis Group, LLC (7.53) Dynamics Review 215 ω, α ω, α O r an at P O an at r rP P E FIGURE 7.14 Rigid body rotation with acceleration. On the left is the body rotating at angular velocity, ω, and accelerating with angular acceleration, α. On the right is a view of this same body looking down the axis toward O. We now will examine the acceleration whose direction is once more specified by the righthand rule. We shall define the position of point, P, by the position vector, r, as shown in Figure 7.14. We can then write the acceleration, a, in terms of radial and circumferential components as discussed earlier. We need to recall Equation 7.42 and note that dr/dt = 0. This leaves us with   dθ 2   d 2θ a =  −r    er +  r 2   dt    dt   eθ  (7.54) Here we need to note that the first term is negative because it acts in the negative radial direction. Now, from our previous definitions we can rewrite Equation 7.54 as a = (−rω2 )er + (rα )eθ (7.55) We can define the normal and tangential components of acceleration as a n = −rω2er and a t = rα eθ (7.56) where an is the normal (centrifugal) acceleration at is the tangential acceleration We can differentiate Equation 7.53 to obtain the more general result a= dr dv dω = × rP + ω P dt dt dt (7.57) If we insert Equations 7.48 and 7.53 into Equation 7.57, we get the general vector form for the acceleration of a rigid body rotating in an inertial coordinate system. a = α × rP + ω × (ω × rP ) (7.58) To move closer toward a more general treatment, we shall now derive the kinematic equations for plane motion of a rigid body using a translating coordinate system (the body is free to rotate). We can break down any planar motion of the rigid body into a translation © 2014 by Taylor & Francis Group, LLC Ballistics: Theory and Design of Guns and Ammunition 216 y΄ rA = Position vector of point A rB = Position vector of point B rB/A = Relative position vector of point B with respect to point A rB/A rB vA = Velocity of point A x΄ vB = Velocity of point B A vB/A = Relative velocity of point B Translating coordinate rA with respect to point A system (translates aA = Acceleration of point A with the body) x O aB = Acceleration of point B Inertial coordinate system aB/A = Acceleration of point B with (fixed in space) respect to point A y B FIGURE 7.15 Definition of vectors associated with rigid body translational and rotational motion. and a rotation about some point. Let us choose point A in Figure 7.15 to be a location about which the body rotates. Equation 7.43 is still valid, but dr/dt no longer equals zero. Thus, from Equation 7.44, we have vB = drB drA drB/A = + dt dt dt (7.59) Not only is drB/A/dt ≠ 0 but also, since we chose point A as one about which rotation will take place, drA/dt is a pure translation and drB/A/dt is a pure rotation about point A. Thus, we can write Equation 7.59 in terms of the velocities as v B = v A + v B/A (7.60) We saw earlier that for a pure rotation, we can write the velocity in the form of Equation 7.53. Thus, we have v B/A = ω × rB/A (7.61) If we substitute Equation 7.61 into Equation 7.60, we obtain the vector equation for planar motion of a rigid body in which our coordinate system translates with a point in the body but does not rotate with the body. v B = v A + ω × rB/A (7.62) To obtain the acceleration of point B, we need to differentiate Equation 7.60, thus aB = dv B dv A dv B/A = + dt dt dt (7.63) Let us examine this equation term by term. The first term is straightforward, showing that the acceleration of the translation is simply the linear acceleration of our chosen reference point. dv A = aA dt © 2014 by Taylor & Francis Group, LLC (7.64) Dynamics Review 217 The second term is differentiated as follows: dv B/A d dω dr = (ω × rB/A ) = × rB/A + ω × B/A dt dt dt dt (7.65) We again need to call upon Equations 7.48 and 7.53 to get Equation 7.65 into a more general form. dv B/A = α × rB/A + ω × (ω × rB/A ) dt (7.66) Inserting Equations 7.64 and 7.66 into Equation 7.63 yields the kinematic equation for the acceleration of a rigid body in a coordinate system that translates with the body. The coordinate system in this case moves with the body but does not rotate allowing the body to rotate relative to the moving coordinate system. a B = a A + α × rB/A + ω × (ω × rB/A ) (7.67) We shall now derive the kinematic equations for plane motion of a rigid body using a translating and rotating coordinate system (the body is capable of movement in both coordinate systems). We choose point A in Figure 7.16 to be a location from which we want to measure the motion of the body. At the instant considered, point A has a position, rA, a velocity, vA, and an acceleration a A, while the x–y axes (and the body) are rotating with angular velocity, ω, and accelerating with angular acceleration, α. Equation 7.43 is still valid for determination of the position vectors. For the velocity of point B, we shall use the form vB = v A + drB/A dt (7.68) If we examine our earlier derivations under the curvilinear motion and replace er with i and eθ with j (i and j represent the unit vectors in our x–y coordinate system), we obtain the following relations: di dθ = j = ωj dt dt (7.69) dj dθ = (−i) = −ωi dt dt (7.70) y΄ y B rB /A x΄ rB A rA Coordinate system (translates and rotates independent the body) x O Inertial coordinate system (fixed in space) FIGURE 7.16 Definition of vectors associated with rigid body translational and rotational motion including a rotating local coordinate system. © 2014 by Taylor & Francis Group, LLC Ballistics: Theory and Design of Guns and Ammunition 218 y B rB/A x A ω FIGURE 7.17 Body rotating in moving coordinate system. Using the definition of the cross product and noting they are orthogonal, Equations 7.69 and 7.70 can be rewritten as di = ω×i dt (7.71) dj = ω× j dt (7.72) If we look at the body and our moving coordinate system as shown in Figure 7.17, we see that if the body translates and rotates relative to our x–y axes we can write, in light of Equation 7.62 drB/A = (v B/A )xyz + ω × rB/A dt (7.73) Substituting this into Equation 7.68 yields the general relation for the velocity of a point in a rigid body as seen from an arbitrary coordinate system. v B = v A + (v B/A )xyz + ω × rB/A (7.74) To obtain the acceleration of our point B, we need to differentiate Equation 7.74 with respect to time, thus aB = drB/A dv A d(v B/A )xyz dω + + × rB/A + ω × dt dt dt dt (7.75) Let us again move term by term through Equation 7.75. The first term is dv A = aA dt (7.76) Since our point A does not rotate, the acceleration of its translation is simply this linear acceleration. © 2014 by Taylor & Francis Group, LLC Dynamics Review 219 The second term is differentiated by first breaking up vB/A into its components along the x and y axes of our moving frame. Since we are looking only at motion in the x–y plane, the z component is nonexistent. (v B/A )xyz = (v B/A )x i + (v B/A )y j (7.77) This allows us to write the second term as d(vB/A ) xyz d(vB/A ) x d(vB/A ) y di dj = i+ j + (vB/A ) x + (vB/A ) y dt dt dt dt dt (7.78) The first pair of terms in Equation 7.78 are the acceleration components of point B relative to point A as seen by an observer moving with the coordinate system at point A. The second pair of terms of Equation 7.78 can be rewritten as the cross product of the angular velocity of the x–y coordinate system and the velocity vector of point B relative to point A. So we can write d(v B/A )xyz = (a B/A )xyz + ω × (v B/A )xyz dt (7.79) Returning now to Equation 7.75, in the third term, we simply rewrite the term dω/dt as α. Finally, for the last term of Equation 7.75, we use Equation 7.73 to obtain a B = a A + (a B/A )xyz + ω × (v B/A )xyz + α × rB/A + ω × (v B/A )xyz + ω × (ω × rB/A ) (7.80) After a slight rearrangement and combination of like terms, we have the general kinematic equation for acceleration of point B a B = a A + α × rB/A + ω × (ω × rB/A ) + 2ω × (v B/A )xyz + (a B/A )xyz (7.81) It is important to review each of the terms in Equations 7.74 and 7.81. First, let us review the generalized velocity and acceleration equations we derived. v B = v A + (v B/A )xyz + ω × rB/A (7.74) a B = a A + α × rB/A + ω × (ω × rB/A ) + 2ω × (v B/A )xyz + (a B/A )xyz (7.81) The terms in these equations have meanings as tabulated in Table 7.1. As one can imagine the addition of the third dimension in these equations adds significant complexity to the expressions although the basic principles remain the same. The interested reader is referred to Ref. [1] or any similar text on dynamics to familiarize themselves with the three-dimensional uses of these equations. © 2014 by Taylor & Francis Group, LLC Ballistics: Theory and Design of Guns and Ammunition 220 TABLE 7.1 Vector Terms Used in Equations 7.74 and 7.81 Variable rB/A vA vB (vB/A)xyz aA aB (aB/A)xyz ω α Definition Relative position vector of point B with respect to point A Velocity of point A in the inertial coordinate system Velocity of point B in the inertial coordinate system Relative velocity of point B with respect to point A in the xyz coordinate system Acceleration of point A in the inertial coordinate system Acceleration of point B in the inertial coordinate system Acceleration of point B with respect to point A in the xyz coordinate system Angular velocity of the xyz coordinate system measured from the inertial coordinate system Angular acceleration of the xyz coordinate system measured from the inertial coordinate system Problem 1 A 155 mm projectile is in flight at its maximum ordinate. At this instant in time, the nose of the projectile is pointing along (and spinning about) the unit vector: x = (0.998e1 + 0.030e2 + 0.056e3 ) The projectile velocity vector is  ft  V = (1199e1 + 0e2 + 49e3 )   s In both of these cases, e1, e2, and e3 are unit vectors in the x, y, and z planes, respectively. Also at this location the air density, spin rate, and projectile mass are as follows:  lbm   rev  ρ = 0.052  3  , p = ω = 150  , and m = 100 [lbm]  ft   s  The projectile characteristics are assumed to be CD = 0.29 CMα = 3.0 (CMq + CM αɺ ) = −10.2 CLα = 2.12 (CNq + CN αɺ ) = 0.002 CNpα = −0.010 = −10.2 CMpα = 0.51 Clp = −0.015 Please answer the following questions: 1. Draw the situation. 2. Determine the drag force vector. Answer: FD = (−68.39e1 − 2.80e3) [lbf] © 2014 by Taylor & Francis Group, LLC Dynamics Review 221 3. Determine the lift force vector. Answer: FL = (−0.310e1 + 14.980e2 + 7.600e3) [lbf] 4. Determine the overturning moment vector. Answer: MM = (−0.441e1 − 5.468e2 + 10.783e3) [ft-lbf] 5. Determine the magnus moment vector. Answer: MMpα = (0.043e1 − 0.732e2 − 0.369e3 ) [ft-lbf] Reference 1. Greenwood, D.T., Principles of Dynamics, 2nd edn., Prentice Hall, Englewood Cliffs, NJ, 1988. Further Readings Beer, F.P. and Johnston, E.R., Vector Mechanics for Engineers—Statics and Dynamics, 7th edn., McGraw-Hill, New York, 2004. Colley, S.J., Vector Calculus, Prentice Hall, Upper Saddle River, NJ, 1998. Hibbeler, R.C., Engineering Mechanics—Statics and Dynamics, 7th edn., Prentice Hall, Englewood Cliffs, NJ, 1995. O’Neil, P.V., Advanced Engineering Mathematics, 2nd edn., Wadsworth Publishing Co., Belmont, CA, 1986. © 2014 by Taylor & Francis Group, LLC © 2014 by Taylor & Francis Group, LLC 8 Trajectories Now that the basics of the terminology and the dynamic equations have been presented, we shall begin to look at their uses in the form of prediction of trajectories. The aero-ballistician is usually faced with one of two problems: “If I want to hit a target at position x, to what elevation (and perhaps with how much propelling charge) do I have to elevate the weapon?” or “My weapon is elevated to elevation x and I expect muzzle velocity y—where is the projectile going to end up?” To approach this in a logical and easily understandable fashion, we shall begin with a great many simplifying assumptions, relieving these as we progress. Each section builds upon the previous one so that we recommend even seasoned veterans progress in numerical order. Initially, we will only look at the effect that gravity imposes on the projectile, a vacuum trajectory, so that even the air is removed from our area of concern, thus neutralizing the fluid mechanics for a while. As we progress, we shall add in the atmosphere but neglect dynamics, atmospheric perturbations, and earth rotation. One by one we shall continually step up the complexity until finally we shall introduce the full six degree-of-freedom (6 DOF) equations. One might initially think that these simplified models have no practical use, but are merely educational stepping stones. Nothing could be further from the truth. In many instances, some of the complications only slightly affect the solution and a ballistician is well placed to assume them away. Some of these common situations will be pointed out as they arise. 8.1 Vacuum Trajectory In this section, we will make two broad assumptions: First, that the projectile mass is concentrated at a point (which allows us to neglect body dynamics affected by mass distribution), and second, that the only force acting on the projectile is that due to the acceleration of gravity (this allows us to neglect the rather complicated fluid dynamic effects when a solid body moves through a fluid). With these assumptions, the two governing differential equations of motion are mxɺɺ = 0 (8.1) yɺɺ = − g (8.2) The solutions to these equations, found by integrating with respect to time, are x = V0t cos φ0 y = V0t sin φ0 − 1 2 gt 2 (8.3) (8.4) 223 © 2014 by Taylor & Francis Group, LLC Ballistics: Theory and Design of Guns and Ammunition 224 y V0 = 800 m/s V mg 0= ? x R = 20,000 m FIGURE 8.1 Vacuum trajectory. A sketch of this simplified trajectory is seen in Figure 8.1. Notice that unlike the generalized trajectory shown previously in Figure 6.1 the terminal point is on the line y = 0 and the entire trajectory is in the x–y plane. But from Equations 8.3 and 8.4, we can write y gt = tan φ0 − 2V0 cos φ0 x (8.5) By solving Equation 8.3 for time, t, we get t= x V0 cos φ0 (8.6) We can then write Equation 8.5 putting y in terms of x only. Therefore, y = x tan φ0 − gx 2 2V02 cos 2 φ0 (8.7) This equation is in the form of a parabola in x and y coordinates, the path the projectile will follow in a vacuum. Solving for the range, x, when y = 0 gives x(2V02 cos φ0 sin φ0 − gx) = 0 (8.8) This says that either x = 0 (the trivial solution) or x= V2 2V02 cos φ0 sin φ0 = 0 sin 2φ0 g g (8.9) Because the trajectory is a parabola, maximum range is attained at π 4 (8.10) π =1 2 (8.11) φ0 = since sin © 2014 by Taylor & Francis Group, LLC Trajectories 225 When we substitute this into Equation 8.9, the maximum range can be found to be xmax = V02 g (8.12) The maximum ordinate of the trajectory is at half the maximum range and is ymax = V02 4g (8.13) If we differentiate Equation 8.9 with respect to ϕ 0 and set this equal to 0, we can prove Equation 8.10 as shown next: dxmax 2V02 = cos 2φ0 = 0 dφ0 g (8.14) This gives the launch angle for maximum range in a vacuum as π/4. Except for the maximum range, there are two quadrant elevations (QE) that will allow a projectile to impact at a given distance. We will designate the second QE with a caret, “∧.” Its existence is due to the identity sin φ = sin(180 − φ ) (8.15) sin 2φ0 = sin 2(90° − φ0 ) (8.16) 1  gR  φˆ0 = 90° − φ0 = 90° − sin −1  2  2  V0  (8.17) Then, Thus, where x has been replaced by the range, R. The maximum ordinate is achieved when the y-component of the velocity is 0. By differentiating Equation 8.4 with respect to time and setting the result equal to 0, we get V0 sin φ0 − gts = 0 (8.18) or ts = V0 sin φ g (8.19) Substituting this into Equation 8.4 gives 2  V sin φ0  1  V0 sin φ0  1 V02 sin 2 φ0 = − y s = V0 sin φ0  0 g   2  g g g 2  © 2014 by Taylor & Francis Group, LLC (8.20) Ballistics: Theory and Design of Guns and Ammunition 226 If we note that at impact the y-coordinate is zero, we can find the time of flight to impact with Equation 8.4 0 = V0tI sin φ0 − 1 2 gtI 2 (8.21) or tI = 2V0 sin φ0 g (8.22) This is double the time to the maximum ordinate and the trajectory in a vacuum is symmetrical about this ordinate. Further evidence of the symmetry may be seen by examining the angle of fall. If we differentiate Equation 8.7 with respect to x and substitute the value of x we found at impact in Equation 8.9 in the differentiated result, we see that dy sin 2φ0 = tan φ0 − cos 2 φ0 dx I (8.23) sin 2φ0 = 2 sin φ0 cos φ0 (8.24) But Therefore, tan φI = tan φ0 − 2 sin φ0 = − tan φ0 cos φ0 (8.25) Thus, in a vacuum trajectory, the projectile impacts at the mirror image of the angle it had when it was launched. For any given launch velocity, V0, maximum range in a vacuum is achieved with an initial launch angle of 45°. To reach any range shorter than the maximum, there are two launch elevations, one greater than 45° and the other less, a high angle and a low angle of fire. The trajectory envelope is a curve that bounds all possible trajectories that attempt to reach all ranges from zero to the maximum range possible for the given launch velocity [1]. We shall now mathematically describe this curve. We know from Equation 8.7 that y = x tan φ0 − gx 2 2V02 cos 2 φ0 (8.7) y = x tan φ0 − gx 2 sec 2 φ0 2V02 (8.26) This can also be written as © 2014 by Taylor & Francis Group, LLC Trajectories 227 If we make use of the trigonometric identity sec2 ϕ = 1 + tan2 ϕ, we can, with substitution and manipulation, write tan 2 φ0 − 2V02 2V 2 tan φ0 + 02 y + 1 = 0 gx gx (8.27) Equation 8.27 is quadratic in ϕ 0 and as such, when solved, yields two roots that correspond to the two elevations that achieve the same range as discussed earlier. The exceptions to this are when the range is zero or the range is maximum. These conditions yield a repeated root. The other instances a repeated root occurs are whenever the trajectory touches the trajectory envelope. This occurs only once at any given elevation. If the roots of this equation are complex conjugates, the range in question cannot be achieved with the given muzzle velocity. We can solve for all of the double roots to obtain the equation of the trajectory envelope. We proceed by first completing the square in Equation 8.27 noting that 2  V2   2V 2 V2  tan φ0 − 0 tan φ0 +  0  =  tan φ0 − 0  gx gx   gx   2 2 ( ) (8.28) 2 By adding and subtracting a term, V02 gx , to Equation 8.27, we complete the square of a part of the equation and can operate on the remainder of it: 2 2  V 2   V 2  2V 2 2V 2 tan φ0 − 0 tan φ0 +  0  −  0  + 02 y + 1 = 0 gx gx  gx   gx  2 (8.29) Breaking apart Equation 8.29 into two terms and setting each equal to zero gives us from Equation 8.28 2 tan 2 φ0 − 2  V2   2V02 V2  tan φ0 +  0  =  tan φ0 − 0  = 0 gx gx   gx   (8.30) and 2  V2  2V02 y − 0  +1 = 0 2 gx  gx  (8.31) The double root in Equation 8.30 occurs when tan φ0 = V02 gxe (8.32) where xe = x on the envelope curve. We can also pursue the equation for the envelope curve more directly 2  V2  2V02 ye −  0  + 1 = 0 2 gxe  gxe  © 2014 by Taylor & Francis Group, LLC (8.33) Ballistics: Theory and Design of Guns and Ammunition 228 2000 1800 1600 Altitude (ft) 1400 1200 1000 800 600 400 200 0 0 500 1000 1500 2000 2500 Range (ft) 3000 3500 4000 4500 FIGURE 8.2 Trajectory envelope. where ye is the y-coordinate on the envelope curve. Equation 8.33 can be further rearranged to yield the final equation of the trajectory envelope: ye = 1 V02 gxe2 − 2 g 2V02 (8.34) A typical trajectory envelope is illustrated in Figure 8.2. To move to a different subject in the study of the vacuum trajectory, when the trajectory of the projectile is relatively flat, certain simplifying assumptions may be made, which allow the equations of motion to be solved with greater ease. In particular, if we rewrite Equation 8.7 as y = x tan φ0 − gx 2 sec 2 φ0 2V02 (8.35) we now take its derivative with respect to ϕ 0, we get gx dy gx 2   = x sec 2 φ0 − 2 tan φ0 sec 2 φ0 = x  1 − 2 tan φ0  sec 2 φ0 V dφ0 V0 0   (8.36) Now because sec 2 φ0 = 1 + tan 2 φ0 and if tan2 ϕ 0 ≪ 1 then sec2 ϕ 0 ≈ 1. This occurs when ϕ 0 < 5°. This is the requirement for what is commonly called the flat fire approximation to be valid. We can then translate Equations 8.35 and 8.36 into y ≈ x tan φ0 − © 2014 by Taylor & Francis Group, LLC gx 2 2V02 (8.37) Trajectories 229 and dy gx   ≈ x  1 − 2 tan φ0  dφ0  V0  Equation 8.38 can be even further simplified for short ranges if − dy ≈x dφ0 (8.38) gx tan φ0 ≪ 1 then V02 (8.39) This is sometimes known as the rigid trajectory because the trajectory appears to rotate rigidly with the elevation angle. The vertical error that arises from use of the flat fire approximation in a vacuum trajectory is εy = gx 2 tan 2 φ0 2V02 (8.40) which, as is readily seen, states that as the range or launch angle increases, the error increases. Flat fire is characteristic of the engagements experienced with high-powered, high-velocity tank cannons where initial launch angles for direct-fire ranges of several kilometers are less than 5°. Elevation changes to correct fire are measured in fractions of a degree (known as mils) as well. One mil is equal to 1/6400 of a circle. Problem 1 A target is located at 20 km. A projectile muzzle velocity is 800 m/s, assuming a vacuum trajectory, at what quadrant elevation (QE) should one set the weapon to hit the target? Answer: ϕ 0 = 158.7 [mil]. Problem 2 The enemy in the aforementioned problem is very smart and has located his unit on the reverse slope of a hill that is 3,000 m in height with its peak located 18,000 m from your firing position. Assuming that the target is at the same level as you (just behind the hill), determine a firing solution (QE, if there is one) to hit him assuming a vacuum trajectory. Answer: It can be hit.—You find the initial QE. Problem 3 The U.S. pattern 1917 (M1917) “Enfield” rifle was the most numerous rifle used by our troops in the First World War. It was an easier rifle to manufacture than the M1903 “Springfield” (even though the Springfield was officially the U.S. Army’s service rifle) and the troops liked its accuracy better. In fact, the famous Sergeant Alvin York was actually armed with an Enfield, not a Springfield as is commonly believed, when he single-handedly captured over 100 German soldiers in the Argonne Forest in 1918. The pattern 1917 used the standard M1 30-’06 cartridge in U.S. service. The bullet had a mass of 174 grains (a grain is a common unit of measure in © 2014 by Taylor & Francis Group, LLC 230 Ballistics: Theory and Design of Guns and Ammunition small arms ammunition and is defined as 1/7000 of a lbm) and a diameter of 0.308 in. This cartridge–rifle combination has a muzzle velocity of 2800 ft/s. Assuming a vacuum trajectory: 1. Determine the angle in degrees to set the sights on the rifle (i.e., the QE) if the target is level with the firer and at 200 yards range. Answer: ϕ 0 = 0.0705°. 2. If the target is at the same horizontal range but 20 yards higher, and the firer does not adjust the sights, how much higher or lower will the bullet strike? Answer: ymiss = −0.0125 [in.]. Problem 4 You are asked to create a rough safety fan for a maximum range test at Yuma Proving Ground. The test consists of a U.S. M198 155 mm howitzer firing an M549 projectile at maximum charge with rocket off. The projectile weighs 96 lbm. The muzzle velocity is 880 m/s. 1. Using a vacuum trajectory, calculate and plot the trajectory envelope for the test. Answer: R = 78,940 [m]. 2. Determine the longest time of flight of the projectile. Answer: tI = 179.4 [s]. Problem 5 It is desired to use a 105 mm battery to set up a line of illumination candles over an enemy defensive position. The enemy is positioned in depth of about 2 km. Assuming that four weapons are available and the candle array is to be placed such that the closest candle is 9 km away from the gun position (assume all guns occupy the same point in space) and there is one candle every 500 m (i.e., four candles total at 9.0, 9.5, 10.0, and 10.5 km), determine the QE for each weapon, the time set on each fuze, and the time and order of fire of each gun such that all of the expulsion events occur simultaneously at an altitude of 750 m over the heads of the enemy. Assume a muzzle velocity of 600 m/s and a projectile mass of 30 lbm. Use a vacuum trajectory for the purposes of this problem. (Note that this would NEVER be acceptable in actual practice.) Could the same result be achieved with less than 4 guns? Show why or why not with a calculation. Problem 6 The fire control problem: During the age of battleships, it was essential that a fire control computer be installed on the ships. From 1915 onwards, these were mechanical devices by which one could put in their own ship’s course and speed as well as the target ship’s estimated course and speed. These data and the expected muzzle velocity of the projectile (as well as time of flight) allowed the guns to be aimed where the enemy ship would be when the shells landed. To get a feel for the magnitude of the problem, we are going to examine it in a greatly simplified form assuming a vacuum trajectory. Given the data that follow, and assuming a vacuum trajectory, provide a firing angle off the bow, elevation and timing to fire each of four guns so that a pattern is created to hit if the target veers 10° to the port or starboard of its present course. A hit can be assumed to occur if the shell lands on the point where the enemy ship will actually be or if it lands in its “danger space.” Because of the trajectory of the shells, a hit will occur if the trajectory passes over the target and lands within 100 yards behind it—the ship creates a “shadow” or danger space. An example pattern might look like the one drawn in the figure but feel free to create your own if it meets the aforementioned criteria. The shells © 2014 by Taylor & Francis Group, LLC Trajectories 231 should all be fired at the same time. Assume your speed estimate of the target is exact. Also assume the four guns are mounted in two pairs so that you only have two azimuths to work with but the elevations can be varied independently. The target is about 400 ft long so some error in azimuth is acceptable—but the error should be quantified. Plot the impact points and target position at the time of impact. Remember that your ship is moving! The weapons are British 12” mark IX naval guns with a muzzle velocity of 2800 ft/s and a maximum elevation of 20°. 20 knots 25° 4,000 yards 15 knots 12,000 yards 8.2 Simple Air trajectory (Flat Fire) As we progress in our study of exterior ballistics, we now introduce the concept of drag by substituting air for the medium through which our point mass projectile flies. We do this so that projectile dynamics do not enter yet into the equations of motion. We are essentially still dealing with a spherical, nonrotating cannon ball. Furthermore, to simplify the mathematics, we will insist on a flat fire trajectory, with launch angles below about 6°. A flat fire trajectory is depicted in Figure 8.3. The methods and equations we will develop were used in the 1950s for direct-fire calculations over relatively short ranges [1]. We begin with Newton’s second law in an inertial reference frame and use vector representations (boldface) where appropriate. F = ma m (8.41) dV = ΣF + m g dt (8.42) V0 y Vy 0 0 Vx 0 V0 j V 0 i FIGURE 8.3 Flat fire trajectory. © 2014 by Taylor & Francis Group, LLC mg x Ballistics: Theory and Design of Guns and Ammunition 232 where m is the projectile mass V is the projectile velocity vector t is the time dv a= = Vector acceleration dt ∑ F is the vector sum of all aerodynamic forces g is the vector acceleration due to gravity The inertial reference frame allows us to neglect the Coriolis acceleration, which is the result of the earth’s rotation. Since there is no angle of yaw, the lift and drag forces due to yaw and the Magnus force due to spin are also negligibly small. These will be discussed in detail later. Thus, only the projectile drag forces (base, wave, and skin-friction) are working to slow the projectile down and gravity is pulling it toward the earth. The aerodynamic drag force acting on the projectile is then given by 1 1 FD = − ρ SCD VV = ρV 2SCD 2 2 (8.43) where CD is the drag coefficient, introduced earlier. Another coefficient in common use in ballistics is the ballistic coefficient, C, which is defined as C= m d2 (8.44) where m and d are the mass and diameter of the projectile. As a matter of convenience, we also define ρ SCD ρπ CD = (8.45) Cˆ D* = 2m 8 C This allows us to save a little energy in typing since this combination of parameters appears so often. It is known as a starred coefficient [2]. Equation 8.45 stems from the fact that S, the frontal area of the projectile, is S= π d2 4 (8.46) We can combine Equations 8.42 and 8.45 and divide by the mass to get an expression for the time rate of change of velocity (acceleration) dV 1 =− ρ SCD VV + g dt 2m (8.47) The negative sign was placed in front of the force (Equation 8.43) because the drag always opposes the velocity vector (otherwise, it is called thrust). We can separate the velocity, acceleration, and gravitational vectors into components along the coordinate axes so that they will be convenient to work with: dV  = Vxi + V y j + V z k + (Vxi + Vy j + Vz k ) dt © 2014 by Taylor & Francis Group, LLC (8.48) Trajectories 233 and g = −gj (8.49) But because we are in an inertial frame, i̇ = j̇ = k̇ = 0 and therefore dV  = Vxi + V y j + V z k dt (8.50) If we break Equation 8.47 into its components, we get three coupled, ordinary, nonlinear differential equations V x = −Cˆ D* VVx V y = −Cˆ D* VVy − g V z = −Cˆ D* VVz (8.51) (8.52) (8.53) The equation that couples Equations 8.51 through 8.53 is V = Vx2 + Vy2 + Vz2 (8.54) We can linearize these equations by making a few assumptions. First, let us assume that there is no crosswind, so Vz = 0, and if we further constrain the ratio of the vertical velocity to the horizontal velocity to |Vy/Vx| = tan ϕ < 0.1, then V and Vx are within 0.5% of each other, and we have constrained the launch and fall angles to be less than 5.7°, the angles introduced in the preceding section for the flat fire approximation. So with the assumptions that V = Vx and Vz = 0, we can develop Equations 8.51 through 8.53 into V x = −Cˆ D* Vx2 (8.55) V y = −Cˆ D* VxVy − g (8.56) V z = 0 (8.57) These differential equations use time as the independent variable. It is often convenient to use distance as the independent variable. By making a common transformation of variables to allow distance along the trajectory to be the independent variable instead of time, we can improve our ability to work with these expressions. Performing the transformation results in equations of the form VxVx′ = −Cˆ D* Vx2 (8.58) VxVy′ = −Cˆ D* VxVy − g (8.59) where the prime denotes differentiation with respect to distance. © 2014 by Taylor & Francis Group, LLC Ballistics: Theory and Design of Guns and Ammunition 234 By dividing both equations by Vx we obtain Equations 8.60 and 8.61 that use the downrange distance, x, as the independent variable. For these equations, an analytic solution does exist: Vx′ = −Cˆ D* Vx (8.60) g Vy′ = −Cˆ D* Vy − Vx (8.61) Equation 8.60 can be integrated by separation of variables as x   Vx = Vx0 exp  − Cˆ D* dx1    0   ∫ (8.62) In this equation and future equations, we use a variable xi or ti as a dummy variable of integration. Equation 8.61 can also be solved by quadrature methods since it is of the form dVy ˆ * g + CDVy = − Vx dx (8.63) Equation 8.61 can be solved for Vy for initial conditions at x = 0, t = 0, and Vy = Vy0 as  x  ˆ *  Vy = exp − C D dx1  Vy 0 −    0  ∫ x ∫ 0    x2  g  ˆ  *    V  exp  C D dx1  dx2   x  0  ∫ (8.64) If we take the ratio of the x and y velocity components, we can obtain the relation for the angle the velocity vector makes with the horizontal. This can be shown to be  1 tan φ =  tan φ0 −  Vx  x ∫ 0  g V  x   x2    Cˆ *D dx1  dx2  exp      0   ∫ (8.65) where ϕ 0 is the initial launch angle. To complete our study of the flat fire trajectory, we need to find the elements of it, i.e., the x and y values along it, from launch to termination. To do this, we must integrate over time the velocities we have found in Equations 8.62 and 8.64, which we had earlier transformed into distance variables. We know by definition t ∫ t ∫ y = Vydt and x = Vxdt 0 © 2014 by Taylor & Francis Group, LLC 0 (8.66) Trajectories 235 Substituting Equations 8.62 and 8.64 into each of the equations of 8.66 in turn and performing the integrations, we can show that with the initial conditions of x = 0, t = 0, and y = y0 x  x   x2   g  * ˆ ˆ *       y = t exp C Ddx1 Vy0 −   exp C D dx1 dx2 + y0       Vx  0 0  0   ∫ ∫ ∫ (8.67) Now we can find t from Vx = dx/dt. Separating the variables and substituting Equation 8.62 for Vx, we get  x2  exp  Cˆ D* dx1 dx2   0 0  x t= 1 Vx0 ∫ ∫ (8.68) Through a somewhat tedious set of algebraic substitutions and manipulations that are contained in Ref. [1], we can arrive at our desired equation in x and y; the launch angle, ϕ 0; the dummy range variables x1, x2, and x3; the initial launch velocity, Vx0; the initial ordinate, y0; and the drag coefficient, ĈD* : y = y0 + x tan φ0 −  gx 2  2 2Vx20  x 2  x x3 ∫∫ 0 0   x2  exp  2 Cˆ *D dx1  dx2dx3       0  ∫ (8.69) The disadvantage of these equations is that the variation of drag coefficient has to be simple to evaluate the integrals. Since the drag coefficient does not vary in a simple manner with Mach number, this makes the analytic solutions inaccurate and difficult to accomplish. Figure 8.4 0.48 0.46 0.44 0.42 0.40 0.38 0.36 0.34 0.32 0.30 0.28 0.26 0.24 0.22 0.20 0.18 0.16 0.14 0.12 0.10 0.08 0.06 0.04 0.02 0.00 Plot area 0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.4 1.6 1.8 2.0 2.2 2.4 2.6 2.8 3.0 3.2 3.4 3.6 3.8 4.0 4.2 4.4 4.6 4.8 5.0 5.2 5.4 5.6 5.8 6.0 6.2 6.4 6.6 6.8 7.0 7.2 7.4 7.6 7.8 Zero yaw drag coefficient Zero yaw drag Mach FIGURE 8.4 Drag coefficient versus Mach number for a typical projectile. © 2014 by Taylor & Francis Group, LLC Ballistics: Theory and Design of Guns and Ammunition 236 depicts a typical drag curve that varies with Mach number. One can see from this figure that there is no simple analytic solution to this variation. With computer power nowadays we usually solve or approximate the exact solutions numerically, doing the quadratures by breaking the area under the curve into quadrilaterals and summing the areas. To integrate these equations analytically, we will examine three forms of the drag coefficient: 1. Constant CD that is useful for the subsonic flight regime, M < 1 2. CD inversely proportional to the Mach number that is characteristic of the highsupersonic flight regime, M ≫ 1 3. CD inversely proportional to the square root of the Mach number that is useful in the low-supersonic flight regime, M ≥ 1 First, we will examine the case of a constant drag coefficient. If we examine Figure 8.4, we can see that this would be a useful approximation for our projectile behavior if the launch velocity was, say, between Mach 0.8 and 0. We assume that the drag force varies with the square of the velocity (the drag coefficient was the drag force divided by the 2 1 dynamic pressure, 2 ρV ) and we set the drag coefficient equal to a constant, K1. We shall use terminology consistent with Ref. [1] so that ĈD* = ρS ρS CD = K 1 = k1 2m 2m (8.70) which we then substitute in Equation 8.62: x    Vx = Vx0 exp −k1 dx1  = Vx0 exp(− k1x)   0   ∫ (8.71) We can find t by substituting Equation 8.70 into Equation 8.68 to give   x2 1 exp  k1dx1  dx2 = Vx0   0  0 x 1 t= Vx0 ∫ ∫ x ∫ exp(k x )dx 1 2 2 (8.72) 0 or t= 1 1 (exp[k1x] − exp[0]) = (exp[k1x] − 1) Vx0 k1 Vx0 k1 (8.73) Noting that from Equation 8.62  x  V exp  − Cˆ *D dx1  = x   Vx0  0  ∫ © 2014 by Taylor & Francis Group, LLC x  V x and exp  Cˆ D* dx1  = 0   Vx 0  ∫ (8.74) Trajectories 237 and also noting that (Vy0 Vx0 ) = tan φ0 , we can show through manipulation [1] that  gt Vy = Vx  tan φ0 − V x0   Vx0 k1t   1+  2   (8.75) To find the angle of fall, ϕ, as a function of range, x, and the instantaneous velocity at x, Vx, we solve Equation 8.71 for k1: k1 = 1  Vx0  ln x  Vx  (8.76) We now substitute Equation 8.76 into Equation 8.73, for t. Taking the result and recalling that tan ϕ = Vy/Vx for any x, we use this new equation for t and transform Equation 8.65 into tan φ = tan φ0 − gt Vx0  Vx0 t 1  Vx0   ln 1 +  2 x  Vx    (8.77) Finally, to find the altitude, y, at any point along the trajectory as a function of the range and the velocity at that range, we transform Equation 8.65 with the constant drag coefficient, k1, use the new equation for t that we derived earlier, and after manipulation arrive at  g x 1 y = y0 + x tan φ0 −  2  Vx0  Vx0 ln    Vx         2 2  1  Vx  Vx    Vx   0 − 1 +  0 − 1 − ln  0  Vx    Vx  2  Vx     (8.78) A constant drag coefficient is useful when analyzing low-subsonic projectiles since most of them have nearly constant drag coefficients. Also, projectiles at hypersonic speeds (usually described as a Mach number greater than five) can be analyzed with this assumption (look again at Figure 8.4). Essentially, we are linearizing the problem when we do this. Our next effort will be to examine a nonconstant drag coefficient, one varying as the inverse of the Mach number. In this case, we assume that the drag force varies linearly with the velocity (because the drag coefficient is the drag force divided by the dynamic pressure, 12 ρV 2 , and when we divide by the Mach number, we essentially divide by the velocity times a constant). Now we set the drag coefficient equal to K 2/M, then K2 M (8.79) ρS ρS K2 CD = 2m 2m M (8.80) CD = and ĈD* = © 2014 by Taylor & Francis Group, LLC Ballistics: Theory and Design of Guns and Ammunition 238 Recall that the Mach number is V/a where a is the speed of sound in air. Then for our flat fire approximation, we can define a constant k2 such that ρS K2a 2m (8.81) ρ S K 2 a k2 = 2m Vx Vx (8.82) k2 = Then, ĈD* = From Equations 8.55 and 8.56, we see that Vɺ x = −k 2Vx (8.83) Vɺ y = −k 2Vy − g (8.84) Vx′ = −Cˆ D* Vx (8.85) and Also from Equation 8.60, we see that Using these three equations and proceeding in the same fashion as we did with the constant CD, we can derive equations for the x- and y-velocities; the time of flight to any range, x; the angle of fall, ϕ, and the trajectory ordinate at any range. These equations are (details in Ref. [1]) Vx = Vx0 exp(−k 2t) (in terms of t) g g  Vy =  Vy0 +  exp(−k 2t) − k2 k2   (in terms of t) x  Vx0  ln Vx0  Vx  t=  Vx  1−   Vx0  gx tan φ = tan φ0 + 2 Vx0 © 2014 by Taylor & Francis Group, LLC Vx0   1− V x   1 − Vx  Vx0  (8.86) (8.87) (8.88)       (8.89) Trajectories 239   gt 2 y = y0 + x tan φ0 −   2 ln Vx0  Vx  Vx      1− V    x0   1−    Vx    ln  0       Vx   (8.90) These relations, for CD proportional to 1/M, are useful in the analysis of high-supersonic projectiles such as kinetic energy armor penetrators where 2.5 < M < ∼5. For the case where the drag coefficient varies as the M , we assume that the drag varies with velocity to the 3/2 power and we set the drag coefficient equal to K 3/ M , then ρS K3 Cˆ D* = 2m M (8.91) Since M= Vx a (8.92) we can define a new constant as ρS K3 a 2m k3 = (8.93) We can then write ρS a k Cˆ D* = K3 = 3 2m Vx Vx (8.94) Proceeding as we did in the earlier two cases, we can derive Vx, Vy, t, ϕ, and the ordinate, y. The details of the derivations are again available in Ref. [1]: Vx = Vy = − (k 3 Vx0 t + 2 ) ) ( t= x Vx0 (8.95) (in terms of t) 4Vy0  k 32Vx0 t 3  + 2k 3 Vx0 t 2 + 4t  +  Vx0 t + 2  3  k 3 Vx0 t + 2 g 3 (k 4Vx0 Vx0 Vx ) 2 (in terms of t) (8.96) (8.97) gt Vx0  1  Vx  Vx0 + 1     0 + Vx  3  Vx   (8.98) y = y0 + x tan φ0 − 1 2 1  Vx   gt   1 + 2   2 Vx0    3  (8.99) tan φ = tan φ0 − © 2014 by Taylor & Francis Group, LLC 240 Ballistics: Theory and Design of Guns and Ammunition These last Equations 8.95 through 8.99 are useful for flight in the low- to moderatesupersonic regime, 1 < M < 2.5. In summary, we have derived the equations of motion assuming a flat fire trajectory. We use them when the angle of departure and angle of fall are both below 5.7°. We have solved them with three drag assumptions: 1. A constant drag coefficient that is useful in the subsonic and hypersonic regimes and can be used over short distances in all Mach regimes. 2. A drag coefficient inversely proportional to the Mach number that is useful in the high-supersonic regime. 3. A drag coefficient inversely proportional to the square root of the Mach number that is useful in the low-supersonic flight regime. Problem 7 The French infantry rifle model 1886 called the Lebel was their standard weapon from 1886 into First World War and even saw limited use in the Second World War. You can see this 51 in. long monster in any movie involving the French Foreign Legion. It used an 8 mm cartridge called the balle D with a bullet mass of 198 grains and a diameter of 0.319 in. This cartridge–rifle combination has a muzzle velocity of 2296 ft/s. Assuming flat fire with K 3 = 0.5 and using standard sea level met data (ρ = 0.0751 lbm/ft3, a = 1120 ft/s) 1. Create a table containing range (yards), impact velocity (ft/s), time of flight (s), initial QE angle (min), and angle at impact (min) in 200 yard increments out to 1000 yards. 2. If an infantryman is looking at a target at 2000 yards, what angle will the sight have relative to the tube assuming they used standard met in the design? Answer: About 10.3°. 3. Comment on the validity of this method with respect to (2). Problem 8 British 0.303 in. ball ammunition is to be fired in an Mk.1 Maxim machine gun. The bullets’ mass is 175 grains. When used in this weapon, it has a muzzle velocity of 1820 ft/s. Assuming flat fire with K3 = 0.5 and using standard sea level met data (ρ = 0.0751 lbm/ft3, a = 1120 ft/s) 1. Create a table containing range (yards), impact velocity (ft/s), time of flight (s), initial QE angle (min), and angle at impact (min) in 200 yard increments out to 1000 yards. 2. The weapon was used by British units assigned to bolster the Italians in the Alps during the First World War (Italy came in on the Allied side because they wanted the Tyrol region from Austria more than they wanted the Nice region from France). At an altitude of 3000 ft, how much higher or lower will a bullet fired from this weapon impact a level target if the sights are set using the sea level conditions given earlier and the target is at 600 yards? At this altitude assume the density and temperature of the atmosphere are ρ = 0.0551 lbm/ft3 and T = 20°F. Answer: y = 3.078 [ft] (too high). © 2014 by Taylor & Francis Group, LLC Trajectories 241 Problem 9 The main armament in the Italian M13-40 during the Second World War was a 47 mm/ 32-caliber weapon designed and built by the Ansaldo Arms company. The most effective antitank projectile it carried was an APBC (Armor-Piercing, Ballistic Capped) round, which had a muzzle velocity of 2060 ft/s. With this particular projectile–weapon combination, the assumption of constant drag coefficient seems to yield reasonable results. The k1 value for this case is 0.00025 [1/m]. Using the flat fire, point mass trajectory create a table of range (yards), velocity (ft/s), initial QE (min), and impact QE (min) out to 1000 yards in 200 yard increments. Problem 10 A U.S. 37 mm projectile is fired with a muzzle velocity of 2600 [ft/s]. The projectile weighs 1.61 lbm. Assuming K 2 = 0.841 [unitless] and using standard sea level met data (ρ = 0.0751 lbm/ft3, a = 1120 ft/s, R = 1716 [ft · lbf/slug · R]) 1. Determine the drag coefficient CD and drag force on the projectile if the projectile is fired in still air. Answer: FD = 33.04 [lbf]. 2. Create a table containing range (yards), impact velocity (ft/s), time of flight (s), initial QE angle (min), and angle at impact (min) in 100 yard increments out to 800 yards. 3. If this weapon is used at an increased altitude and assuming the density and temperature of the atmosphere are ρ = 0.060 lbm/ft3 and T = 30°F, how much higher or lower will the weapon have to be aimed to hit a target at 800 yards? Answer: The weapon must be aimed 0.28 mil or 0.98 min lower. Problem 11 You are asked to create a rough safety fan for at maximum range test at Yuma Proving Ground. The test consists of an experimental 155 mm howitzer projectile at a severe overpressure charge. The projectile weighs 106 lbm. The muzzle velocity is 980 m/s. a. Using a vacuum trajectory calculate and plot the trajectory envelope for the test. b. Determine the longest time of flight of the projectile. If we assume an average constant drag coefficient (on the way up) of CD = 0.5/M and the projectile were fired vertically, what would the maximum ordinate be? Assume standard sea level met data (ρ = 0.0751 lbm/ft3, a = 1120 ft/s, R = 1716 [ft · lbf/slug · R]) Hint: Look at the flat fire assumptions and re-derive the equation of motion assuming there never is an x-velocity. Problem 12 The drag of a sphere below Mach number of 0.5 is well approximated by CD = AM2 + BM + C, where A = 0.0262, B = 0.0456 and C = 0.4666, and M is the Mach number. If we would like to analyze the motion of a cannon ball fired from a Demi-Culverin circa 1646, which fired a spherical shot weighing 9 lbm and of 4.5 in. diameter. Please develop the equation of motion for Vx (only) as a function of downrange distance based on the flat fire assumptions. Then using standard sea level met data (ρ = 0.0751 lbm/ft3, a = 1120 ft/s, R = 1716[ft ⋅ lbf/slug ⋅ R ]), determine Vx at a range of 1800 yards (the reported maximum range of the weapon). Assume a muzzle velocity of 550 ft/s. © 2014 by Taylor & Francis Group, LLC 242 Ballistics: Theory and Design of Guns and Ammunition Problem 13 A French 240 mm projectile is fired from a model 1873 cannon with a muzzle velocity of 440 [m/s]. The projectile mass is 144 kg. Assuming K1 = 0.55 [unitless] and using standard sea level met data (ρ = 0.0751 lbm/ft3, a = 1120 ft/s, R = 1716 [ft · lbf/slug · R]) a. Create a table containing range (m), impact velocity (m/s), time of flight (s), initial quadrant elevation angle (degrees), and angle at impact (degrees) in 600 m increments out to 7800 m. b. Experiments were conducted in France in 1884 to determine the effect of rotating and locations on the range of this projectile. These experiments were reported on by Breger in Notes on the Construction of Ordnance No. 27, 10 June 1884. The farther rearward the rotating band was placed, the greater the projectile yaw at muzzle. The same QE that provided a range of 7800 m actually only made 7688 m with a (assume constant) yaw angle of 4°. Assume that the 7800 m range was achieved with zero initial yaw. With this information calculate the yaw drag coefficient, CDδ 2. Problem 14 It is desired to develop a close protection system using a 0.50 caliber machine gun. The muzzle velocity of the weapon is 2950 ft/s. The projectile is an M8 bullet with a diameter of 0.50 in. and a mass of 0.09257 lbm. Since the projectile interception mission is to occur relatively close to the firing platform, we can assume the projectile behaves according to the flat fire assumption with a constant drag coefficient of K1 = 0.45 [unitless]. Assuming the flat fire assumption is valid for the trajectory and the intercept point is to be 20 ft above the ground, develop a firing table for the intercept out to 300 yards in 50 yard increments. Assume standard sea level met data (ρ = 0.0751 lbm/ft3, a = 1,120 ft/s, R = 1716 [ft · lbf/slug · R]) Create a table containing range (yards), impact velocity (ft/s), time of flight (s), and initial quadrant elevation angle (degrees) in 50 yard increments out to 300 yards. Comment on the accuracy of your answer. Problem 15 If we were to assume that the flat fire conditions were to hold in a coordinate system that were elevated to an angle θ, re-derive the differential equations in time and space coordinates— DO NOT solve them, just put them in the form of Equations 8.60 and 8.61. Also write the transformation equations between the normal and slant velocity (Vn and Vs) and Vx and Vy. Problem 16 One of the problems with hit-to-kill close protection systems is the ability to accurately point the weapon and fire in a timely fashion at a very small target. Assume that we must impact a sphere 4 in. in diameter at the ranges developed in problem 14. For each range, assuming perfect timing as well as perfect projectile tracking, determine the allowable tolerance in QE to impact the target. Problem 17 Let us now assume that the pointing against the incoming round in problem 14 is absolutely perfect. Determine the tolerance in lock time (officially the time from pulling the trigger to weapon firing—but we will assume it is to muzzle exit) to hit the target at the conditions of problem 14. Assume that the incoming projectile is moving at 300 ft/s. How does this change if the velocity estimate is ±20 ft/s? © 2014 by Taylor & Francis Group, LLC Trajectories 243 Problem 18 The main armament of the last pre-war U.S. Heavy Cruisers (known as the “tin-clads”) was an 8 in./55 cal weapon. The effective range of this weapon was 30,000 yards at an elevation of 40°43′. During the Second World War, there were many night actions in the Pacific where these weapons were used at an extremely short range (less than 10,000 yards). You are asked to create a firing table for this weapon at the short ranges. The muzzle velocity of the weapon/projectile/propellant combination is 2500 ft/s. The projectile is an APC (Armor Piercing, Capped) with a diameter of 8 in. and a mass of 335 lbm. Since the range is short, we can assume the projectile behaves according to the flat fire assumption with a drag coefficient inversely proportional to the Mach number of K 2 = 0.62 [unitless] (note that this is not really a great fit for this projectile). Assuming the flat fire assumption is valid for the trajectory, develop a firing table for the system to 10,000 yards in 1,000 yard increments. Assume standard sea level met data (ρ = 0.0751 lbm/ft3, a = 1,120 ft/s, R = 1716 [ft · lbf/slug · R]). Create a table containing range (yards), impact velocity (ft/s), time of flight (s), and initial quadrant elevation angle (degrees) in 1,000 yard increments out to 10,000 yards. Problem 19 The analog fire control computers installed on board ships during the Second World War were amazing devices. The inputs required were course and speed of the firing ship, estimated range to the target and course and speed of the target. Inaccuracies in the target course and speed estimates were compensated for by generating a “ladder”; this was a shell pattern that was a linear array using as many guns as were available in one “salvo.” The ideal result was that the target would end up directly in the middle of the “salvo” and be “straddled.” Because of the relatively flat projection of the shells, being straddled usually guaranteed that the target was hit by at least one projectile. The shorter the range to a target, the larger the “danger space” offered and the better the chance of a hit. In this problem you assume the role of the fire control computer. The weapons available are nine 8”/55 caliber guns with the ballistic performance from problem 18. Your ship is moving due north at 30 knots (one knot is one nautical mile per hour or 2000 yards per 3600 s). At the instant of fire, the enemy ship is dead ahead of your ship traveling at 35 knots on course 090 (see the figure) at a range of 8000 yards. Ignore the effect of the launch platform motion on the drag (only) of the projectile. ± 30º enemy ship 8,000 yds θ your ship a. Determine the firing solution assuming both ships continue straight ahead (QE and relative angle to the bow of your ship) for one shell to impact the enemy. (Because of the flat trajectory, it is good to aim so the shell falls a little behind the enemy ship)—Hint: Remember the projectiles are leaving from a ship that is moving! b. Perform the same calculation if the enemy turns 30° toward the firing ship and 30° away from the firing ship. © 2014 by Taylor & Francis Group, LLC Ballistics: Theory and Design of Guns and Ammunition 244 c. Using a method of your choosing, examine the sensitivity of the fire control problem to (i) an incorrect firing ship speed, (ii) an incorrect target ship course, and (iii) an incorrect target ship speed. Quantify their relative importance. Problem 20 The U.S. 7.62 mm Ball M80 (projectile diameter = 0.308 in., mass m = 147 grains) is fired in a test range. Based on data given next, estimate the coefficient CD. Assume the projectile is fired with a muzzle velocity of 2810 ft/s, under standard sea level met conditions (ρ = 0.0751 lbm/ft3, a = 1120 ft/s). Justify your answer by explaining why you chose the appropriate drag model. Validate your answer with an appropriate calculation. Range (yards) V0 (ft/s) Vx (ft/s) 2810 2810 2810 1960 1765 1580 400 500 600 Problem 21 Many times all of the data we need for a projectile are not provided to us and we have to extract the information from different sources. You are given the following information about a British 2 pounder projectile [3]: Projectile diameter: 40 mm Projectile weight: 2.375 lbm Muzzle velocity: 2600 ft/s Armor penetration as a function of distance Distance (yards) thickness Perforated (mm) 100 500 1000 1500 55 47 37 27 In terminal ballistics, we will find that, based on some work by Zener and Holloman in 1942, the penetration of this type of projectile is proportional to the velocity as follows: V ∼t d m where t is the target thickness d is the projectile diameter m is the projectile mass With only this information at your disposal a. Determine the best drag model for this projectile b. Generate the proper coefficient from the data c. Create a table of range (yards), velocity (ft/s), time of flight (s), launch angle (minutes), impact angle (minutes) if the projectile is fired with no wind at each position Please justify your answer. © 2014 by Taylor & Francis Group, LLC Trajectories 245 Problem 22 A Hornady 0.308” diameter 208 grain Amax bullet is to be analyzed. Based on data collected by Litz [4], it has the following drag characteristics: V [ft/s] 1500 2000 2500 3000 CD 0.354 0.306 0.274 0.250 If it is fired with a muzzle velocity of 2780 ft/s, assuming standard sea level met data (ρ = 0.0751 lbm/ft3, a = 1120 ft/s) a. Determine a proper drag model to use over this range. Justify your answer by comparing numbers from all three drag models that were introduced. b. Assuming K3 = 0.409, create a table of range (yards), velocity (ft/s), time of flight (s), launch angle (min), impact angle (min), and drift (yards) if the projectile is fired with no wind every 100 yards out to 600 yards. c. Determine the deflection in inches assuming the projectile experiences a headwind of 20 ft/s for the entire flight. Tabulate every 100 yards out to 600 yards. d. Determine the deflection in inches assuming the projectile experiences a crosswind (left or right—your choice) of 20 ft/s for the entire flight. Tabulate every 100 yards out to 600 yards. Problem 23 Many precision shooters develop specific propellant loads for competition shooting for a given propellant type, projectile, barrel, cartridge and primer combination. The strategy for optimizing a propellant loading is to incrementally increase the propelling charge with the weapon aimed at the same location and determine the charge at which the projectiles “group” the most. This method, called a Creighton Audette ladder test, finds the so-called “sweet spot” of the gun. The projectiles “group” due to a combination of barrel motion owing to recoil and vibration. To illustrate this point, let us assume that we have a perfectly rigid (i.e., no vibration) 22 in. barrel that fires the projectile of problem 22. Assume a range of 300 yards. Assume that the projectile velocity during barrel transit can be approximated as ½ of the muzzle velocity. Also assume that the recoil is fairly constant and causes a constant upward rotation of the muzzle of 3.3 rad/s. For the following 20 velocities, determine the muzzle velocity (which would correspond to a propellant load) that is in the “sweet spot.” Please plot impact height versus muzzle velocity. Hint: The muzzle rise will affect the initial launch angle but the flatter trajectory will compensate to some degree. Although not done in practice, for this calculation, start at the high muzzle velocity and track the bullet impact heights as you lower the velocity. © 2014 by Taylor & Francis Group, LLC Ballistics: Theory and Design of Guns and Ammunition 246 V0 [ft/s] 2780 2775 2770 2765 2760 2755 2750 2745 2740 2735 2730 2725 2720 2715 2710 2705 2700 2695 2690 2685 Problem 24 Normally on a fin-stabilized projectile the spin damping due to the body is much smaller than that due to the fins. A 155 mm projectile weighs 101 lbm and is designed so that it leaves the muzzle of the weapon at 600 m/s and spinning at 220 Hz. After 0.5 s fins are deployed. At this instant in time the spin rate is 197 Hz and the velocity is 500 m/s. After another 0.5 s the projectile has achieved its steady state spin rate of 12 Hz and is at a velocity of 400 m/s. The changes in polar moment associated with the fin deployment are provided next (be sure to think about which one is before and which is after). Determine the following: a. The drag coefficient of the both flight configurations b. The spin damping coefficient for the body (only) and the fins (only) c. The roll coefficient (Clδ) of the fins assuming a 1° cant Assume  lbm  m ρ = 0.076  3  a ≈ 330   I PA = 10.1[lbm ⋅ ft 2 ] I PB = 10.8 [lbm ⋅ ft 2 ] ft   s It is important in this problem, since it is pretty open ended, that you list all of your assumptions and present your answer in a manner appropriate to those assumptions. Problem 25 In a test range, a 0.50 caliber M33 ball projectile is fired at an elevation of 10° with a muzzle velocity of 3013 ft/s. The initial pitch and yaw angles are 1.030° and 1.263°, respectively. The initial pitch and yaw rates are 2 rad/s nose down and 1 rad/s nose left, respectively. If the projectile has the coefficients below at this particular instant, write the acceleration vector and the angular momentum vector. © 2014 by Taylor & Francis Group, LLC Trajectories 247 Please ignore the Coriolis acceleration and assume the weapon has a right-hand twist. Projectile information: CD = 0.2938 (C Mq + CMαɺ = −5.5 ) I P = 7.85 [g-cm 2 ] CMα = 2.88 (C Nq + CNαɺ = 0.004 ) I T = 74.5 [g-cm 2 ] CLα = 2.69 CMpα = 0.05 m = 42.02 [g] CNpα = −0.01  lbm  ρ = 0.0751  3   ft   rad  p = 15, 404   s  Please supply all answers in an inertial coordinate system labeled 1, 2, 3 with 1 being along the downrange direction and 3 being to the right side. Treat all missing coefficients as equal to zero. It is very important that you DRAW the situation. Problem 26 The projectile given in problem 25 is fired off of a fast attack boat chasing down some pirates. All of the conditions in problem 25 are identical except for the mounting of the weapon and its movement. The weapon is mounted on the starboard (right) side of the boat. At the instant of firing the boat is moving at 40 knots in a straight line. The boat is rolling at 6 Hz. in a counterclockwise direction as viewed from behind. The boat is pitching into a swell so that the bow is dropping at 3 Hz. The gunner is slewing the weapon toward the bow at 1.5 rad/s while simultaneously lowering the muzzle at 2 rad/s. Assuming, at the instant of muzzle exit, the weapon is pointed directly to starboard and has the same elevation as in the test firing (i.e., 10°), find the values for the acceleration and angular momentum as was done in problem 25. Comment on the results. For a proper comparison, use the same coordinate system as in problem 25 but now with the 1 direction pointing to starboard and the 3 direction pointing to the stern of the boat. Once again the drawings are important. Problem 27 In a test range, a modified 105 mm M1 projectile is fired at an elevation of 7° with a muzzle velocity of 1022 ft/s. The initial pitch and yaw angles are 1.0° and 1.5°, respectively. The initial pitch and yaw rates are 3 rad/s nose down and 2 rad/s nose left, respectively. If the projectile has the coefficients below at this particular instant, write the acceleration vector and the angular momentum vector. Please ignore the Coriolis acceleration and assume the weapon has a right-hand twist. Projectile information: CDδ 2 = 4.20 (C (C CMα = 4.30 C Mpα = −0.892 CLα = 1.65 Clp = −0.028 CNpα = −0.55  lbm  ρ = 0.0751  3   ft  CD = 0.131 © 2014 by Taylor & Francis Group, LLC ) )≈0 Mq + CMαɺ = −8.7 Nq + C Nα I P = 0.547 [lbm-ft 2 ] I T = 5.377 [lbm-ft 2 ] m = 32.1[lbm]  rad  p = 932   s  Ballistics: Theory and Design of Guns and Ammunition 248 Please supply all answers in an inertial coordinate system labeled 1, 2, 3 with 1 being along the downrange direction and 3 being to the right side. Treat all missing coefficients as equal to zero. It is very important that you DRAW the situation. This will have a great deal of influence in obtaining the correct answer Problem 28 A modified 105 mm M1 projectile is fired downward at an angle of −45° from the horizontal from an aircraft moving horizontally at 200 knots with a muzzle velocity of 1022 ft/s. The initial pitch and yaw angles are 1.0° and 1.5°, respectively. The initial pitch and yaw rates are 3.5 rad/s nose down and 2.5 rad/s nose left, respectively. a. If the projectile is fired off the right side of the aircraft and has the coefficients below at this particular instant, write the acceleration vector and the angular momentum vector. b. Write the acceleration vector and the angular momentum vector assuming everything is the same except now the projectile is fired off the left side of the aircraft. c. Comment on the differences between parts (a) and (b). Please ignore the Coriolis acceleration and assume the weapon has a right-hand twist. Projectile information: CDδ 2 = 4.20 (C (C CMα = 4.30 C Mpα = −0.892 CLα = 1.65 Clp = −0.028 CNpα = −0.55  lbm  ρ = 0.0751  3   ft  CD = 0.131 ) )≈0 Mq + CMαɺ = −8.7 Nq + C Nα I P = 0.547 [lbm-ft 2 ] I T = 5.377 [lbm-ft 2 ] m = 32.1[lbm]  rad  p = 932   s  Please supply all answers in an inertial coordinate system labeled 1, 2, 3 with 1 being along the downrange direction and 3 being to the right side of the gun looking from the breech (be careful as this will change between parts (a) and (b)). Treat all missing coefficients as equal to zero. 8.3 Wind Effects on a Simple Air Trajectory We continue with our study of a point mass projectile model by adding a further complication to its flat fire trajectory—a crosswind or a range wind, as dynamic atmospheric phenomena. In the basic equations, we have neglected any change in air density with a change in altitude since the effect is small. We also have assumed the equations could be solved in closed form. We want to be able to solve them with winds that are both constant and variable along the flight path. © 2014 by Taylor & Francis Group, LLC Trajectories 249 We begin with a modified version of Equation 8.47 using vector notation dV ρ SCD ɶ =− V (V − W) + g dt 2m (8.100) dV ɶ (V − W) + g = −Cˆ D* V dt (8.101) or where m is the projectile mass W is the wind velocity vector V is the projectile velocity vector g is the vector acceleration due to gravity t is the time ρ is the air density dV a= is the vector acceleration dt S is the projectile reference area ρ SCD ĈD* = 2m CD is the dimensionless drag coefficient In the aforementioned equations, we have replaced the velocity vector, V, by the vector (V − W) because drag measurements are made relative to the air stream not relative to the ground. We have also replaced the scalar velocity (the speed) with ɶ =| V − W | V This is the scalar difference of the projectile and wind velocities. A diagram of the problem is shown in Figure 8.5. Height y Wy Wx Wz FD V0 j on cti fle i Range De k mg 0 z FIGURE 8.5 Coordinate system for projectile launch including wind effects. © 2014 by Taylor & Francis Group, LLC V x Ballistics: Theory and Design of Guns and Ammunition 250 We can resolve V, W, and g into components along the coordinate axes as follows: V = Vxi + Vy j + Vz k (8.102) W = Wxi + Wy j + wz k (8.103) g = −g j (8.104) Note that ɶ 2 =|V − W|⋅|V − W| V and |V| = Vx2 + Vy2 + Vz2 This leads us to |V − W|⋅|V − W|= (Vx − Wx )2 + (Vy − Wy )2 + (Vz − Wz )2 Then, ɶ = (Vx − Wx )2 + (Vy − Wy )2 + (Vz − Wz )2 V (8.105) If we insert Equations 8.102 through 8.104 into Equation 8.101, we get dV ɶ (Vx − Wx )]i + [ − Cˆ *D V ɶ (Vy − Wy ) − g]j + [− Cˆ *D V ɶ (Vz − Wz )]k = [− Cˆ *D V dt We can separate this vector equation into its three scalar components: dVx ɶ (Vx − Wx ) = −Cˆ D*V Vɺ x = dt (8.106) dVy ɶ (Vy − Wy ) − g = −Cˆ D* V Vɺ y = dt (8.107) dVz ɶ (Vz − Wz ) = −Cˆ D* V Vɺ z = dt (8.108) Equations 8.106 through 8.108 are the exact equations for a point mass trajectory of a projectile acted upon by gravity, wind, and aerodynamic drag. They are first-order, nonlinear, coupled, ordinary differential equations that are coupled through Equation 8.105. © 2014 by Taylor & Francis Group, LLC Trajectories 251 The nonlinearity, as previously discussed, creates difficulties when we attempt to solve these expressions analytically. We can only solve the exact equations using numerical methods. This will necessitate making the simplifying assumption of flat fire, which will allow us to solve them in closed form. We can alter Equation 8.105 by multiplying by the fraction (Vx − Wx)/(Vx − Wx) = 1 and then simplifying to get ɶ = (Vx − Wx ) 1 + ε y2 + ε z2 V (8.109) where εy = (Vy − Wy ) (Vx − Wx ) (8.110) εz = (Vz − Wz ) (Vx − Wx ) (8.111) and Using the binomial expansion of the form 1  1  1 + ζ = 1 + ζ − ζ 2 + ⋯ 2 8   we can operate on the radical of Equation 8.109 arriving at ɶ = (Vx − Wx ) 1 + 1 ( ε y2 + ε z2 ) − 1 ( ε y2 + ε z2 )2 + ⋯ V  2  8 (8.112) Because a projectile’s velocity is usually much greater than winds of even hurricane force, we can assume that |Wx |, |Wy |, and |Wz |≪ Vx (8.113) and ε y2 and ε z2 ≪ 1 (8.114) If we look at the first inequality of Equation 8.114 and consider the assumptions of Equation 8.113, we find that εy = Vy Wy Vy − ≈ = tan φ ≪ 1 Vx Wx Vx This was the approximation developed around a similar binomial expansion in Section 8.2. If we recall that this relation restricted us to Vy/Vx < 0.1, which, by squaring, results in © 2014 by Taylor & Francis Group, LLC Ballistics: Theory and Design of Guns and Ammunition 252 the requirement that winds be at least two orders of magnitude smaller than the velocity, Vx, and this is easily the case. The second inequality of Equation 8.114 is also satisfied if Wz and Vz are comparable in size from εz = (Vz − Wz ) (Vz − Wz ) = ≪1 (Vx − Wx ) (Vx ) All this results in Ṽ and (Vx − Wx) being within about 1% of each other. So if Ṽ ≈ (Vx − Wx), we can rewrite Equations 8.106 through 8.108 as dVx = −Cˆ D* (Vx − Wx )2 Vɺ x = dt (8.115) dVy = −Cˆ D* (Vx − Wx )(Vy − Wy ) − g Vɺ y = dt (8.116) dVz = −Cˆ D* (Vx − Wx )(Vz − Wz ) Vɺ z = dt (8.117) Updrafts and downdrafts are usually so small (and usually have the same effect as a crosswind for reasons we shall later describe) that we neglect them completely. Thus, we shall set Wy equal to zero from now on. We will now look first at the effect where only a crosswind is present (i.e., where Wx = Wy = 0) and then examine the effect of a headwind or tailwind. If we make this substitution into Equations 8.115 through 8.117, we obtain dVx = −Cˆ D* Vx2 Vɺ x = dt (8.118) dVy = −Cˆ D* VxVy − g Vɺ y = dt (8.119) dVz = −Cˆ D* Vx (Vz − Wz ) Vɺ z = dt (8.120) Equations 8.118 and 8.119 are identical to Equations 8.55 and 8.56 from our earlier work in the zero wind case. If we now change from time to space variables, as we did in the zero wind case, and recall that dt/dx = 1/(dx′/dt), then we arrive at the equations as follows: © 2014 by Taylor & Francis Group, LLC Vx′ = −Cˆ D* Vx (8.121) g Vy′ = −Cˆ D* Vy − Vx (8.122) Vz′ = −Cˆ D* (Vz − Wz ) (8.123) Trajectories 253 Once again, the prime symbol represents differentiation with respect to x, and Equations 8.121 and 8.122 are identical to those developed for the zero wind case. Now we have already solved differential Equations 8.121 and 8.122 under their previous guise with the result of  x  Vx = Vx0 exp  − Cˆ D* dx1     0  ∫  x  Vy = exp  − Cˆ *D dx1  Vy0 −    0  ∫ x ∫ 0  g V  x (8.124)    x2   Cˆ *D dx1  dx2  exp       0  ∫ (8.125) Equation 8.123 is somewhat more difficult to solve. It is a first-order, linear differential equation of the form y΄ + P(x)y = Q, whose solution, after the necessary integrations and substitution of initial conditions that at x = 0, Vz = 0, is  x x x  Vz = exp  − Cˆ *D dx1  Cˆ *D Wz exp  Cˆ *D dx1  dx2      0 0 0  ∫ ∫ ∫ (8.126) From Equation 8.124, we see that the exponential is Vx /Vx0, and this can be inserted directly into Equation 8.126. Also if we assume that Wz is a constant, it can be removed from the integral to give Vz = We can integrate ∫  Cˆ D* exp  0  x x  x Vx Wz Cˆ D* exp  Cˆ D* dx1  dx2 Vx0   0  0 ∫ ∫ (8.127)  x ∫ Cˆ * dx  dx by parts in Equation 8.127 to yield D 1 2 0 x x  x Cˆ D* exp  Cˆ D* dx1  dx2 = exp  Cˆ D* dx1  Cˆ D* dx2 −     0 0 0  0 x ∫ ∫ ∫ ∫ x  x  x  Cˆ D* dx2  Cˆ D* exp  Cˆ D* dx1  dx1      0  0 ∫∫ 0 ∫ (8.128) The integral of the last term can be solved through a series of substitutions and evaluations at the limits to yield  x  x ˆ ˆ * *   CD exp CDdx1 = exp  Cˆ D* dx1  − 1     0  0  0 x ∫ © 2014 by Taylor & Francis Group, LLC ∫ ∫ (8.129) Ballistics: Theory and Design of Guns and Ammunition 254 If this is inserted into Equation 8.127, the result is Vz =    x Vx Wz exp  CD* dx1  − 1  Vx0      0  ∫ (8.130) We can further manipulate Equation 8.130 by inserting the value of the exponential from Equation 8.124. In doing so, we get Vz =  Vx V   Vx  Wz  0 − 1  = Wz  1 − x  V Vx0 V x0   x   (8.131) Since 0 < Vx < Vx0 , Vz always has to be less than the wind speed Wz. Thus, Wz is an upper bound on Vz. If we examine the deflection due to a constant crosswind, we can write t t  t  1 V  z = Vzdt = Wz  1 − x  dt = Wz  t 0 − V V  x0 x0   0 0  ∫ ∫  x  z = Wz  t −  V x0    Vxdt   0  t ∫ (8.132) Equation 8.132 is known as the lag rule for predicting crosswind effects. It is an exact solution for a constant crosswind. The quantity in the brackets is known as the lag time because a projectile in a real atmosphere would take longer to reach the same range than one fired in a vacuum. Another interesting point is seen from examination of Equation 8.131. If Vx is always equal to the initial x velocity, no matter how hard the wind blows, the projectile will not be affected. Thus, a rocket motor that maintains the initial x velocity could make the projectile insensitive to wind, a concept called automet. Note also that if the thrust is greater than the initial velocity, the projectile will actually move into the wind. We consider next the effect of a variable crosswind. A simple way to model this effect on a projectile is to superimpose solutions for constant crosswinds over incremental distances and piece the resultant trajectory together. This technique of superposition works only with linear phenomena. However, since Equation 8.132 is linear in x and t, we can apply this method. An alternative approach would be to apply Equation 8.132 in a piecewise fashion using the information from the previous calculation in the subsequent one. To do this, we shall rewrite Equation 8.132 as a difference equation  ∆x  ∆zi = wzi  ∆ti − i  V xi0   (8.133) where ∆zi is the distance traveled in the z-direction from time i − 1 to the time i ∆xi is the distance traveled in the x-direction from time i − 1 to the time i ∆ti is the time between time i − 1 to the time i wzi is the constant crosswind acting on the projectile between time i − 1 and time i Vxi0 is the x-velocity at time i − 1 © 2014 by Taylor & Francis Group, LLC Trajectories 255 We can rewrite Equation 8.133 as  ( x − xi − 1 )  zi − zi −1 = wzi (ti − ti −1 ) − i  Vxi 0   (8.134) To use this method, one must first tabulate t, Vx, and x as described earlier and then perform the calculation for z at each interval. With some modifications, a forward difference technique can also be used. These tedious calculations are best done with a computer program for small intervals of time. We will now examine the effects of a constant range wind, both head-on and a tailwind. We do this by comparing the effects to a flat fire, no-wind flight and will determine the effects on time of flight, impact, and velocity at impact. We make the initial assumption that there is no crosswind, i.e., Wy = Wz = 0, and insert this into Equations 8.115 through 8.117, the component differential equations for a pointmass, flat fire trajectory: dVx = −Cˆ D* (Vx − Wx )2 Vɺ x = dt (8.135) dVy = −Cˆ D* (Vx − Wx )Vy − g Vɺ y = dt (8.136) dVz = −Cˆ D* (Vx − Wx )Vz Vɺ z = dt (8.137) Because there is no crosswind, Equation 8.137 reduces to Vz = 0 By a change of time to space variables and various algebraic manipulations, we can change Equation 8.135 to W   Vx′ + Cˆ D* Vx = Cˆ D* Wx  2 − x  Vx   (8.138) Similarly, we do the same to Equation 8.136 and arrive at a distance equation in a y-variable only  W Vy′ + Cˆ D*  1 − x Vx  −g   Vy = V x  (8.139) Recall from our earlier discussion that the wind speed is about two orders of magnitude smaller than the projectile velocity, so mathematically we can express this condition as Wx ≤ 0.01 Vx © 2014 by Taylor & Francis Group, LLC Ballistics: Theory and Design of Guns and Ammunition 256 We can then rewrite Equations 8.138 and 8.139 allowing them to be equalities as follows: Vx′ + Cˆ D* Vx = 2Cˆ D* Wx (8.140) g Vy′ + Cˆ D* Vy = − Vx (8.141) and We can then solve these equations for Vx and Vy. As we saw in the earlier solutions for the constant crosswind, with appropriate integrations, algebraic manipulation, and the insertion of the initial condition that at x = 0, Vx = Vx0, we see that  x  x  x   x  ˆ ˆ ˆ * * *     Vx = exp − C D dx1 2 C D Wx exp C D dx1 dx2 + Vx0 exp  − Cˆ D* dx1         0 0 0   0  ∫ ∫ ∫ ∫ (8.142) From Equation 8.129, recall that  x  x ˆ ˆ * *   CD exp CDdx1 = exp  Cˆ D* dx1  − 1.     0  0  0 x ∫ ∫ ∫ Using this fact and by substituting it into Equation 8.142, factoring the result, and considering that Wx is constant, we arrive at   x   x  Vx = Vx0 exp  − Cˆ D* dx1  + 2Wx 1 − exp  − Cˆ D* dx1         0   0    ∫ ∫ (8.143) The first term on the RHS of Equation 8.143 is simply the velocity decay caused by drag of the projectile. The second term is the effect of the range wind on it. If we examine Equation 8.124, which was an analysis for a firing in the absence of range wind, the first term of Equation 8.143 represents Vx, the x-velocity with no wind. The second term, when we substitute for the exponential, then represents the effect of the range wind on the flight. Thus, we can see the range wind effects shown as the variable of interest with a tilde (∼) in the following: ɶx = Vx + 2Wx 1 − Vx  V    Vx0  (8.144) This equation shows that at any time, t, a tailwind (i.e., one blowing in the positive x-direction) has the effect of increasing the velocity (relative to the ground), while the opposite is true of a headwind. This is important because if we had a table of velocities versus range for the no-wind case, we could then tabulate the effect of range wind. © 2014 by Taylor & Francis Group, LLC Trajectories 257 If we now look at the y-velocity, we can operate on Equation 8.141 with the initial condition that at x = 0, Vy = Vy0. This provides us with the solution of the space variable equation  x x x    x g * * ˆ ˆ     exp C D dx1 dx2 + Vy0 exp  − Cˆ D* dx1  Vy = − exp − C D dx1   Vx       0  0 0 0  ∫ ∫ ∫ ∫ (8.145) At this point, we can introduce the range wind by inserting Equation 8.143 for Vx arriving at  x x Vy = − g exp  − Cˆ *D dx1     0 0 ∫ ∫  + Vy0 exp  −   ∫  x  exp  Cˆ *D dx1   0        x Vx0 exp  − Cˆ *D dx1  + 2Wx 1 − exp  −   0       ∫ ∫ ∫  * ˆ C D dx1    0   dx2 x  Cˆ *D dx1   0  x (8.146) This rather complicated integral can be simplified somewhat; however, another approach [1] to the problem that makes use of the no-wind method used previously in Equation 8.144 simplifies things even further. This is seen as  x   x x x  Vy0 g * * ˆ* ˆ ˆ ɶ       Vy = − exp − C D dx1 C V exp D dx1 dx 2 + x0 exp − C D dx1 ɶx     V   Vx0  0   0 0 0  ∫ ∫ ∫ ∫ or since Vy0 Vx0 = tan φ0  x x x  1 * ˆ ɶ    Cˆ *D dx1  dx2 + Vx tan φ0 Vy = − g exp − C D dx1 exp ɶx   V    0 0 0  ∫ (8.147) ∫ ∫ (8.148) Further use of Ṽx and some algebraic manipulation gives x ɶy = Vx tan φ0 − gVx V 1 ∫ V Vɶ dx 0 x 2 (8.149) x Recalling Equation 8.144, we can rewrite the denominator of the integral as 2 ɶx = Vx2 + 2Wx  Vx − Vx  = Vx2  1 + 2Wx − 2Wx  VxV     Vx0  Vx Vx0    © 2014 by Taylor & Francis Group, LLC (8.150) Ballistics: Theory and Design of Guns and Ammunition 258 If we again use the fact that the wind velocity is at least two orders of magnitude smaller than the projectile velocity, the last two terms in the product on the RHS vanish, leaving ɶx ≈ Vx2 VxV (8.151) Then we can rewrite Equation 8.149 as x ɶy = Vx tan φ0 − gVx V 1 ∫V 0 2 x dx2 (8.152) This equation has exactly the same form as the flat fire equation for Vy. Hence, we can say that for a flat fire trajectory, with a small range wind compared to the projectile velocity, the vertical component of velocity is not appreciably affected. We can now turn our attention to the time of flight of a projectile with a constant range wind by first defining an average downrange velocity following the procedure by McCoy [1] as x ɶx = 1 V ɶxdx1 V avg x ∫ (8.153) 0 For Ṽx, we substitute Equation 8.144 giving ɶx = 1 V avg x x  Vx  dx1 x0  ∫  V + 2W − 2W V x x 0 x Performing the integration on the second term of the integrand and rewriting, we get x  ɶx = 1  2Wx x + Vxdx1 − 2 Wx V avg x Vx0 0  ∫  Vxdx1   0  x ∫ (8.154) Rearranging Equation 8.154 and knowing that the velocity averaging also applies to the no-wind case, i.e., x Vxavg = 1 Vxdx1 x ∫ (8.155) 0 we get ɶx = Vx + 2Wx  1 − Vxavg  V   avg avg Vx0   © 2014 by Taylor & Francis Group, LLC (8.156) Trajectories 259 The time of flight can be expressed as the range divided by the average velocity for either the case of no range wind or with range wind included. Thus, we can write R Vxavg (8.157) ɶt = R ɶx V avg (8.158) t= or By taking the reciprocal of Equation 8.158, performing judicious substitutions, gathering terms, and finally taking the reciprocal of the result, we can write   t ɶt =   t 1  1 + 2Wx  −   R Vx0        (8.159) This shows, as we would expect, that a tailwind (Wx positive) reduces the time of flight while a headwind (Wx negative) increases it. Let us summarize what we have done for crosswinds and range winds. We modified the flat fire equations to account for crosswind and range wind. We use them when the angle of departure and angle of fall are both below 5.7°. We solved the crosswind equations assuming constant and variable crosswinds and introduced the classic lag rule. With variable crosswinds, we saw it is fairly accurate to piece the trajectory together using locally constant values for the crosswind. We have solved the range wind equations assuming only constant range wind. We could treat variable range wind in a manner similar to variable crosswinds, but the difference in results is usually not worth the added effort. For range wind, we used a solution technique that compared the velocities, positions, and time to the no-wind case. Problem 29 A U.S. 37-mm AP projectile is fired with a muzzle velocity of 2600 [ft/s]. The projectile weighs 1.61 lbm. Assuming flat fire with K 2 = 0.841 [unitless] and using standard sea level met data (ρ = 0.0751 lbm/ft3, a = 1120 ft/s) 1. Create a table containing range (yards), impact velocity (ft/s), time of flight (TOF) (s), initial QE angle (min), and angle at impact (min) in 200 yard increments out to 1000 yards assuming no-wind effects. Answer: At 1000 yards, V = 1837 [ft/s]. 2. Determine the deflection of the projectile with a 20-mi/h crosswind blowing from left to right as viewed from behind the weapon. Answer: At 1000 yards, z = 6.217 [ft]. 3. Determine the impact velocity, change in TOF, and how high the projectile will hit if fired at the same QE’s with a 20-mi/h tailwind and no crosswind. Answer: The projectile will hit 1.402 in. higher than expected. © 2014 by Taylor & Francis Group, LLC Ballistics: Theory and Design of Guns and Ammunition 260 Problem 30 A British 12-in. projectile has a K3 of 0.8 and a weight of 850 lbm. If it is fired at an initial QE of 130 mil with a muzzle velocity of 2800 ft/s 1. Create a table of range (yards), altitude (yards), velocity (ft/s), time of flight (s), inclination angle (degrees), and drift (yards) if the projectile is fired with no wind. 2. Repeat part (1) if the projectile is fired with a headwind of 25 ft/s for the first 3000 yards of flight and a crosswind (left or right—your choice) of 35 ft/s for the remainder of the flight. Tabulate every 1000 yards with the impact location as the last entry in the table. Problem 31 The U.S. 0.30 caliber Ball M2 (projectile diameter = 0.308in.) was the standard infantry rifle cartridge in WWII. Based on data collected by McCoy [1], it is a 150 grain, flat-based Spitzer shape with a K3 of 0.491 from Mach 1.2 to 3. If it is fired with a muzzle velocity of 2780 ft/s, assuming standard sea level met data (ρ = 0.0751 lbm/ft3, a = 1120 ft/s): a. Create a table of range (yards), velocity (ft/s), time of flight (s), launch angle (minutes), impact angle (min), and drift (yards) if the projectile is fired with no wind every 100 yards out to 600 yards. b. Determine the deflection in inches assuming the projectile experiences a headwind of 20 ft/s for the entire flight. Tabulate every 100 yards out to 600 yards. c. Determine the deflection in inches assuming the projectile experiences a crosswind (left or right—your choice) of 20 ft/s for the entire flight. Tabulate every 100 yards out to 600 yards. Problem 32 Precision shooters are always in search of “tight groupings.” That is, the grouping of the impact points of the projectiles at a given range. We shall examine the rifle and projectile combination given in problem 31 and determine the effect each of several parameters has on the precision. Vary each of the following parameters individually by 1% (up and down) from problem 31 and determine the miss distance in inches for a 200 yard range. Please carry answers to five significant figures as a baseline of comparison: a. b. c. d. e. f. g. h. i. Muzzle velocity Projectile mass Drag coefficient (CD) Air density Air temperature (in Rankine) Launch angle of departure (we will assume this is due to weapon or shooter motion) Headwind (0 ± 0.2 ft/s) Crosswind (0 ± 0.2 ft/s) Choose any two of these and vary them together—what is the result? Is the answer simply a linear superposition of the two individual errors? Why or Why not? Is this true for all of the parameters? Can you draw any conclusions from looking at all of the results? © 2014 by Taylor & Francis Group, LLC Trajectories 261 8.4 Generalized Point Mass Trajectory In keeping with our practice of introducing ever-increasing complexity into our theory, we will now remove most of the restrictions of the earlier work. We will examine the effects of an unrestricted launch angle and make the high-angle fire of mortars and howitzers amenable to trajectory analysis. We still reserve for later study the effects on flight of a three-dimensional body whose shape, physical properties, and motions add a significant level of complexity to trajectory analysis. The aerodynamic behavior of a projectile can be examined from three separate viewpoints: motion affected only by the acceleration of gravity and the initial velocity (vacuum trajectory); motion affected by gravity, initial velocity, and aerodynamic drag (point mass trajectory); motion affected by the projectile’s shape, physical properties, and dynamics (which actually manifests itself as changing drag) as well as gravity and launch conditions. We will concentrate on the second viewpoint in this section. In the equations that follow, we are assuming that the projectile is still a cannon ball with all of its mass concentrated at one point. This allows us to continue to neglect the rigid body kinematics that would be present in a distributed mass. However, we shall include wind effects and earth-rotational effects, and therefore three-dimensional motion. As stated, flat fire restrictions are removed so that the analysis is applicable to all launch angles. We begin with the same set of equations of motion, except for the addition of a term for the Coriolis force, mΛ: F = ma m dV = dt ∑ F + m g + mΛ (8.160) (8.161a) where m is the projectile mass V is the velocity vector t is the time dV a= is the vector acceleration dt ΣF is the vector sum of all aerodynamic forces g is the vector acceleration due to gravity Λ is the vector Coriolis acceleration due to rotation of the earth We also recall from our earlier work with wind effects dV ɶ (V − W) + g = −Cˆ D* V dt (8.162) ˆ * = ρ SC /2m. In the aforementioned equations, where W is the wind velocity vector and C D D we have replaced the velocity vector V by the vector (V − W) because drag measurements are made relative to the air stream, not relative to the ground. We have also again replaced the scalar velocity (the speed) with Ṽ = |V − W|, which is the scalar difference of the projectile and wind velocities. A diagram of this is shown in Figure 8.6. © 2014 by Taylor & Francis Group, LLC Ballistics: Theory and Design of Guns and Ammunition 262 Wy Height y Wx Wz FD V0 j mg V De fle c ti on 0 i k Range x z FIGURE 8.6 Generalized point mass trajectory. Without repeating the entire procedure, it can be shown that we may separate Equation 8.162 into individual components to obtain the differential equations for a point mass: dVx ɶ (Vx − Wx ) = − Cˆ *D V Vɺ x = dt (8.163) dVy ɶ (Vy − Wy ) − g = − Cˆ *D V Vɺ y = dt (8.164) dVz ɶ (Vz − Wz ) = − Cˆ *D V Vɺ z = dt (8.165) ɶ = (Vx − Wx )2 + (Vy − Wy )2 + (Vz − Wz )2 V (8.166) The scalar velocity, Ṽ, is again In all of the aforementioned equations, the wind velocity is variable and is considered positive when it blows in the positive direction of one of the coordinate axes. Equations 8.163 through 8.165 are nonlinear, coupled differential equations which are the exact solution to Newton’s laws governing the motion of a projectile affected by wind, gravity, and aerodynamic drag. These equations are coupled through Equation 8.166. Now, as we did in our discussion of flat fire, we would like to evaluate Equations 8.163 through 8.165 by using the downrange distance, x, as the independent variable. To do this, we simply note that for each of the time derivatives, we can write dVx dVx dVx dt dx = = Vx = VxVx′ Vɺ x = dt dt dx dt dx (8.167) Vɺ y = VxVy′ (8.168) Vɺ z = VxVz′ (8.169) And similarly © 2014 by Taylor & Francis Group, LLC Trajectories 263 We can now write the three equations of motion with x as the independent variable as follows: Vx′ = Vy′ = ɶ V 1 ɺ Vx = − Cˆ *D  Vx  Vx ɶ V 1 ɺ Vy = − Cˆ *D  Vx  Vx Vz′ =   (Vx − Wx )  (8.170)   g   (Vy − Wy ) −    Vx   (8.171) ɶ  V 1 ɺ Vz = − Cˆ *D   (Vz − Wz ) Vx  Vx  (8.172) As we noted earlier, the vertical component of the wind, Wy, is usually extremely small and will be neglected in further treatment. Further, as we mentioned, these equations are impossible to solve in closed form and we must resort to numerical methods for their solution. Without the restrictions of flat fire, projectiles fired at high angles of departure may traverse the atmosphere to great altitudes. In their flight, they encounter air temperatures and pressures that constantly change. These changes must be accounted for in the numerical computations to adequately solve the trajectory. Hence, knowledge of the standard atmosphere must serve as input to the calculations. There are two standards in common use: Army Standard Metrology and the International Civil Aviation Organization (ICAO) atmosphere. ICAO atmosphere is the most used of the two. Temperature and pressure versus altitude are shown for the ICAO model in Figure 8.7. These atmospheric models are usually incorporated into ballistics codes. Now, to become familiar with the physics of the Coriolis acceleration which was brought to its final form by Gaspard de Coriolis in 1835, we will study the effects of the earth’s rotation on a flat fire, vacuum trajectory example. We do this because we will be able to see these effects without resorting to a computer for calculation. The effect is really due to the fact that the firing point and target are located on the rotating earth, thus when the projectile lands, the earth has rotated through an angle and has thus moved the target. Figure 8.8 shows the geometry of the earth, the latitude of the firing site, and the orientation of the axes. The Coriolis acceleration is defined as 60 60 50 50 Altitude (km) Altitude (km) 2ω × ( v Β/ Α )xyz = 2Ω × ( v )xyz 40 30 20 40 30 20 10 10 0 (a) (8.173) –60 –40 –20 0 20 Temperature (°C) 0 (b) 40 80 Pressure (kPa) 120 FIGURE 8.7 International Civil Aviation Organization (ICAO) models for (a) atmospheric temperature and (b) pressure. © 2014 by Taylor & Francis Group, LLC Ballistics: Theory and Design of Guns and Ammunition 264 Ω x N y AZ x z L .. z FIGURE 8.8 Angles used for Coriolis acceleration calculations. Picture on the right represents a map of the corresponding area on the globe. We have written Equation 8.173 in this way because the angular velocity we are considering is that of the earth and our projectile velocity is relative to our firing position (and therefore the earth), which moves with the x–y–z coordinate system. For this equation to be useful to us, we have to write the earth’s angular velocity, Ω, in terms of our x–y–z coordinate system. We will see that this acceleration is independent of the projectile weight but dependent upon its velocity. From Figure 8.8, we see that we can readily define Ω in terms of our moving coordinate system as Ω = Ω cos L cos AZi + Ω sin Lj − Ω cos L sin AZk (8.174) If we also note that v is defined as (v )xyz = Vxi + Vy j + Vz k (8.175) Then inserting Equations 8.174 and 8.175 into Equation 8.173 gives us 2Ω × ( v )xyz (Vz sin L + Vy cos L sin AZ)i    = 2Ω (−Vz cos L cos AZ − Vx cos L sin AZ))j (Vy cos L cos AZ − Vx sin L)k  (8.176) We will write the Coriolis acceleration in terms of a D’Alembert force (i.e., the negative of what we have in Equation 8.176), so we shall define the Coriolis term in our equation of motion (Equation 8.161) as Λ = −2Ω × ( v )xyz © 2014 by Taylor & Francis Group, LLC (−Vy cos L sin AZ − Vz sin L)i    = 2Ω (Vx cos L cos AZ + Vz cos L cos AZ)j (Vx sin L − Vy cos L cos AZ)k  (8.177) Trajectories 265 Here we need to define the following variables: Λ is the vector Coriolis acceleration Ω is the angular velocity of the earth about its polar axis = 0.00007292 (rad/s) L is the latitude of the firing site, positive in the northern hemisphere, negative in the southern AZ is the azimuth angle of fire, measured clockwise from north Vx, Vy, Vz is the velocity in the x, y, z directions, respectively, positive along the positive coordinate axes Now that we have defined some terminology, we shall examine the effect that the Coriolis acceleration has on a vacuum trajectory. While this is stretching the vacuum trajectory much beyond its usefulness in ballistics, we remind the reader that the purpose is to demonstrate the physics that result from Coriolis effects. We begin by recalling Equation 8.161 m dV = ΣF + m g + mΛ dt (8.161) Now, since this is a vacuum trajectory, the force term on the RHS is zero, and we can divide by the mass, m, to obtain the vector equation for a vacuum trajectory dV =g+Λ dt (8.178) Rewriting Equation 8.178 in terms of its vector components gives (note that the “g” term appears only in Equation 8.180) dVx = 2Ω(−Vy cos L sin AZ − Vz sin L) dt (8.179) dVy = 2Ω(−Vx cos L sin AZ + Vz cos L cos AZ) − g dt (8.180) dVz = 2Ω(Vx sin L − Vy cos L cos AZ) dt (8.181) We shall now provide examples of the effect. These examples are based on the work of McCoy and can also be found in his work [1]. Let us consider first a purely vertical firing (i.e., Vx = Vz = 0). One may, initially, consider this a trivial example, but for test purposes we occasionally do fire vertically. And, by the way, as we will see, what goes up does not come straight down. Let us also choose due east as positive x, so AZ = 90°. With these assumptions, Equations 8.179 through 8.181 become © 2014 by Taylor & Francis Group, LLC dVx = −2Ω ΩVy cos L dt (8.182) dVy = −g dt (8.183) dVz =0 dt (8.184) Ballistics: Theory and Design of Guns and Ammunition 266 These equations are well behaved and no longer coupled, so we can solve them independently. We shall integrate Equation 8.182 by first rewriting it, then integrating it: dy dVx = −2Ω cos L dt dt (8.185) Vx = −2Ωy cos L + C (8.186) To determine C, we know that at y = y0, Vx = 0 so we can write Vx = −2Ωy cos L + 2Ωy0 cos L = −2Ω cos L( y − y0 ) (8.187) If you recall our coordinate system, this means a projectile fired straight up will drift to the west and one fired (or dropped) straight down will drift to the east. Now we will integrate Equation 8.183 to get Vy = − gt + C (8.188) Again, solving for the constant by inserting the initial conditions that at t = 0, Vy = Vy0 , we get Vy = Vy0 − gt (8.189) Now we shall rewrite and integrate Equation 8.189 a second time making use of the fact that at t = 0, y = y0, to obtain y = Vy0 t − 1 2 gt + y0 2 (8.190) We can now insert Equation 8.190 into Equation 8.187 and rewrite it as dx 1 1     = −2Ω cos L  y0 + Vy0 t − gt 2 − y0  = −2Ω cos L  Vy0 t − gt 2  dt 2 2     (8.191) This can be integrated using the initial conditions that at t = 0, x = 0 to give 1   x = −Ω cos L  Vy0 t 2 − gt 3  3   (8.192) Let us now look at the special case of a bomb dropped from a given height with Vy0 = 0 and let y = 0. If we know the altitude from which we are dropping the bomb, we can determine its time of flight from Equation 8.190, thus y0 = © 2014 by Taylor & Francis Group, LLC 1 2 gt 2 (8.193) Trajectories 267 or t= 2 y0 g (8.194) If we insert this into Equation 8.192, we get  2y  1 x = gΩ cos L  0  3  g  3/ 2 (8.195) This says that since we are on the positive x-axis, the bomb will drift to the east. This drift would be greatest at the equator and zero at the poles. Another example that uses the vacuum trajectory analysis is a projectile that is fired vertically upward with velocity, Vy0 . We can find the time to apogee from Equation 8.189 knowing that at apogee, Vy = 0: t= Vy0 g (8.196) If we insert this value of t into Equation 8.192, we get 2 x = − Ωgt 3 cos L 3 (8.197) The time to apogee can be put in terms of the height at apogee, ys, through Equation 8.190 ys = 1 2 gt 2 (8.198) ts = 2 ys g (8.199) Therefore, the time to apogee is And therefore the Coriolis-caused displacement at apogee along the x-axis is found by inserting Equation 8.199 into Equation 8.197 giving 2 ys3 4 xs = − Ω cos L g 3 (8.200) Lastly, we can apply the Coriolis analysis to the vacuum trajectory, flat fire situation and determine a correction for the acceleration in that case. We begin by making the usual assumptions for the flat fire trajectory of Vy ≪ Vx © 2014 by Taylor & Francis Group, LLC and Vz ≪ Vx (8.201) Ballistics: Theory and Design of Guns and Ammunition 268 We substitute these into Equations 8.179 through 8.181 yielding dVx ≈0 dt (8.202) dVy ≈ 2ΩVx cos L sin AZ − g dt (8.203) dVz ≈ 2ΩVx sin L dt (8.204) Solution of Equation 8.202 with the initial conditions of Vx = Vx0 at t = 0 yields Vx ≈ Vx0 (8.205) Solution of Equation 8.203 after insertion ofV Equation 8.205 with the initial conditions of Vy = Vy0 at t = 0 and integrating yields    2ΩVx0  Vy ≈ Vy0 − gt 1 −   cos L sin AZ     g  (8.206) Solution of Equation 8.204 with the initial conditions of Vz = 0 at t = 0 yields after insertion of Equation 8.205 and integrating Vz ≈ 2ΩVx0 t sin L (8.207) If we now integrate Equations 8.205 through 8.207 subject to x = 0, y = y0, and z = 0 at t = 0 to get the displacements in the x, y, and z directions, we get x ≈ Vx0 t y ≈ y0 + Vy0 t − gt 2 2   2ΩVx0 1 −    g z = ΩVx0 t 2 sin L (8.208)    cos L sin AZ    (8.209) (8.210) If we want to parameterize Equations 8.209 and 8.210 in terms of the downrange distance, x, we can rewrite Equation 8.208 as t≈ © 2014 by Taylor & Francis Group, LLC x Vx0 (8.211) Trajectories 269 We can then insert this value of time into Equations 8.209 and 8.210 to obtain y ≈ y0 + Vy0 Vx0 x−  gx 2   2ΩVx0  1−  cos L sin AZ  2  2Vx0   g   (8.212) and z≈ Ωx 2 sin L Vx0 (8.213) Equation 8.212 was rearranged in this form (including the substitution of tan ϕ 0) for comparison with Equation 8.37 also modified to include a y0: gx 2 2V02 (8.37)  gx 2   2ΩVx0  1−  cos L sin AZ  2  2Vx0   g   (8.214) y ≈ y0 + x tan φ0 − y ≈ y0 + x tan φ0 − From this comparison, we see that the incorporation of the Coriolis acceleration in the flat fire vacuum trajectory manifests itself in a modification to the gravitational term. Thus, as defined in Ref. [1], we can define a Coriolis factor, fC, as    2ΩVx0  f C = 1 −   cos L sin AZ     g  (8.215) and we could rewrite Equation 8.214 as y ≈ y0 + x tan φ0 − f C gx 2 2Vx20 (8.216) If we look closely at Equation 8.214, we note several things: the value of cos L is anywhere between 0 and 1 for all possible latitudes; thus, if we were firing due north or due south, there would be no effect on the vertical component of impact; if we fired due east (AZ = 90°), the Coriolis effect essentially weakens the gravity term and the bullet would hit high; a due west firing would strike low; and the maximum effect on gravity is to alter it by 1.8%. Since sin L fluctuates between +1 and −1, the drift, the z-component, will vary right or left depending on the hemisphere where the firing occurs. Now that the physics of the Coriolis effect are understood, the only difference when applied to the nonvacuum point mass trajectory is the fact that the velocity is changing with time due to drag. This is best handled numerically and will not be covered here. © 2014 by Taylor & Francis Group, LLC Ballistics: Theory and Design of Guns and Ammunition 270 y V0 = 2800 ft/s V mg 0 = 130 mils x R = ?, z = ? FIGURE 8.9 Graphical representation of long-range fire for problem 33. In summary, for the generalized point mass trajectory, we included drag but ignored the projectile’s dynamic effects on drag. We described the origins of the Coriolis acceleration acting on a projectile. The physics was demonstrated through the vacuum trajectory and further examined with the flat fire assumptions. Incorporation of the Coriolis acceleration into the generalized point mass assumption is only affected by the variation of velocity over the trajectory and best handled numerically. Problem 33 A projectile fired from a British 12 in. Mark IX naval gun had a muzzle velocity of 2800 ft/s and was fired at a QE of 130 mil (Figure 8.9). Assuming a vacuum trajectory, at what deflection would the shot hit the ground? Assume the firing is taking place at 50° south latitude and the round is being fired due north. Answer: z = −75.8 [ft]. 8.5 Six Degree-of-Freedom (6-DOF) Trajectory In keeping with our plan of increasing the complexity of our analyses to approach more closely the physical realities of projectile flight, we will now consider the projectile as a distributed mass. Since projectiles are relatively stiff structures, a 6-DOF model can adequately represent its position and attitude at any time. Each degree of freedom is tied to a coordinate necessary to completely describe the position of a body. While this model is necessarily more complex than anything we have studied so far, the underlying physical principles remain the same. In the following work, we will use vectors (boldfaced, nonitalicized letters) in many of the derivations. We continue to do this because of the brevity and elegance of the notation. In the equations that follow, we assume that the projectile is a rigid body of finite length with its mass distributed based on its geometry. This allows us to account for the effect of projectile attitude on drag and also allows the full dynamics to come into play. We shall use direction cosines with respect to the projectile axis of symmetry (and thus a coordinate system with unit vectors i, j, k that translates with the CG but does not rotate and remains aligned with the projectile axis) as opposed to Eulerian angles (angles that are measured relative to the inertial coordinate system). This is illustrated in Figure 8.10. © 2014 by Taylor & Francis Group, LLC Trajectories 271 i y, 2 V at j k mg x, 1 z, 3 FIGURE 8.10 Coordinate system for 6-DOF model. Once again we restate the equations of motion, which for generality includes a term for rocket propulsion of the projectile. However, because this force is usually assumed to be aligned with the projectile’s longitudinal axis, its effect on the motions we will study is uncoupled from the other motions and may be added in afterward. Consequently, we will ignore it in our further work: F = ma m dV = dt (8.217) ∑ F + m g + mΛ + ∑ R T (8.218) where m is the projectile mass V is the projectile velocity vector t is the time dV a= is the vector acceleration dt ΣF is the vector sum of all aerodynamic forces g is the vector acceleration due to gravity Λ is the vector Coriolis acceleration due to rotation of the earth ΣRT is the vector sum of all rocket thrust forces (to be ignored) We can also write the equation for the conservation of angular momentum as dH = dt ∑M + ∑R M where H is the vector angular momentum of the projectile ΣM is the vector sum of all aerodynamic moments ΣR M is the vector sum of all rocket thrust moments (to be ignored) © 2014 by Taylor & Francis Group, LLC (8.219) 272 Ballistics: Theory and Design of Guns and Ammunition Because the projectile is assumed to be symmetric, every axis transverse to the longitudinal axis through the CG is a principal axis of inertia. The longitudinal axis itself is also, of course, a principal axis of inertia. The definition of the inertia tensor, which we will use, is  I xx  I =  −I yx  −I zx −I xy I yy −I zy −I xz   −I yz  I zz  (8.220) Here the diagonal terms are called the moments of inertia and the off-diagonal terms are called the products of inertia. We know that there is a rotation that can be applied to this tensor such that the off-diagonal elements go to zero. In this orientation, the axes are said to be principal axes of inertia and the tensor is written as I x  I = 0  0 0 Iy 0 0  0 I z  (8.221) In our coordinate system, we shall define the unit vectors, i, j, and k so that they all lie along the projectile’s principal axes. Because of this unique situation, the total angular momentum of the projectile can be expressed as the sum of two vectors: the angular momentum about i and the angular momentum about any axis perpendicular to i through the CG. Since the i-axis is what we usually call the polar axis, we will denote the polar moment of inertia as IP. With the symmetry of the projectile, the other moments of inertia about axes perpendicular to i are known as the transverse moments of inertia, Iy = Iz = IT. We can then rewrite the inertia tensor as IP  I= 0  0 0 IT 0 0  0 I T  (8.222) If a projectile is spinning at spin rate, p, the angular momentum about the polar axis is defined as HP = I P pi (8.223) The angular momentum about any transverse axis is defined as  di  HT = I T  i ×   dt  (8.224) With this, we can write the total momentum vector as  di  H = I P pi + I T  i ×   dt  © 2014 by Taylor & Francis Group, LLC (8.225) Trajectories 273 By defining a specific angular momentum, h = H/IT, we can write h= I P p  di  i + i×  IT  dt  (8.226) If we take the derivative of Equation 8.226 with respect to time, we get I p di  di di   d 2i  dh I P = pɺ i + P +  ×  + i×  I T dt  dt dt   dt 2  dt I T (8.227) Since the cross-product of a vector with itself is zero, we get I p di  d 2i  dh I P = pɺ i + P + i×  I T dt  dt 2  dt I T (8.228) In anticipation of a later need, we shall take the dot product and cross-product of the vector h with the unit vector i to get  I p  di   I p h ⋅ i =  P i +  i ×  ⋅ i = P d I t IT   T   (8.229)  I p  di   di h × i =  P i +  i ×  × i = d d I t t    T (8.230) In Equations 8.229 and 8.230, we have used the orthogonality properties of vectors as follows: i⋅i = 1 i⋅j = i⋅k = 0 i×i = 0 i × j = k → (i × j) × i = (k) × i = j ∴ (i × j) × i = j We will now examine all of the forces and then the moments acting on the projectile and combine them into Equations 8.218 and 8.219. We have discussed all of these items in Chapter 6, so we shall simply refresh their meanings briefly and move on. The first force acting on the projectile is the drag force, which acts opposite to the velocity vector, so we have 1 Drag Force = FD = − ρ SCD VV 2 (8.231) The second force is the lift force, which we modified for our coordinate system as follows: 1 Lift Force = FL = − ρ SCLα [V × (i × V )] 2 © 2014 by Taylor & Francis Group, LLC (8.232) Ballistics: Theory and Design of Guns and Ammunition 274 This equation contains a vector triple product in it that we replace with the relationship from vector algebra A × (B × C) = (A ⋅ C)B − (A ⋅ B)C (8.233) which, for the product in Equation 8.232, can be written as V × (i × V ) = V 2i − (V ⋅ i)V When inserted into Equation 8.232, we have Lift force = FL = 1 ρ SCLα [V 2i − (V ⋅ i)V] 2 (8.234) The next force is the Magnus force, brought on by the spin or roll of the projectile and taken from Equation 6.13 Magnus force = FM = 1  pd  ρ SV   CNpα (V × i) 2 V  (8.235) However, from Equation 8.229, we know that p = (IT/IP)(h · i) and, from vector algebra, V × i = −i × V. Then, we can manipulate Equation 8.235 to the form 1 I Magnus force = FM = − ρ SdCNpα  T 2  IP   (h ⋅ i)(i × V )  (8.236) Next we need to include the pitch damping force from Equation 6.15 where we will write v′ as the unit vector along the velocity vector: Pitch damping force = 1 1  di   di dv′  ρVSd   CNq + ρVSdCNαɺ  −  2 2  dt   dt dt  (8.237) If we assume dv′/dt = di/dt (this means that the rate at which the velocity vector is rotating to follow the curve of the trajectory is much smaller than the rate at which the axis of the projectile is moving) and we include Equation 8.230, we get a relation similar to Equation 6.18 Pitch damping force = 1  di   di  1 ρVSdCNq   + ρVSdCNαɺ   2  dt   dt  2 (8.238) 1 ρVSd(CNq + CNαɺ )(h × i) 2 (8.239) or by Equation 8.230 Pitch damping force = © 2014 by Taylor & Francis Group, LLC Trajectories 275 With all our forces now expressed in terms of our defined coefficients, we can divide Equation 8.218, omitting the rocket motor, by the projectile mass and inserting the coefficients to give ρ SdCNpα  I T dV ρVSCD ρ SCLα 2 =− V+ [V i − (V ⋅ i)V] − dt 2m 2m 2m  I P + ρVSd ( CNq + CNαɺ ) 2m   (h ⋅ i)(i × V )  (h × i) + g + Λ (8.240) We will now examine the moments involved in Equation 8.219, the first of which is the spin damping moment written as 1  pd  ρV 2Sd   Clp i 2 V  Spin damping moment = MS = (8.241) However, if we again insert Equation 8.229 into the aforementioned equation, we get 1 I ρVSd 2Clp T (h ⋅ i)i 2 IP Spin damping moment = MS = (8.242) The rolling moment comes from Equation 6.7 and is Rolling moment = MR = 1 ρV 2Sdδ FClδ i 2 (8.243) The overturning moment can be written from Equation 6.10 as Overturning moment = Mα = 1 ρ SdVCMα (V × i) 2 (8.244) The Magnus moment can be written from Equation 6.14 and, by using the relations of Equations 8.233 and 8.236, we get Magnus moment = Mpα = 1 I ρ Sd 2CMpα T (h ⋅ i)[V − (V ⋅ i)i] 2 IP (8.245) We can obtain the pitch damping moment by rewriting Equation 6.19 as well as using the relation of Equation 8.229 to get Pitch damping moment = Mq = ( ) 1 ρVSd 2 CMq + CMαɺ [h − (h ⋅ i)i] 2 (8.246) We can now place all of these relations into Equation 8.219, again omitting the rocket term, to yield dH = MS + MR + Mα + Mpα + Mq dt © 2014 by Taylor & Francis Group, LLC (8.247) Ballistics: Theory and Design of Guns and Ammunition 276 Equation 8.247 can be changed to a more desirable form by dividing by IT, which yields dh MS MR Mα Mpα Mq = + + + + IT IT IT IT IT dt (8.248) This, in turn, may be rewritten by inserting the various moment equations derived earlier as ρ V 2Sdδ FClδ dh ρ VSd 2Clp ρVSdCMα = (h ⋅ i)i + (V × i) i+ dt 2I P 2I T 2I T + ρVSd 2 (CMq + CMαɺ ) ρ Sd 2CMpα (h ⋅ i)[V − (V ⋅ i)i] + [h − (h ⋅ i)i] 2I P 2I T (8.249) Note that the equations of motion are highly coupled to one another and the reason we call the model a 6 DOF is readily apparent. When we break the equations up into their individual components, we have six equations and six unknowns (x, y, z, p, α, and β). Let us recall that the x, y, z axes are axes fixed to the earth, independent of the projectile, while i is the unit vector along the axis of symmetry of the projectile and has components along the x, y, z earth axes and p, α, and β are the spin rate, pitch, and yaw angles, respectively. For convenience, clarity, and to facilitate analysis, we will relabel the x, y, z unit vectors (normally i, j, k) as e1, e2, and e 3, respectively, letting the subscripts denote the x, y, z axes in that order (see Figure 8.10). Then, in terms of components in the earthfixed system, h = h1e1 + h2e2 + h3e3 (8.250) i = i1e1 + i2e2 + i3e3 (8.251) V = V1e1 + V2e2 + V3e3 (8.252) W = W1e1 + W2e2 + W3e3 (8.253) v = V − W = (V1 − W1 )e1 + (V2 − W2 )e2 + (V3 − W3 )e3 (8.254) v1 = (V1 − W1 ) v2 = (V2 − W2 ) v3 = (V3 − W3 ) (8.255) v = v12 + v22 + v32 (8.256) We shall also define and further defining and © 2014 by Taylor & Francis Group, LLC Trajectories 277 We can insert the definition for v in place of V in Equations 8.245 and 8.249 to yield ρ SdCNpα  I T dV ρ vSCD ρ SCLα 2 =− v+ [v i − (v ⋅ i)v] − dt 2m 2m 2m  I P +   (h ⋅ i)(i × v )  ρ vSd(CNq + CNαɺ ) (h × i) + g + Λ 2m (8.257) and ρ v 2Sdδ FClδ ρ vSdCMα dh ρ vSd 2Clp = i+ ( v × i) (h ⋅ i)i + dt 2I P 2I T 2I T + ρ Sd 2CMpα ρ vSd 2 (CMq + CMαɺ ) (h ⋅ i) [ v − (v ⋅ i)i ] + [h − (h ⋅ i)i] 2I P 2I T (8.258) Our goal is to examine Equations 8.257 and 8.258 and break each into three equations, one for each coordinate direction (actually for the acceleration in each coordinate direction). But before we attempt to break Equations 8.257 and 8.258 into their components, it will be best to solve for some of the vector quantities that occur in them. Beginning with the second term of Equation 8.257 we can, with appropriate vector multiplication, obtain ( ) ( ) (v ⋅ i)v = v12i1 + v1v2i2 + v1v3i3 e1 + v1v2i1 + v22i2 + v2v3i3 e2 ( ) + v1v3i1 + v2v3i2 + v32i3 e3 (8.259) Another useful relation is that (v ⋅ i) = (v1e1 + v2e2 + v3e3 ) ⋅ (i1e1 + i2e2 + i3e3 ) (8.260) (v ⋅ i) = v1i1 + v2i2 + v3i3 (8.261) or However, we can show that cos α t = (v ⋅ i) v1i1 + v2i2 + v3i3 = v v (8.262) The next relations in Equation 8.257 are (h ⋅ i) = h1i1 + h2i2 + h3i3 (8.263) e1 e2 e3 (i × v ) = i1 i2 i3 = (i2v3 − i3v2 )e1 + (i3v1 − i1v3 )e2 + (i1v2 − i2v1 )e3 v1 v2 v3 (8.264) and © 2014 by Taylor & Francis Group, LLC Ballistics: Theory and Design of Guns and Ammunition 278 Now, also in Equation 8.257 is the term (h ⋅ i)(i × v ) = ( h1i1 + h2i2 + h3i3 )[(i2v3 − i3v2 )e1 + (i3v1 − i1v3 )e2 + (i1v2 − i2v1 )e3 ] (8.265) But with the fact that (h · i) = IPp/IT as shown earlier, then we can write (h ⋅ i)(i × v ) = I p I p IP p (i2v3 − i3v2 )e1 + P (i3v1 − i1v3 )e2 + P (i1v2 − i2v1 )e3 IT IT IT (8.266) Another relation we have to deal with is e1 e2 e3 (h × i) = h1 h2 h3 = ( h2i3 − h3i2 )e1 + ( h3i1 − h1i3 )e2 + ( h1i2 − h2i1 )e3 i1 i2 i3 (8.267) The next relation we generate with the help of Equation 8.229 is (h ⋅ i)i = I p I p IP p i1e1 + P i2e2 + P i3e3 IT IT IT (8.268) In Equation 8.258, we need e1 e2 e3 (v × i) = v1 v2 v3 = (i3v2 − i2v3 )e1 + (i1v3 − i3v1 )e2 + (i2v1 − i1v2 )e3 i1 i2 i3 (8.269) And finally in Equation 8.258 (v ⋅ i)i = (v1i1 + v2i2 + v3i3 )(i1e1 + i2e2 + i3e3 ) (8.270) or ( ) ( ) (v ⋅ i)i = v1i12 + v2i1i2 + v3i1i3 e1 + v1i1i2 + v2i22 + v3i2i3 e2 ( ) + v1i1i3 + v2i2i3 + v3i32 e3 (8.271) Let us now look at Equation 8.257 with all of the vector quantities broken into their components dV1 dV2 dV3 ρ vSCD (v1e1 + v2e2 + v3e3 ) e1 + e2 + e3 = − dt dt dt 2m ρ SCLα 2 [v i1e1 + v 2i2e2 + v 2i3e3 − v cos α t (v1e1 + v2e2 + v3e3 )] + 2m ρ SdCNpα  I T   I P p  − [(v3i2 − v2i3 )]e1 + (v1i3 − v3i1 )e2 + (v2i1 − v1i2 )e3 ] 2m  I P   I T  + ρ vSd ( CNq + CNαɺ ) [( h2i3 − h3i2 )e1 + ( h3i1 − h1i3 )e2 + ( h1i2 − h2i1 )e3 ] 2m + g1e1 + g 2e2 + g 3e3 + Λ 1e1 + Λ 2e2 + Λ 3e3 © 2014 by Taylor & Francis Group, LLC (8.272) Trajectories 279 Similarly, let us perform the same operation on Equation 8.258 ρ vSd 2Clp  I P p  dh1 dh dh e1 + 2 e2 + 3 e3 =  I  (i1e1 + i2e2 + i3e3 ) dt dt dt 2I P  T  + ρ v 2Sdδ FClδ ρ vSdCMα (i1e1 + i2e2 + i3e3 ) + [(v2i3 − v3i2 )e1 + (v3i1 − v1i3 )e2 + (v1i2 − v2i1 )e3 ] 2I T 2I T + ρ Sd 2CMpα  I P p   I  [(v1e1 + v2e2 + v3e3 ) − v cos α t (i1e1 + i2e2 + i3e3 )] 2I P  T  + ρ vSd 2 ( CMq + CMαɺ )    IP p  (i1e1 + i2e2 + i3e3 ) ( h1e1 + h2e2 + h3e3 ) −   2I T  IT    (8.273) We will first operate on Equation 8.272 by collecting all of the terms with the unit vectors e1, then e2 and e3, and by putting them into the equations for linear and angular momentum: ρ SdCNpα p pSCLα 2 dV1 ρ vSCD v i1 − vv1 cos α t  − v1 + =− (v3i2 − v2i3 ) dt 2m 2m 2m + ρ vSd ( CNq + CNαɺ ) 2m ( h2i3 − h3i2 ) + g1 + Λ1 (8.274) ρ SdCNpα p dV2 ρ vSCD ρ SCLα 2 v i2 − vv2 cos α t  − v2 + =− (v1i3 − v3i1 ) dt 2m 2m 2m + ρ vSd ( CNq + CNαɺ ) 2m ( h3i1 − h1i3 ) + g 2 + Λ 2 (8.275) ρ SdCNpα p dV3 ρ vSCD ρ SCLα 2 v i3 − vv3 cos α t  − v3 + =− (v2i1 − v1i2 ) dt 2m 2m 2m + ρ vSd ( CNq + CNαɺ ) 2m ( h1i2 − h2i1 ) + g 3 + Λ 3 (8.276) Next is Equation 8.273, where the same procedure will be followed: dh1 ρ vSd 2ClP p ρ v 2Sdδ FClδ ρ vSdCMα i1 + i1 + = (v2i3 − v3i2 ) dt 2I T 2I T 2I T ρ vSd 2 ( CMq + CMαɺ )  ρ Sd 2CMpα p  IP p   [v1 − vi1 cos α t ] + +  h1 −   i1  2I T 2I T  IT    © 2014 by Taylor & Francis Group, LLC (8.277) Ballistics: Theory and Design of Guns and Ammunition 280 dh2 ρ vSd 2ClP p ρ v 2Sdδ FClδ ρ vSdCMα i2 + i2 + = (v3i1 − v1i3 ) dt 2I T 2I T 2I T + ρ vSd 2 ( CMq + CMαɺ )  ρ Sd 2CMpα p  IP p   [v2 − vi2 cos α t ] +  h2 −   i2  2I T 2I T  IT    (8.278) dh3 ρ vSd 2ClP p ρ v 2Sdδ FClδ ρ vSdCMα i3 + i3 + = (v1i2 − v2i1 ) dt 2I T 2I T 2I T ρ vSd 2 ( CMq + CMαɺ )  ρ Sd 2CMpα p  IP p   [v3 − vi3 cos α t ] + +  h3 −   i3  2I T 2I T  IT    (8.279) We can simplify Equations 8.274 through 8.279 considerably by defining the following coefficients: ɶ D = ρ vSCD C 2m 2 ɶl = ρ vSd ClP p C P 2I T ɶL = ρ SCLα C α 2m ρ ɶ N = SdCNpα p C pα 2m 2 ɶl = ρ v Sdδ FCl δ C δ 2I T ɶ M = ρ vSdCMα C α 2I T ɶN = C q 2 ɶ M = ρ Sd CMpα p C pα 2I T ρ vSd ( CNq + CNαɺ ) 2m ρ vSd ( CMq + CMα ) 2 ɶM = C q 2I T With these coefficients, we can write Equations 8.274 through 8.276 in a more compact form: dV1 ɶ Dv1 + C ɶL (v 2i1 − vv1 cos α t ) − C ɶ N (v3i2 − v2i3 ) = −C pα α dt ɶ N ( h2i3 − h3i2 ) + g1 + Λ1 +C q (8.280) dV2 ɶ D v2 + C ɶL (v 2i2 − vv2 cos α t ) − C ɶ N (v1i3 − v3i1 ) = −C pα α dt ɶ N ( h3i1 − h1i3 ) + g 2 + Λ 2 +C q (8.281) dV3 ɶ D v3 + C ɶL (v 2i3 − vv3 cos α t ) − C ɶ N (v2i1 − v1i2 ) = −C pα α dt ɶ N ( h1i2 − h2i1 ) + g 3 + Λ 3 +C q © 2014 by Taylor & Francis Group, LLC (8.282) Trajectories 281 We can do the same with Equations 8.277 through 8.279: ( )   IP p    h1 −   i1   IT    (8.283) ( )   IP p    h2 −   i2   IT    (8.284) dh3 ɶl + C ɶl i3 + C ɶ M (v1i2 − v2i1 ) + C ɶ M (v3 − vi3 cos α t ) + C ɶ M  h3 −  I P p  i3  = C α q  δ P pα  I   dt  T    (8.285) dh1 ɶl + C ɶl i1 + C ɶ M (v2i3 − v3i2 ) + C ɶ M (v1 − vi1 cos α t ) + C ɶM = C α q δ P pα dt dh2 ɶl + C ɶl i2 + C ɶ M (v3i1 − v1i3 ) + C ɶ M (v2 − vi2 cos α t ) + C ɶM = C α q δ P pα dt ( ) Now that we have the six, coupled, equations for our six accelerations, we would like to determine the position of the projectile in space and time. We do this by creating a vector, X, to the center of mass of the projectile. If we note that X = [xe1 + ye2 + ze3] in the earth-fixed coordinate system, then we can break the individual components into t ∫ x = x0 + V1dt (8.286) 0 t ∫ y = y0 + V2dt (8.287) 0 ∫ t z = z0 + V3dt (8.288) 0 Recognize that when firing a long-range weapon we usually do so with grid coordinates on a map of the earth. A map is, in theory, created by peeling the geometry off a globe. Thus, the coordinates and distances are correct in the downrange and cross-range directions (x and z). However, the altitude, y, has to be corrected for the curvature of the earth. This is depicted with the applicable equations in Figure 8.11. A similar rotation occurs with the gravity vector as depicted in Figure 8.12. With this relationship, we can write the projectile position vector in earth coordinates as  x 2     x2  x  [e e3 E ≈ [E1e1 + E2e2 + E3e3 ] =  xe1 +  y +  e2 + ze3  =  y +  1 R 2 2       z R   e3 ] (8.289) where R is the average radius of the earth, taken to be 6,951,844 yards or 6,356,766 m. The use of earth coordinates is recommended at ranges beyond about 2000 yards (at 2000 yards there is a 10.36-in. difference in height [1]). Furthermore, the acceleration of gravity varies with altitude (and, in fact, latitude and longitude as well) and we need to consider this. © 2014 by Taylor & Francis Group, LLC Ballistics: Theory and Design of Guns and Ammunition 282 θ 2 x = Rθ θ E 2 2R sin θ 2 θ R E2 cos θ ≈ E2 2R sin2 E2 ≈ E2cos θ = 2R sin2 θ 2 θ2 θ ≈ 2R 4 2 Rθ2 = R x2 2 2 R2 E2 ≈ x2 2R FIGURE 8.11 Altitude error over long trajectories. θ θ g g g cos θ = g cos 2 θ 2 = g 1−2sin2 x = Rθ θ g=g R g sin θ ≈ gθ = g x R y x e e + g 1−2 R 1 R 2 θ x2 =g 1− 2 2 2R x R y g=g 1 1−2 R 0 [ e1 e2 e3 ] g = 32.174 [1 – 0.0026 cos (2L)] ft2 s FIGURE 8.12 Rotation of the gravity vector due to earth curvature and associated equations. To complete the equations of motion, we must consider the form of the Coriolis acceleration vector. We have discussed this extensively previously so we shall simply write the components of this vector as Λ1 = 2Ω(− V2 cos L sin AZ − V3 sin L) (8.290) Λ 2 = 2Ω(V1 cos L sin AZ + V3 cos L cos AZ) (8.291) Λ 3 = 2Ω(V1 sin L − V2 cos L cos AZ) (8.292) Or as a vector  Λ1    Λ =  Λ 2  e1  Λ 3  e2  −V2 cos L sin AZ − V3 sin L    e3  = 2Ω V1 cos L sin AZ + V3 cos L cos AZ  e1   V1 sin L − V2 cos L cos AZ © 2014 by Taylor & Francis Group, LLC e2 e3  (8.293) Trajectories 283 We now have the differential equations of motion but need initial conditions to solve them. Let us examine the projectile at the instant of muzzle exit without worrying about how it attained its state of motion there (this is the job of the interior ballistician). We shall define the initial tube angle in azimuth and elevation as θ 0 and ϕ 0, respectively. Then our initial velocity vector can be defined as  Λ1  cos φ0 cos θ 0      V0 =  Λ 2  [e1 e2 e3 ] = V0  sinφ0 cos θ 0  [e1 e2 e3 ]  Λ 3   sinφ0  (8.294) And, if we also take the wind into account, we have  V10 − W10   v10      v 0 = V0 − W0 =  v20  [e1 e2 e3 ] = V20 − W20  [e1 e2 e3 ] V30 − W30  v30  (8.295) Here the usual relationships for these vectors apply. These are V0 = V120 + V220 + V320 (8.296) v0 = v120 + v220 + v320 (8.297) The initial orientations of the body-fixed unit vectors in the earth-fixed system are cos((φ0 + α 0 )cos(θ 0 + β 0 )  i10      i0 = i10 e1 + i20 e2 + i30 e3 =  i20  [e1 e2 e3 ] =  sin(φ0 + α 0 )cos(θ 0 + β 0 )  [e1 e2 e3 ]   i30  sin(θ 0 + β 0 ) (8.298)  − cos 2 (θ 0 + β 0 ) sin(φ0 + α 0 )cos(φ0 + α 0 )   j10   2  1   j0 = j10 e1 + j20 e2 + j30 e3 =  j20  [e] = cos (θ 0 + β 0 )cos 2 (φ0 + α 0 ) + sin 2 (θ 0 + β 0 ) [e1 e2 e3 ]  Q   j30   − sin(θ 0 + β 0 )cos(θ 0 + β 0 )sin(φ0 + α 0 )  (8.299) − sin(θ 0 + β 0 )  k10    1     0 k 0 = k10 e1 + k 20 e2 + k 30 e3 =  k 20  [e] =  [e1 Q  k 30  cos(θ 0 + β 0 )cos(φ 0 + α 0 ) e2 e3 ] (8.300) In the aforementioned equations, α 0 and β 0 are the initial pitch and yaw angles, respectively, of the projectile. Thus, they add directly to the weapon azimuth and elevation angles. The quantity Q we define following Ref. [1] as Q = sin 2 (θ 0 + β 0 ) + cos 2 (θ 0 + β 0 )cos 2 (φ0 + α 0 ) © 2014 by Taylor & Francis Group, LLC (8.301) Ballistics: Theory and Design of Guns and Ammunition 284 If we now consider the rotation, (ω)ijk, of the projectile about its axis of symmetry (thus relative to the i-j-k triad) and we define an arbitrary initial projectile rotation as (ω0 )ijk = ωi0 i0 + ω j0 j0 + ωk0 k 0 (8.302) Here this initial angular velocity is dependent upon the initial orientation of the unit vector, i0. Then, the initial velocity of the unit vector can be written as follows: i0 j0 di0 = (ω0 )ijk × i0 = ωi0 ω j0 dt 1 0 iɺ10    ωk0 = [ωk0 j0 − ω j0 k 0 ] = iɺ20  [e] ɺ  0 i30  k0 ωk0 j10 − ω j0 k10    = ωk0 j20 − ω j0 k 20  [e1 e2 e3 ]   ωk j3 − ω j k 3  0 0   0 0 (8.303) Note that Equation 8.303 is a tensor equation. Tensors are higher order vectors but can be treated the same. If we insert the results of Equations 8.299 and 8.300 into the aforementioned equation, we get 1 = ω j0 si n(θ 0 + β 0 ) − ωk0 cos 2 (θ 0 + β 0 )sin(φ0 + α 0 )cos(φ0 + α 0 )] iɺ10 = Q (8.304) 1 = ωk0 cos 2 (θ 0 + β 0 )cos 2 (φ0 + α 0 ) + ωk0 sin 2 (θ 0 + β 0 ) iɺ20 =  Q  (8.305) 1  −ω j0 cos(θ 0 + β 0 )cos(φ0 + α 0 ) − ωk0 sin (θ 0 + β 0 )cos(θ 0 + β 0 )siin(φ0 + α 0 ) iɺ30 = Q (8.306) Continuing with our statement of the initial conditions, a positive pitch rotates the nose of the projectile upward and a positive yaw rotates the nose to the left as viewed from the rear. The initial value of the modified angular momentum vector is given by I P p0 di   i0 +  i0 × 0  dt  IT  (8.307) di0 ɺ = i10 e1 + iɺ20 e2 + iɺ30 e3 dt (8.308) h0 = We can rewrite di0/dt as © 2014 by Taylor & Francis Group, LLC Trajectories 285 which then allows us to write e1 i0 × di0 = i10 dt iɺ10 e2 e3 i20 ( ) ( ) ( ) i30 = i20 iɺ30 − i30 iɺ20 e1 + i30 iɺ10 − i10 iɺ30 e2 + i10 iɺ20 − i20 iɺ10 e3 (8.309) iɺ20 iɺ30 We can then incorporate Equation 8.309 into Equation 8.307 to yield  I P p0 ɺ ɺ   I i1 0 + i20 i30 − i30 i20  T    h10    I p   h0 =  h20  [e] =  P 0 i2 0 + i30 iɺ10 − i10 iɺ30  [e1 e2 e3 ]   IT  h30     I P p0 i + i iɺ − i iɺ   I T 3 0 10 20 20 10  (8.310) Here the initial value of the spin rate p0 is determined by the axial velocity and the twist rate, n (in calibers per revolution), of the weapon through p0 = 2πV0/nd. We have thus completed all of the initial conditions necessary to perform the calculation. As we will discuss in a later section, the projectile’s motion can be characterized as epicyclic. The tip of a vector drawn from the CG of the projectile to the nose will trace out a curve that contains two cyclic modes: a fast mode, known as nutation, and a slow mode, known as precession. If the round is stable, these modes will eventually damp down to near zero, leaving only some movement because of nonlinear forces and moments. We shall explore this more later. Some other terms come up in the succeeding sections that require definitions. Since they are essential to the understanding of trajectories, we will define them now. A projectile’s yaw of repose is the yaw created by the action of gravity on the projectile as it attempts to follow its trajectory curve. As stated earlier, the nose of the projectile is usually above the trajectory. There is then a net aerodynamic force through the CP which wants to rotate the nose up. With a right-hand spinning projectile, this results in a yaw of the nose to the right. This is called the yaw of repose. Failure to trail is a situation that arises when the base of the projectile does not follow the nose (it flies base first after apogee). This is depicted in Figure 8.13. The trail angle is the quadrant elevation angle (particular to a gun, projectile, and charge combination) above which the projectile will not turn over and will fail to trail. We can summarize this section by saying that for a rigid projectile the 6-DOF model is as accurate as one can get to the trajectories. If the model yields an inaccurate answer, the problem is usually a wrong assumption in the metrology, initial conditions, or projectile mass properties. Lastly, the only practical method of solving these equations is by numerical methods and, with the speed of computers today, the codes run very efficiently and quickly. This last statement makes it difficult to generate meaningful problems for the interested reader. We have endeavored to create useful exercises by stipulating a large number of conditions and requiring the reader to examine the accelerations of the projectile at a point in space. © 2014 by Taylor & Francis Group, LLC Ballistics: Theory and Design of Guns and Ammunition 286 i y, 2 V αt j k mg mg V x, 1 z, 3 FIGURE 8.13 A projectile that has failed to trail. Problem 34 A British bomber is flying at a speed of 200 mph in still air. If the 0.303 in. machine guns are fired sideways, calculate the axial acceleration vector and the angular acceleration vector acting on the projectile through use of the 6-DOF equations if the projectile is 1. Fired to the right Answer:  ft  a = [ − 1376e1 + 5.89e2 − 1184e3 ]  2  s  dh  rad  = [39.85e1 − 35, 922e2 − 2.45e3 ]  2  dt  s  2. Fired to the left Answer:  ft  a = [ − 1376e1 + 70.29e2 − 1184e3 ]  2  s  dh  rad  = [39.85e1 + 35, 922e2 + 2.45e3 ]  2  dt  s  3. Discuss the effect of the angular momentum on the projectile nose (which way does it tip?) Please ignore the Coriolis acceleration, assume there is no yaw at muzzle exit and assume a muzzle velocity of 2440 ft/s, the weapon has a right-hand twist. © 2014 by Taylor & Francis Group, LLC Trajectories 287 Projectile information: C D 0 = 0.35 C 2 Dδ = 3.46 C M α = 2.36 (C (C ) ) = 0.003 Mq + C M αɺ = −16.2 I P = 0.00026[lbm-in.2 ] Nq + C N αɺ I T = 0.00258[lbm-in.2 ] C M P α = 0.02 m = 0.025[lbm]  lbm  ρ = 0.060  3   ft   rev  p = 2033   s  CL α = 2.81 CN Pα = −0.67 Please supply all answers in an inertial coordinate system labeled 1, 2, and 3 with 1 being along the aircraft flight path and 3 being off the right side of the plane. Treat all missing coefficients as equal to zero. Problem 35 One of the interesting aspects of the forces acting on a projectile occurs as the projectile leaves an aircraft sideways. This problem is encountered all the time in the AC-130 gunship. Let us examine a 105 mm HE projectile being fired into a city from both the top of a building and from the AC-130 in flight. The velocity of the projectile is 1510 ft/s. With the information provided 1. Calculate the total acceleration vector for both cases 2. Comment on the differences Positional information: 33.5° north latitude Azimuth of velocity vector: 80° True Angle of velocity vector to horizontal: −10° Wind is calm α = +2° (nose up), β = −1.5° (nose to the left looking downrange) The projectile nose is rotating to the right of the velocity vector at 0.5 rad/s The aircraft is flying at 300 mph to the north Projectile information: C D2 = 8.0 (C (C C M α = 3.80 C M p α = 0.05 m = 32.1[lbm]  lbm  ρ = 0.060  3   ft   rev  p = 200   s  C D 0 = 0.39 δ ) ) = 0.005 Mq + C M αɺ = −6.5 I P = 0.547 [lbm-ft 2 ] Nq + C N αɺ I T = 5.377 [lbm-ft 2 ] CL α = 1.9 CN p α = −0.01 Please supply all answers in an inertial coordinate system labeled 1, 2, and 3 with 1 being due north and 3 being due east. © 2014 by Taylor & Francis Group, LLC Ballistics: Theory and Design of Guns and Ammunition 288 Problem 36 The Paris gun was built by Germany in the First World War to shell Paris from 75 miles away. The weapon was a 210 mm diameter bore with the shells pre-engraved to compensate for wear of the tube. During firing of this weapon, all things such as wind effects, Coriolis, etc., had to be accounted for (they really could have used a good 6-DOF model and a computer). Write the acceleration vector for this projectile at an instant in its trajectory when the velocity (relative to the ground) is 2500 ft/s and the following conditions apply (please note that there is “no” rocket motor): Positional information: 48.75° north latitude Azimuth of velocity vector: 300° True Angle of velocity vector to horizontal: +10° Wind is blowing at 20 mph due south and horizontal α = 1°, β = 1.5° The projectile nose is rotating up at 2 rad/s Projectile information: C M α =3.50 (C (C CL α = 2.50 C M pα = 0.55 m = 220 [lbm] CN pα = − 0.02  lbm  ρ = 0.060  3   ft   rev  p = 150   s  C D = 0.28 ) ) = 0.005 Mq + C M αɺ = −16.5 I P = 19.13[lbm-ft 2 ] Nq + C N αɺ I T = 66.40[lbm-ft 2 ] Note that the aforementioned numbers are guesses at the projectile’s characteristics; they do not represent the real projectile’s performance as no data are available from any source researched. Please supply all answers in an inertial coordinate system labeled 1, 2, and 3 with 1 being due west and 3 being due north. Answer:  ft  Linear acceleration vector is a = { − 85e1 − 33e2 − 35e3 }  2  s  Angular acceleration vector is dh  rad  = { − 14e1 − 42e2 + 38e3 }  2  dt  s  Problem 37 An AC-130 is flying at a speed of 200 miles per hour in still air. If the 105 mm weapon is fired sideways, calculate the axial acceleration vector and the angular acceleration vector acting on the projectile through use of the 6-DOF equations if the projectile is fired to the left. Discuss the effect of the angular momentum on the projectile nose (which way does it want to tip?). Please ignore the Coriolis acceleration, assume there is no yaw at muzzle exit and assume a muzzle velocity of 1500 ft/s, the weapon has a right-hand twist. © 2014 by Taylor & Francis Group, LLC Trajectories 289 Projectile information: C D2 = 8.0 (C (C C M α = 3.80 C M p α = 0.05 m = 32.1[lbm]  lbm  ρ = 0.060  3   ft   rev  p = 220   s  C D 0 = 0.39 δ ) ) = 0.005 Mq + C M αɺ = −6.5 I P = 0.547 [lbm-ft 2 ] Nq + C N αɺ I T = 5.377 [lbm-ft 2 ] CL α = 1.9 CN p α = −0.01 Please supply all answers in an inertial coordinate system labeled 1, 2, 3 with 1 being along the aircraft flight path and 3 being off the right side of the plane. Treat all missing coefficients as equal to zero. Problem 38 An F-86 Saber jet flying with a speed of 600 mph. The aircraft is pulling up out of a dive and the pilot is experiencing 2.5 g′s. The pilot fires his 0.50 caliber machine guns at a ground target during the pull out. At the instant the plane makes an angle of 30° to the ground, a projectile leaves the muzzle of the weapon with a muzzle velocity of 1800 ft/s, 3° to the right of the bore axis and is rotating to the left at 2 rad/s in the plane of the gun, relative to the gun. Also the projectile is pitched up 1.5° and yawed left 1° to the line of fire. Ignoring Coriolis effects, write the vectors that completely define the initial conditions of the projectile so that you could use them in a 6-DOF model. DO NOT solve the equations of motion—just write the initial conditions. Assume coordinate directions as follows: 1. Horizontal direction in which the aircraft is flying 2. Straight up 3. To the right as viewed from the rear of the aircraft Problem 39 In the study of ballistics, we normally neglect Magnus forces as small and only the moments generated by these forces are significant. In low-velocity projectiles, however, Magnus forces may be considerable. Consider a baseball thrown with spin toward a batter. The ball weighs 5 oz and is 2.9in. in diameter. The distance from the pitcher’s mound to home plate is 60′ 6”. Assume the ball is thrown at 70 miles/h and that there is no wind. The release point is 4 ft above the ground with an upward angle of 4°. We need to determine how much spin is required to move the ball 2 ft to the left as viewed from the pitcher. Please perform the following calculations: a. Assuming constant drag and Magnus force coefficients, develop the equations of motion for the ball assuming flat fire is valid and the spin axis remains vertical for the entire flight. b. Determine the spin rate and direction to cause the desired motion. c. Determine the final velocity, time of flight, and height of the ball. d. Determine if the assumption of flat fire really was valid. © 2014 by Taylor & Francis Group, LLC Ballistics: Theory and Design of Guns and Ammunition 290 Please list all of your assumptions. The following coefficients may be assumed:  lbm  CD = 0.52 ρ = 0.076  3  CNpα = 0.09  ft  8.6 Modified Point Mass Trajectory The 6-DOF model’s equations, when fully developed, described an epicyclical motion with fast (nutational) and slow (precessional) modes that (hopefully) would damp out early on, allowing the projectile to assume a yaw of repose for the remainder of the flight. This yaw of repose, which remains nearly constant, we assume will account for most of the drag induced by the yawing of the projectile. If we can simplify the 6-DOF model, which is computationally expensive to run, by accounting only for the yaw of repose, we could get a model that will allow the projectile to drift the proper amount and still be quite accurate. Once again, following McCoy [1], we will make the point mass assumption in the equations that follow. Recall that because of this assumption the projectile is essentially represented as a cannon ball with all of its mass concentrated at one point. We shall then add some details, which will account for the yawing of the projectile, by assuming the projectile yaw is relatively constant or varies little with time compared to the steady state yaw angle. This assumption is usually valid except in high-angle fire situations. We begin with the usual equations of motion and Newton’s second law F = ma (8.311) particularized as m ∑ F + mg + mΛ dV = dt (8.312) Here the variables are the same as we described in the 6-DOF section. In the aforementioned equations, we replace the velocity vector V by the vector (V − W) because drag measurements are made relative to the air stream not relative to the ground. We will also again replace the scalar velocity (the speed) with the difference between the projectile and wind velocities: ɶ =|V − W| v = V−W →V (8.313) The diagram of the problem is shown in Figure 8.14. From our work on 6-DOF model, recall Equations 8.257 and 8.258, which we rewrite here neglecting the pitch damping and rocket forces: ρ SdC Npα  I T dV ρ vSCD ρ SCLα 2 =− v+ [v i − (v ⋅ i)v] − dt 2m 2m 2m  I P © 2014 by Taylor & Francis Group, LLC   (h ⋅ i)(i × v ) + g + Λ  (8.314) Trajectories 291 y Wy Height Wz Wx FD V0 j on c ti De f le k V mg 0 Range i x z FIGURE 8.14 Modified point mass trajectory. 2 dh ρ vSd Clp ρ v 2Sdδ FCl δ ρ vSdC Mα (h ⋅ i)i + i+ (v × i) = dt 2I P 2I T 2I T + ρ vSd 2 ( CMq + C Mαɺ ) ρ Sd 2C Mpα (h ⋅ i)[v − (v ⋅ i)i] + [h − (h ⋅ i)i] 2I P 2I P (8.315) Recall also from 6-DOF model (Equation 8.228) that introduces the polar and transverse moments of inertia I p di  d 2i  dh I P = pɺ i + P + i×  I T dt  dt 2  dt I T (8.316) Equations 8.314 and 8.315 may be simplified by introducing the tilde (∼) coefficients: ɶ D = ρ vSCD C 2m 2 ɶl = ρ vSd ClP p C P 2I T ɶL = ρ SCLα C α 2m ρ ɶ N = SdCNpα p C pα 2m 2 ɶl = ρ v Sdδ FCl δ C δ 2I T ɶ M = ρ vSdCMα C α 2I T ɶN = C q 2 ɶ M = ρ Sd CMpα p C pα 2I T ρ vSd ( CNq + CNαɺ ) 2m ρ vSd ( CMq + CMα ) 2 ɶM = C q 2I T Using this notation, the modified equations are written as dV ɶD v + C ɶL [v 2i − (v ⋅ i)v] − C ɶN  IT = −C pα  α dt  IP ( )   (h ⋅ i)(i × v) + g + Λ  dh ɶl + C ɶl i + C ɶ M (v × i) + C ɶ M (h ⋅ i)[v − (v ⋅ i)i] + C ɶ M [h − (h ⋅ i)i] = C p δ α pα q dt © 2014 by Taylor & Francis Group, LLC (8.317) (8.318) Ballistics: Theory and Design of Guns and Ammunition 292 or alternatively dV ɶD v + C ɶL [v × (i × v)] − C ɶ N (v × i) + g + Λ =−C α pα dt ( ) (8.319) dh ɶl + C ɶl i + C ɶ M (v × i) + C ɶ M (h ⋅ i)[v − (v ⋅ i)i] + C ɶ M [h − (h ⋅ i)i] = C q P δ α pα dt (8.320) We have shown earlier that since the unit vector, i, is always perpendicular to its derivative di/dt, the dot product of i and di/dt is identically zero. We shall combine Equations 8.316 and 8.320 to yield IP I di  d 2i  ɶ ɶ ɶ ɶ ɶ pɺ i + P p + i×  = ClP + Clδ i + CMα (v × i) + CMpα [i × (v × i)] + CMq IT I T dt  dt 2  ( ) di   i × dt  (8.321) We will now take the dot product of i with Equation 8.321 to yield ( ) ( IP ɶl + C ɶl → dp = I T C ɶl + C ɶl pɺ = C p p δ δ IT dt I P ) (8.322) Here use has been made of the facts that a cross-product results in a vector that is orthogonal to both of the original vectors and that the dot product of orthogonal vectors is identically zero. These relationships are written in mathematical terms here: di = i ⋅ (ω × i) = 0 dt i ⋅ i × (v × i) = 0 i⋅  di  i⋅i×  = 0  dt  i ⋅ (i × v) = i ⋅ (v × i) = 0 Equation 8.322 has an important consequence; for a rotationally symmetric projectile, the spin is decoupled from the yawing motion. Now, if we substitute Equation 8.322 into Equation 8.321, we get (Cɶ lp 2 ɶl i + I P p di +  i × d i  +C   δ I T dt  dt 2  ) ɶ M [i × (v × i)] + C ɶ M i × di  ɶl + C ɶl i + C ɶ M (v × i) + C = C pα q  δ α p  dt  ( ) (8.323) or I P di  d 2i  ɶ ɶ ɶ p + i × 2  = C Mα (v × i ) + CM pα [i × (v × i)] + CMq I T dt  dt   di  i × dt  (8.324) With Equations 8.319, 8.322, and 8.324, we have merely restated our 6-DOF model. Murphy [2] formulated the differential equation of motion as a second-order equation in terms © 2014 by Taylor & Francis Group, LLC Trajectories 293 j l sin αt αt i cos αt FIGURE 8.15 Projectile axial unit vector, l, illustrated. of complex variables and solved it. The particular solution was the (relatively) constant yaw of repose, and the complimentary solution was the transient epicyclic motion. In the modified point mass approach, we extract the particular solution and ignore the transient motion, instead concentrating on the yaw of repose, the drift, and the effect of the yaw drag. We assume that the epicyclic pitching and yawing motion are negligible everywhere along the trajectory, i.e., in many instances, reasonable, since it should damp early in the trajectory and thus contributes little to the drift. We proceed by defining another unit vector triad in the same sense as our i-j-k triad. Instead of it being aligned with the geometric axis of the projectile, we align it with the velocity vector and utilize l-m-n as the principal directions. We can then define l (Figure 8.15) as l= v v (8.325) and formally define our vector yaw of repose as α R = l × (i × l) (8.326) But we know that, where αt is the total angle of attack, l × (i × l) = (1)2i − (l ⋅ i)l = i − (1) (1) cos α t l (8.327) α R = l × (i × l) = i − (cos α t )l (8.328) so that in terms of αt For simplicity [1,2], if we choose the plane that l lies in to be the plane that j lies in as well, we can write Equation 8.328 as α R = i − cos α t (cos α ti + sin α t j) = (1 − cos 2α t ) i + sin α t cos α t j then α R = (1 − cos 2α t )2 + cos 2α t sin 2α t = © 2014 by Taylor & Francis Group, LLC sin 4α t + cos 2α t sin 2α t Ballistics: Theory and Design of Guns and Ammunition 294 and α R = sinα t sin 2α t + cos 2α t = sinα t We shall now differentiate Equation 8.328 with respect to time: dα R di dl = − (cos α t ) + sin α t l dt dt dt (8.329) We have made the assumption early in this analysis that the yaw of repose is relatively constant, thus (dαR/dt) ≈ 0. We also note that for a small yaw angle sin αR ≈ 0 ≪ cos αR. If we incorporate these approximations into Equation 8.329, we get di dl = (cos α t ) dt dt (8.330) Taking the time derivative of Equation 8.330 yields d 2i d2l dl α cos = ( ) − (sin α t ) t 2 2 dt dt dt (8.331) But the small angle approximation still applies so that d 2i d2l = (cos α t ) 2 2 dt dt (8.332) We can also solve Equation 8.328 for i to get i = α R + (cos α t )l (8.333) Since v and l are parallel and cross-products of parallel vectors are zero, we can write i × v = [α R + (cos α t ) l] × v = α R × v + (cos α t ) l × v = α R × v (8.334) v × (i × v ) = v × (α R × v) = v 2 α R − (v ⋅ α R ) v = v 2α R (8.335) and Also in a similar fashion operating on Equation 8.334, we can write v × i = v × [α R + (cos α t ) l] = v × α R + v × (cos α t ) l = v × α R (8.336) We can also show that i × (v × i) = v cos α tα R + v(sin 2 α t )l © 2014 by Taylor & Francis Group, LLC (8.337) Trajectories 295 We now have relations in Equations 8.330, 8.332, and 8.333 for i, di/dt, and d2i/dt 2, respectively, and can substitute them into Equations 8.319 and 8.324 to eliminate i. We shall start with Equation 8.319, and also noting that v × α R = (vl) × α R = v(l × α R ) (8.338) dV ɶ Dv + C ɶ L v 2α R + C ɶ N v(l × α R ) + g + Λ = −C α pα dt (8.339) we get It is also worth noting that dV d dV dl ɺ dl = l+V Vl = = Vl + V dt dt dt dt dt (8.340) We will now attack each term of Equation 8.324, but first we define γ = (l ⋅ i) = cos α t (8.341) Then this and the succeeding relations follow as dl dl I P di I P I ≈ p cos α t =γ P p p dt I T dt I T I T dt (8.342)  d 2i   d2l  2  d2l   i × 2  = γ αR × 2  + γ  l × 2  dt   dt    dt  (8.343) dl  2  dl   di   i × dt  = γ  α R × dt  + γ  l × dt  (8.344) Combining the terms of Equation 8.324, we get γ   d2l  ɶ I P dl d2l  2 ɶ p + γ αR × 2  + γ 2  l × 2  = C Mα v( l × α R ) + CM pα [vγα R + v sin α t l] I T dt d t d t     ɶ M γ  α R × dl  + γ 2  l × dl   +C q      dt   dt     (8.345) At this point we continue with our simplifying assumptions and neglect the Coriolis term in comparison with the gravitational acceleration term and also neglect the sin2 αt in comparison to γ. Thus, we can rewrite Equations 8.339 (including the relation of Equation 8.340) and 8.345 as dl ɶD v + C ɶL v 2α R + C ɶ N v(l × α R ) + g Vɺ l + V = −C α pα dt © 2014 by Taylor & Francis Group, LLC (8.346) Ballistics: Theory and Design of Guns and Ammunition 296 and γ   d 2l  ɶ I P dl d 2l  ɶ ɶ p + γ αR × 2  + γ 2  l × 2  = C Mα v( l × α R ) + CM pα vγα R + CM q I T dt dt    dt    dl  2  dl  γ  α R × dt  + γ  l × dt       (8.347) We shall now take the vector cross-product of l with Equations 8.346 and 8.347 and observe how each term behaves. First the LHS dl  dl  dl   = 0 +V l ×  l ×  Vɺ l + V  = l × Vɺ l + l × V dt  dt  dt   (8.348) Then each term on the RHS of Equation 8.346 ( ɶ D v ) = l × (−C ɶ Dvl) = −C ɶ Dv(l × l) = 0 l × (−C (8.349) ɶL v 2α R = C ɶL v 2 (l × α R ) l×C α α (8.350) ) ɶ N v (l × α R ) = −C ɶ N v[l × (l × α R )] = C ɶ N v[l × (α R × l)] l × −C pα pα pα ɶ N v[α R − (l ⋅ α R )l] =C pα (8.351) l× g = l× g (8.352) and Continuing on Equation 8.347, first the LHS l×γ I P dl I  dl  p = γ P p l×  I T dt I T  dt    d 2l  d 2l  l × γ αR × 2  = γ l × αR × 2  = γ dt  dt     d 2 l  d2l   l ⋅ 2  α R − ( l ⋅ α R ) 2  dt   dt    d2l    d 2 l  d 2 l   d2l  l × γ 2  l × 2  = γ 2  l ×  l × 2   = γ 2  l ⋅ 2  l − 2    dt    dt  dt   dt  (8.353) (8.354) (8.355) Then each term on the RHS of Equation 8.347 ɶ M v(l × α R ) = C ɶ M v[l × (l × α R )] = −C ɶ M v[α R − (l ⋅ α R )l] l×C α α α (8.356) ɶ M vγα R = C ɶ M vγ (l × α R ) l×C pα pα (8.357) © 2014 by Taylor & Francis Group, LLC Trajectories 297 and continuing ɶ M γ  α R × dl  = C ɶ M γ  l ×  α R × dl   l×C q q      d dt   t     ɶ M γ  α R × dl  = C ɶM γ l×C q q   dt    dl  dl dl  ɶ  l ⋅ dt  α R − (l ⋅ α R ) dt  = CMq γ (l ⋅ α R ) dt     (8.358) (8.359) and finally ɶ M γ 2 dl ɶ M γ 2  l × dl  = C ɶ M γ 2  l ×  l × dl   = C ɶ M γ 2 0 − (l ⋅ l) dl  = − C l×C q q q q     dt     d t t d dt        (8.360) We will now insert Equations 8.348 through 8.352 into Equation 8.346 to get  dl  ɶ 2 ɶ V  l×  = C Lα v ( l × α R ) − CN pα v[α R − ( l ⋅ α R )l] + (l × g)  dt  (8.361) We then do the same with Equations 8.353 through 8.360, inserting them into Equation 8.347 yielding γ  d 2 l   d 2 l  d 2 l  d2l  I P  dl  p  l ×  + γ  l ⋅ 2  α R − ( l ⋅ α R ) 2  + γ 2  l ⋅ 2  l − 2  dt  I T  dt   dt   dt  dt  ɶ M γ (l ⋅ α R ) dl + γ 2 dl  ɶ M v[α R − (l ⋅ α R )l] + C ɶ M vγ (l × α R ) − C = −C α pα q  dt dt   (8.362) We now have a pair of equations that essentially are comprised of two vector variables: the yaw of repose, αR, and the vector, l × αR. Cumbersome as it may seem, this is a linear system that can be solved readily through the use of matrices and their determinants to yield αR in terms of l and the other scalars and coefficients. The solution for αR after much manipulation is ɶ M vγ V  l × dl  − (l × g) −C pα  dt      2 2 ɶL v 2γ 2  l ⋅ d l  l − d l  + C ɶL v 2γ I P p  l × dl  + C ɶL C ɶ M v 2γ 2 dl +C  α q α α   2  2 dt I T  dt   dt  dt  αR = 2 ɶN C ɶ M v 2γ − C ɶL v 2γ  l ⋅ d l  + C ɶL C ɶM v3 C  pα pα α α α 2  d t   (8.363) We can further simplify this expression by returning to Equation 8.346 and taking the dot product of it with l. Through vector algebra and the use of the fact that (l · αR) = 0, we see that ɶ Dv + (l ⋅ g) = − ρ SCDv + (l ⋅ g) Vɺ = −C 2m 2 © 2014 by Taylor & Francis Group, LLC (8.364) Ballistics: Theory and Design of Guns and Ammunition 298 When this is substituted back into Equation 8.346, we get ɶ D v + (l ⋅ g)l + V −C dl ɶD v + C ɶL v 2α R + C ɶ N v(l × α R ) + g = −C α pα dt (8.365) which can be rewritten as V dl ɶ 2 ɶ N v(l × α R ) + g − (l ⋅ g)l = CLα v α R + C pα dt (8.366) and by neglecting the Magnus force term as it is small and noting that l × (g × l) = g − (l · g)l, we get V dl ɶ 2 = CLα v α R + [ l × (g × l)] dt (8.367) Remember that we wish to find a useful form with which we can calculate the quasisteady state yaw of repose as shown in Equation 8.363. That equation encompasses different vector functions and time derivatives of l, but our ultimate goal is to find expressions that only involve the measurable quantities of the aeroballistic coefficients, spin, gravity, and velocity and not the unit vector, l. To do this, though it may seem a devious process, we begin by taking the time derivative of Equation 8.367, getting dl d2l d + V 2 = 0 + [ l × (g × l)] Vɺ dt dt dt (8.368) We use Equation 8.364 and substitute it in Equation 8.368, arriving at 2 ɶ Dv dl + (l ⋅ g) dl + V d l = dg − (l ⋅ g) dl −  dl ⋅ g  l −  l ⋅ dg  l −C     dt dt dt 2 dt dt  dt   dt  (8.369) Noticing that dg/dt = 0 (if not, we really will have problems), we can rewrite this as V d 2l dl  d l  ɶ d l = −2(l ⋅ g) −  ⋅ g  l + C Dv dt 2 dt  dt  dt (8.370) If we examine the RHS of Equation 8.370 term by term and realize from Equation 8.367 that dl/dt can be solved for, then −2(l ⋅ g) dl 2 ɶL v 2α R + (l ⋅ g)g − (l ⋅ g)2 l  = − (l ⋅ g)C α  V dt (8.371) also 1 ɶ 2  dl  2 2  ⋅ g  l = CLα v (α R ⋅ g) + g − (l ⋅ g)  l V  dt  © 2014 by Taylor & Francis Group, LLC (8.372) Trajectories 299 and ɶ ɶ Dv dl = CDv C ɶL v 2α R + g − (l ⋅ g)l  C α  dt V  (8.373) Putting these expressions back into Equation 8.370 complicates things considerably, viz., V 2 ɶ 2 1 ɶ 2 d2l 2 = − C Lα v ( l ⋅ g )α R + ( l ⋅ g )g − ( l ⋅ g ) l  −  V CLα v (α R ⋅ g)l V dt 2 ɶ Dv dl + g 2l − (l ⋅ g)2 l  + C dt (8.374) ɶL , and drag, CD̃ , coefficients If we assume that the terms containing our modified lift, C α are either small with respect to the other terms or cancel one another, the third term in Equation 8.374 disappears and the others simplify to give eventually V2 d2l =  3(l ⋅ g)2 − g 2  l − 2(l ⋅ g)g dt 2  (8.375) Recalling that the triad (l, m, n) are unit vectors, with l in the direction of the velocity vector, v, we can write l = v/v′ and can transform Equation 8.367 into 1 dl 1  ɶ 2  = CLα v α R + 2 [ v × (g × v )] dt V  v  (8.376) Likewise our Equation 8.374, by dividing both sides by V, can be transformed to d2l 2 =− 2 2 dt V − 1 V2 1 1 ɶ  2 CLα v( v ⋅ g )α R + v (v ⋅ g) g − v 3 (v ⋅ g ) v  ɶ v dl ɶ  C g2 1 v − 3 (v ⋅ g )2 v  + D CLα v(α R ⋅ g )v + v v   V dt (8.377) Now, if we examine each term of Equation 8.363, we will have to deal with such terms  d2l  d2l dl d2l as l × , l ⋅ 2 , and  l ⋅ 2  l − 2 . We can perform all the substitutions of these terms dt dt  dt  dt with what we have shown in Equations 8.376 and 8.377, and then examine the resulting Equation 8.363 for terms that can be neglected for comparative magnitudes. When all the algebra is completed, the resulting equation, applicable to spinning projectiles [1], is dV   −2I P p  v ×  dt   αR = ρ Sdv 4CMα © 2014 by Taylor & Francis Group, LLC (8.378) 300 Ballistics: Theory and Design of Guns and Ammunition For nonspinning projectiles, we can simplify Equation 8.363 by removing the spin terms and neglecting terms of small magnitude. The resulting expression is αR = ɶ M γ 2 [ v × ( v × g )] C q ɶ M Vv 3 C (8.379) α The vector mechanics work out so that when there is a positive overturning moment (statically unstable projectile), the yaw of repose vector points to the right for a right-hand spin. The yaw of repose for a statically stable nonspinning projectile is such that the nose points slightly above the trajectory. Either Equation 8.378 or 8.379 can be inserted into Equation 8.361 and numerically integrated simultaneously with Equation 8.362 to yield the velocity and position at any time. This forms the basis of the modified point mass method. Problem 40 The Paris gun was built by Germany in the First World War to shell Paris from 75 miles away. The weapon was a 210 mm diameter bore with the shells pre-engraved to account for wear of the tube. During firing of this weapon, all things such as wind effects, Coriolis, etc., had to be accounted for. When the United States entered the war, the doughboys (the nickname for American troops) were to take the St. Mihiel salient where the gun was located. We shall assume that the Germans have turned the gun to fire on the Americans. The projectile is at some point in space defined later. To demonstrate your knowledge of the modified point mass equations 1. Draw the situation. 2. Calculate the vector yaw of repose for this projectile using Equation 8.326. Answer: αR = [0.008e2 + 0.002e3] [rad]. 3. Write the acceleration vector for this projectile using Equation 8.339 at the instant in its trajectory when the velocity (relative to the ground) is 2100 ft/s and the conditions given next apply. Note: You do not need all of the information given next. It is provided to you so you can compare the differences in formulations with the 6-DOF model. Answer: dV  ft  = [ −61e1 − 29e2 − 10e3 ]  2  dt s  4. Why we do not need to obtain the angular acceleration vector dh/dt? Positional information: 48° north latitude Azimuth of velocity vector: 190° True Angle of velocity vector to horizontal: +1° Wind is blowing at 15 mph due south and horizontal α = 0.5°, β = 0.25° The projectile nose is rotating down at 1 rad/s © 2014 by Taylor & Francis Group, LLC Trajectories 301 Projectile information: CMα = 3.50 (C (C CLα = 2.50 CMpα = 0.55 m = 220 [lbm]  lbm  ρ = 0.060  3   ft   rev  p = 130   s  CD = 0.28 ) ) = 0.005 Mq + CMαɺ = −16.5 I P = 19.13 [lbm-ft 2 ] Nq + CNαɺ I T = 66.40 [lbm-ft 2 ] CNpα = −0.02 C1P = −0.01 Supply all answers in an inertial coordinate system labeled 1, 2, and 3 with 1 being due south and 3 being due west. Problem 41 If we were to use a modified point mass assumption for both of the cases cited in Problem 35 1. Calculate the vector yaw of repose for both cases. Answer: αr = [0.0003e1 + 0.0060e2 − 0.0009e3] for the building αr = [−0.0009e1 − 0.0007e2 − 0.0000e3] for the gunship 2. Draw and explain what this vector represents. 3. Comment on whether this model is applicable for each case and why. 8.7 Probability of First Round Hit The theories describing the characteristics of exterior ballistic trajectories that we have just completed, while increasingly complex, do not deal with the problems of real-life gunnery. What an artillery commander or a tank commander wants to know is what the probability will be that the first round fired will land where desired. This is less of a problem for the indirect fire mode, where a spotter can call in corrective advice to the gunner, than it is in the direct fire mode, where the urgency of a first round hit may be a life and death matter. When you fire your weapon at a visible enemy, you may be subject to immediate return fire. So a first round hit with high probability of target destruction is of primary interest to gunner and commander. Predicted fire is what we are talking about, particularly tank gun versus tank target and the variables of the combat situation are what govern these probabilities. The advent of smart, even autonomous munitions, has, of course, reduced the problems of predicted fire, but at great monetary expense and the subjection of such munitions to countermeasures that negate their usefulness. The study of predicted fire is still a worthwhile exercise. Firing real projectiles from real weapon systems involves interactions among many variables. For example, the location (terrain) of the launching platform; the condition of the gun, its age, wear history, mounting method; aiming procedures including fire control; condition of the crew including fatigue, experience, fear of attack; weather conditions, © 2014 by Taylor & Francis Group, LLC 302 Ballistics: Theory and Design of Guns and Ammunition including temperature, humidity, air density and pressure; target location relative to the launching platform, target size, and not least, the characteristics of the projectile itself. Leaving these generalities, let us examine two of the areas where hit probabilities are generated: the proving ground and troop training locations. In the proving ground, great care can be taken to reduce, and even eliminate, some of the variables. For the ammunition designer, this is the opportunity to assess how precisely the design functions round to round: the gunner can take all the time he needs to aim the weapon—no one is shooting back at him. The weapon can be new or in the early part of its wear life; the propellant temperature can be closely controlled, reducing temperature effects on muzzle velocity variations. In some cases, special mounts can be built so that even large caliber cannon can be fired from what amounts to a bench rest similar to ones used with small arms. The object of such care is to find out the round-to-round precision (sometimes called dispersion) of the ammunition. This is not its accuracy, but is an important component of what is known as “first round hit probability.” The group (usually five or more projectiles made as identical as possible) fired under such controlled conditions clusters about an average point called the “center of impact” or CI. It is the predictability and reproducibility of the CI that assures the gunner and tank commander that when the first round of an engagement is fired it will go where it is intended. Rarely are second and third rounds fired at the same target in combat. The normal course of a development program sees multiple cycles of design, manufacture of test samples, test and evaluation until a stable design is achieved. These cycles are usually followed by a series of confirmatory tests that evaluate the performance of the stable design under varying conditions of temperature, gun wear, range performance, and a host of other tests. By normalizing the gun super-elevations of each group after firing, it is possible then to compare the CIs of the groups. At the proving ground, the azimuth or left-right setting of the gun would be closely maintained as a constant. This comparison speaks to the accuracy of the design, that is, the reproducibility of the CIs under varying conditions of proving ground firings. The precision (dispersion) of the design would also be found under these variations. How would this scatter of the CIs relate to accuracy under real-world combat situations? What could we then say about the probability of first round hit (PFRH) under combat conditions? To answer these questions, serious studies have been conducted in probabilistic terms about the conditions to be found around the world in combat in what are termed “Quasi-combat” conditions [6]. Extensive computer programs have been devised, as well, more recently, to compute PFRHs for designs of direct fire munitions under “Quasicombat” conditions [7]. Much of these theories and computations are dependent on the vehicle’s fire control system and data collections of conditions of terrain and atmospherics around the world. The original work in this area was done by H. Brodkin of the PittmanDunn Laboratory at the now defunct Frankford Arsenal in 1958 and remains classified. As shown by Christman [5], the PFRH is dependent on certain fixed biases, variable biases, and random errors. Hit probability computed this way is heavily dependent, as it should be, on both the type of fire control system employed and the flight properties of the projectiles being judged. For the predicted fire hit probabilities Christman computed in 1959, laser range finding with its very small ranging errors was not in use. Neither were the short times of flight of the fin-stabilized, small diameter but highly dense, armor defeating projectiles (APFSDS). They were not yet invented. The super-elevation of the gun, that is, the amount of elevation above the line of sight to the target that the gunner must use to lay the weapon, is greatly reduced when both the flight velocity is increased and the drag of the projectile is reduced with its consequently shortened time-of-flight. These factors have greatly aided the improved PRFH with the © 2014 by Taylor & Francis Group, LLC Trajectories 303 latest tank cannon and fire control systems now in use. It is nevertheless instructive to show the error sources considered in a stationary-to stationary engagement of a tank to a vertical 2.3 × 2.3 m target. Such computations are now carried out on computer programs similar to those shown in [6]. Some error sources are tabulated in Tables 8.1 and 8.2 [5]. As can be seen from Tables 8.1 and 8.2 there are many error sources that affect PFRH in a quasi-combat situation. Some of these require explanation. For example: Zeroing: The weapon is zeroed by firing a small group usually at 1200 m range, in a noncombat situation, recording the center of impact, CI, of the group and adjusting the optical fire control system so that it has a basis from which range-affected variables can be set. Once zeroed, the other fixed and variable biases and random errors remain to determine the PFRH at the time. Cant: This error, caused by the unevenness of terrain, causes a tilt of the weapon’s trunnion in both the roll and pitch attitudes and thus a consequent error in super-elevation (gun elevation above boresight) dictated by the fire control. Drift: This error is usually the result of the Coriolis effect, that is, the effect of the earth’s rotation under the trajectory causing a deviation in the CI. This is dependent on the latitude of the firing and target sites as well as the hemisphere (northern vs. southern hemispheres reverse the direction of the deviation). Fire control and parallax: These errors are the result of the slop in the mechanical linkages of the controls and the small offset in the alignment of the optical sight with the axis of the gun tube. TABLE 8.1 Horizontal Errors on First Rounds— Quasi-Combat Conditions Fixed biases Drift Fire control Jump, mean Parallax Variable biases Cant Fire control Jump variation Wind, cross Zeroing Cant Fire control Group center of impact Jump variation Observation of CI Wind, cross Random errors Round to round Dispersion Laying error © 2014 by Taylor & Francis Group, LLC Ballistics: Theory and Design of Guns and Ammunition 304 TABLE 8.2 Vertical Errors on First Rounds—Quasi Battle Conditions Fixed biases Fire control Jump Parallax (at 0 range) Range estimation Variable biases Jump variation Cant Fire control Muzzle velocity (lot to lot) Range estimation Zeroing Fire control Cant Group CI Jump Muzzle velocity (lot to lot) Observation of CI Range estimation Random errors Round to round dispersion Laying error Cross wind: A variable bias usually affecting the firing of different groups on different occasions Laying error: This is the result of random performance of gunners sighting in on targets and laying the gun from different directions occasion-to-occasion. Range estimation: This is a bias which is fixed for laser ranging but range dependent for optical ranging. Round-to-round dispersion: This is a random error dependent on the type of ammunition being fired. It also can vary from lot-to-lot in the same ammunition type due to muzzle velocity variations shot-to-shot in the lot. Jump: This is what is left over in the error budget when all other errors are accounted for. The test procedure of [7] meticulously lays out the testing procedures used in qualifying ammunition for service use. Appendix A of this reference provides a nomogram for computing PFRH. It also provides techniques for measuring some of the error sources mentioned. We must emphasize that this discussion only points out the complexity of the problem of computing PFRH. If the model by Bunn [6] is used to run such computations, it requires considerable detailed data input, which for modern tank, fire control, and ammunition designs is not available to general audiences. Furthermore, recall that the discussion was only for stationary-to-stationary situations where the opposing tanks are firing head on. Targets moving must provide lead data. Firing from moving vehicles with stabilized turrets as in the U.S. M1 Abrams tank, adds further error sources inherent in the stabilization system, but whose advanced computers and fire control systems may eliminate some of the other errors. © 2014 by Taylor & Francis Group, LLC Trajectories 305 References 1. McCoy, R.L., Modern Exterior Ballistics, Schiffer Military History, Atglen, PA, 1999. 2. Murphy, C.H., Free flight motion of symmetric missiles, Ballistics Research Laboratory Report No. 1216, Aberdeen Proving Ground, MD, 1963. 3. Jentz, T.L., Tank Combat in North Africa, the Opening Rounds, Schiffer Military History, Atglen, PA, 1998. 4. Litz, B., Applied Ballistics for Long Range Shooting, 2nd edn., Applied Ballistics, Cedar Springs, MI, 2011. 5. Christman, E.C., The effect of system design characteristics on first round hit probability of tank fired projectiles, Ballistics Research Laboratory Report No. 1192, Aberdeen Proving Ground, MD, 1959 (DTIC AD 316221). 6. Bunn, F.L., The tank accuracy model, U.S. Army Research Laboratory Report No. ARL-MR-A8, Aberdeen Proving Ground, MD, 1993 (DTIC AD 262812). 7. U.S. Army Armor and Engineer Board, Common Service Test Procedure, Round-to-Round Dispersion, U.S. Army Armor and Engineer Board, Fort Knox, KY, 1970 (DTIC AD 872085). Further Readings Capecelatro, A. and Salzarulo, L., Quantitative Physics for Scientists and Engineers—Mechanics, Auric Associates, Newark, NJ, 1980. Sears, F.W., Zemanski, M.W., and Young, H.D., University Physics, 6th edn., Addison-Wesley, Reading, MA, 1982. © 2014 by Taylor & Francis Group, LLC © 2014 by Taylor & Francis Group, LLC 9 Linearized Aeroballistics The aeroballistics topics discussed so far have built up to where the reader has an appreciation for the techniques required to analyze projectile motion to a great degree of accuracy. The culmination of this study was the development of the equations for a six degree-of-freedom (6-DOF) model, which accurately describes the motion of a rigid body through air. With a 6-DOF model in hand, the aeroballistician can examine the effects of a given configuration. The word given was italicized for emphasis because the aeroballistician must know the configuration properties before he or she analyzes the projectile. The implications of this are that without other tools to determine what needs to be changed in a design to alter the projectile behavior, one must simply guess at a new configuration, determine the aerodynamic coefficients, and reanalyze. This process can be very inefficient. The solution to this problem is to develop a theory that can be used to quickly determine what must be changed in a projectile to alter its flight behavior, make the changes, and reassess. This will be the topic for the remainder of this section. Linearized theory was (at least in the opinion of the authors) refined to an exceptional degree by Murphy [1] in 1963. Other authors before and since [2–4] have developed similar theories, and a good description of these can be found in McCoy [5]. Linearized theory is so named because the aerodynamic coefficients are assumed to be linear functions of the angle of attack. In other words, Fj ∝ C j0 (sin α t ) or M j ∝ C j0 d(sin α t ) (9.1) where the subscript j indicates any parameter of interest as introduced earlier. There are good points and bad points (as always) with this technique. The good news is that the mathematics become simple enough to determine quantities of interest extremely quickly and find means of changing a projectile’s flight characteristics quickly. The bad news is that the use of linear coefficients prevents us from duplicating some motions that occur frequently enough in projectile flight to warrant the inclusion of their nonlinear brethren— and the math becomes complicated to boot. We will continue the practice of using the definitions of the appropriate vectors and scalars based on Ref. [5]. The choice is somewhat arbitrary, but for several years now the authors have used this lucid work as a supplementary textbook and it is a matter of convenience. Our coordinate system is defined as in Figure 9.1. The aerodynamic coefficients introduced in the beginning of this chapter were written for both forces and moments as Fj = 1 ρV 2SC j 2 (9.2) Mj = 1 ρV 2SdC j 2 (9.3) 307 © 2014 by Taylor & Francis Group, LLC Ballistics: Theory and Design of Guns and Ammunition 308 y, j ≈ n CL Y, J α CN = CY x, i Cn I CD Z, K X, I V, I Cm α CX Projection of the velocity vector onto the x–z plane β z, k ≈ m CZ FIGURE 9.1 Coordinate system for projectile aerodynamic coefficients. We have also defined the angular rates of the projectile as p = Roll (spin) rate (9.4) q = Pitch rate (9.5) r = Yaw rate (9.6) The projectile angular position with respect to the velocity vector was given by α = Angle of attack (9.7) β = Angle of sideslip (9.8) The aerodynamic coefficients are functions of the rates expressed in Equations 9.4 through 9.6, as well as angular positions expressed in Equations 9.7 and 9.8. Additionally, these coefficients are also functions of the time rate of change of α and β that do not normally coincide with q and r. Thus, we can write C j = C j (α , β , αɺ , βɺ , p, q, r ) (9.9) With this nomenclature, any coefficient can normally be expressed as a series expansion in the seven variables  βɺ d   qd   pd   rd   αɺ d  C j = C j0 + C jα α + C jβ β + C jα   + C j p   + C jq   + C jr   + ⋯  + C jβ  V V V V V          (9.10) In Equation 9.10, we have included the terms in parentheses to maintain the nondimensional characteristics of the coefficient. We can see that this expansion results in a large number of terms that must be carried. Seldom in aeroballistics do we require terms in this expression beyond second order, but they can be included if data are available. When we discuss linear aeroballistics, we are limiting ourselves to the eight terms displayed in Equation 9.10. © 2014 by Taylor & Francis Group, LLC Linearized Aeroballistics 309 The linearization implies that C jk = ∂C j ∂k (9.11) k =0 Further simplifications will be made as we progress which will assist us in tackling the mathematics. We shall make use in this section of starred coefficients. These coefficients are defined in terms of their unstarred counterparts as C j*k = ρ Sd C jk 2m (9.12) 9.1 Linearized Pitching and Yawing Motions In the beginning of this chapter, we discussed terminology that allowed us to describe the pitching and yawing motion of a projectile. Because of the symmetry of typical projectiles, we combined pitch and yaw into a total yaw without any explanation. In this section, we will discuss the two motions separately and then formally make the assumptions that allowed us to combine them. This approach was formulated by Murphy [1] and what follows is basically that development with the coordinate system altered to fit our needs. If we have a projectile as depicted in Figure 9.1 and allow it only to move in a truly pitching motion, we can look down the z-axis and we would see what is depicted in Figure 9.2. Some interesting observations can be made from this figure. First, we see that the velocity vector, V, and the associated unit vector, l, are pitched up at angle ϕ to the earth-fixed coordinate system. The projectile is actually pointed above this angle by the pitch angle α. The vector along which the projectile is pointed is the geometric axis unit vector, x, and the spin (principal) axis unit vector, i. In an axially symmetric projectile, these are identical. We can see that through a rigid body rotation this forces the unit vectors of the transverse geometric axis, y, and transverse principal axis, j, to be rotated from the earth-fixed Y-axis through an angle of ϕ + α. If we assume that the projectile is constrained to pitch only, then the time rate of change of this total angle is q and the yaw angle and yaw rate are equal to zero as depicted in the figure. d( + α) =q dt β=0 y, j r=0 Y, J +α x, i α V, l X, I Earth referenced coordinate system FIGURE 9.2 Projectile in a pure pitching motion. © 2014 by Taylor & Francis Group, LLC Ballistics: Theory and Design of Guns and Ammunition 310 d(θ + β) =r dt α=0 q=0 X, I θ β z, k V, l x, i θ+β Z, K Earth referenced coordinate system FIGURE 9.3 Projectile in a pure yawing motion. In a similar manner, we can constrain our projectile to motion in the yaw plane only, which is depicted in Figure 9.3. In this case, the velocity and its associated unit vector are yawed with respect to the earth-fixed coordinate system by angle θ. The projectile geometric axis as well as the principal axis is yawed at angle β with respect to the velocity vector. This results in a rotation of the transverse principle axis from the earth-fixed coordinate system of θ + β as depicted in the illustration. The rate of change of this total angle is the yaw rate, r, and because of our constraints there is no pitching motion as identified in the figure. We shall now develop the equations of motion for each of these two specialized cases with the purpose of combining them in the end. For the purpose of this development, we shall define the force in the Y- and Z-directions using force coefficients CY and CZ, respectively. If we examine our projectile constrained to a pitching motion only, we can define the force coefficient as  qd   αɺ d  CY = CY0 + CYα α + CYαɺ   + CYq    V  V  (9.13) where we have restricted ourselves to the linear coefficients. We can see that this pitching motion causes a force in the Y-direction that is affected by angle of attack, rate of change of angle of attack, and pitching rate. An item worthy of note is that for a perfectly symmetrical projectile, CY0 would be zero. It is included here for completeness and can be present if an asymmetry exists. The corresponding moment for pitching motion only is given by  qd   αɺ d  Cm = Cm0 + Cmα α + Cmαɺ   + Cm q   V  V  (9.14) where the same comments about the nondimensionalization and Cm0 apply as well. Now we will examine the equations of motion. The force and moment equations are given by © 2014 by Taylor & Francis Group, LLC F = ma (9.15) M = Iαɺ (9.16) Linearized Aeroballistics 311 If we define angles ϕ̂ and θ̂ as φˆ = φ + α (9.17) θˆ = θ + β (9.18) then our scalar equations of a projectile in flight exhibiting pure pitching motion are dVx dV ≈m = Fx dt dt (9.19) d 2Y = FY cos φˆ + Fx sin φˆ − mg dt 2 (9.20) m m d 2φˆ MZ = dt 2 IZ (9.21) If we examine a small time of the projectile flight, we can assume constant velocity. If we further limit the pitching motion to small angles, we can assume Fx = − FD ≈ 0 (9.22) cos φ̂ ≈ 1 (9.23) sin φˆ ≈ φˆ (9.24) These assumptions can be used in Equations 9.19 and 9.20 to yield dV =0 dt (9.25) d 2Y = FY − mg dt 2 (9.26) m m We know that FY = 1 ρV 2SCY 2 (9.27) If we then substitute Equations 9.13 and 9.27 into Equation 9.26, we obtain m  d 2Y 1 = ρV 2S CY 0 + CYαα + CY αɺ 2 2 dt  © 2014 by Taylor & Francis Group, LLC  qd    αɺ d    + CYq    − mg  V   V  (9.28) Ballistics: Theory and Design of Guns and Ammunition 312 or, using our definition of starred coefficients, we have d 2Y V 2  *  qd    αɺ d  CY0 + CY*α α + CY*α  =  + CY*q    − g  2 d  dt  V   V  (9.29) Equation 9.29 can be combined with Equation 9.21 to develop a single equation for projectile motion. With this, the dynamic equation for the pure pitching motion of a projectile can then be described as ˆ + Gˆ ˆ 1α = A αɺɺ + Hˆ 1dαɺ − M 1 d (9.30) This linear, second-order differential equation with constant coefficients was established by Murphy [1] and modified here (the terms with the “d” subscript) to account for the assumption of zero drag. In this expression, we identify the coefficients as follows: ) ( 1  * αɺ   V  Ĥ1d = − CY*α + 2 Cm* q + Cm  d k Z     1 *  V  M̂1 =  2 Cm α    kZ  d  ˆ =  1 C*   V  A 1  k 2 m0   d   Z   2 (9.32) 2 (9.33)  1 *  g  Ĝd = −  2 Cm q    kZ  d  k Z2 = (9.31) (9.34) IZ md 2 (9.35) If we include drag (and thus ignore Equation 9.22) yet leave all of the other assumptions in place, we obtain a result identical to Murphy [1]. This results in Equations 9.30, 9.31, and 9.34 being modified to ˆ + Gˆ ˆ 1α = A αɺɺ + Hˆ 1αɺ − M 1 ( (9.36) ) 1  * q + Cm * αɺ   V  Ĥ1 = − CY*α + CD* + 2 Cm  d kZ    (9.37)  1 * *   g  Ĝ = −  2 Cm q − CD    d   kZ (9.38) © 2014 by Taylor & Francis Group, LLC Linearized Aeroballistics 313 It is more convenient to examine the differential Equations 9.30 and 9.36 with dimensionless distance (defined as s/d) instead of time as the independent variable. The time derivatives of dimensionless distance can then be written as ds  V  =  dt  d  (9.39) d 2s  Vɺ  =  dt 2  d  (9.40) d d ds  V  d () = () =   () dt ds dt  d  ds (9.41) 2 2  Vɺ  d d2 V  d () = () +   ()   2 2 dt  d  ds  d  ds (9.42) and With this, we can use the relations to rewrite Equations 9.30 and 9.36, respectively, as α ′′ + H1dα ′ − M1α = A1 + Gd (9.43) α ′′ + H1α ′ − M1α = A1 + G (9.44) The coefficients in these equations are given by ) ( 1  * q + Cm * αɺ  H1d = − CY*α + 2 Cm kY   ( (9.45) ) 1  * q + Cm * αɺ  H1 = − CY*α + 2CD* + 2 Cm k Z   M1 = 1 * Cmα k Y2 (9.47) A1 = 1 * Cm 0 k Z2 (9.48)  1 *   gd  Gd = −  2 Cm q  2   kZ   V0  (9.49)  1 * *   gd2  G = −  2 Cm q − CD   k Z   V0   (9.50) where V0 is the muzzle (or a reference) velocity of the projectile. © 2014 by Taylor & Francis Group, LLC (9.46) Ballistics: Theory and Design of Guns and Ammunition 314 A similar procedure can be followed to define motion constrained to the yaw plane only. This gives the result (details covered in Ref. [1]) of β ′′ + H 2d β ′ − M2 β = A2 (9.51) β ′′ + H 2 β ′ − M2 β = A2 (9.52) The coefficients in these equations are given by ( ) 1   H 2d = − CZ*β + 2 Cn*r + Cn*βɺ  k Z   ( ) 1   H 2 = − CZ*β + 2CD* 2 Cn*r + Cn*βɺ  kY   (9.53) (9.54) M2 = 1 * Cm β k Y2 (9.55) A2 = 1 * Cn 0 k Y2 (9.56) IY md 2 (9.57) k Y2 = Assuming a projectile is axially symmetric implies that any plane orthogonal to the polar axis is a principal axis. This forces the two transverse moments of inertia to be equal and, with an assumption of small yaw, allows us to write IT = Iy = Iz ≈ IY = IZ (9.58) This symmetry also allows us to equate the pitch and yaw coefficients. Thus, we define CNα ≡ CYα = CZβ (9.59) C M q ≡ Cm q = C n r (9.60) CMα ≡ Cmα = Cn β (9.61) CMαɺ ≡ Cmαɺ = Cn βɺ (9.62) Complex numbers are commonly used to define pitch and yaw angles. This is extremely convenient because it allows us to collapse two differential equations into one. We shall define the complex yaw angle as ξ, which shall be defined thus ξ ≡ α + iβ © 2014 by Taylor & Francis Group, LLC (9.63) Linearized Aeroballistics 315 α iβ FIGURE 9.4 Complex yaw plane. This definition allows one to look downrange as a projectile flies along a trajectory and visualize the imaginary part of the equation affecting the yaw of the projectile and the real part of the equation as affecting pitch. This is illustrated in Figure 9.4. In this figure, the origin is the trajectory of the projectile looking downrange. The two differential equations of motion Equations 9.44 and 9.52 can then be combined by first multiplying Equation 9.52 by the imaginary number, i, and adding them together. This results in V  ξ ′′ + Hξ ′ − Mξ = A + G  0  V  2 (9.64) The coefficients in this equation are given by ( ) 1  * q + CM * αɺ  H = − CN* α + 2CD* + 2 CM  k T   M= A= 1 *α CM kT2 ( 1 * 0 + iCn*0 Cm kT2 (9.65) (9.66) )  gd   1 * * 2 G = −  2 CM q − CD    kZ   V0  (9.67) (9.68) The solution to Equation 9.64 can be found for a nonspinning projectile to be [1] ξ = K1 exp(iψ 1 ) + K 2 exp(iψ 2 ) + K 3 exp(iψ 30 ) + ξ g (9.69) In this equation, each term Kj is known as an arm to be described subsequently. Mathematically, we can express these terms as K j = K j0 exp(λjs) © 2014 by Taylor & Francis Group, LLC (9.70) 316 Ballistics: Theory and Design of Guns and Ammunition Here we see that each arm is a function of its initial value (that occurring at the muzzle of the gun) and an exponential damping term. The exponential damping term decides whether the amplitude of the motion will decay, grow, or remain constant. The damping terms are given by [1,5] 1 λ1 = λ2 = − H 2 (9.71) The exponential terms in Equation 9.69 contain phase angles, ψ j. These phase angles represent the instantaneous angle that each arm makes with the imaginary axis. These can be written in terms of their initial value and a turning frequency as ψ j = ψ j0 + ψ ′js (9.72) The turning frequencies are given for a nonspinning projectile by [1,5] ψ 1′ = −ψ 2′ = − M (9.73) The third term on the RHS of Equation 9.69 is the so-called trim arm. This is a measure of the amount that a fin-stabilized projectile will trim (i.e., fly with constant pitch or yaw) during flight. It is given by K 3 exp(iψ 30 ) = − i(Cm0 + iCn0 ) CMα (9.74) The fourth term on the RHS of Equation 9.69 is the yaw caused by interaction of the projectile with the gravity vector, sometimes called the yaw of repose. It is defined as  gd  i CMq − kT2CD  2  V  ξg = CMα ( ) (9.75) To visualize the physical meaning of Equation 9.69, we shall imagine we have a projectile and we are looking downrange along the trajectory such that the complex plane lies perpendicular to the trajectory curve. Our projectile will be at some arbitrary yaw angle. This is depicted in Figure 9.5. We need to note that the arms usually do not point to the nose of the projectile, they point to the symmetry axis; however, it is easiest to visualize the situation by scaling them to point to the nose. Imagine that we follow the projectile depicted in Figure 9.5 as it traverses the trajectory. We would see the nose motion swirling around. Throughout this time, we would also see the length of each of the arms changing (growing, decaying, or remaining the same) as dictated by Equation 9.69. Additionally, we would see the arms rotating around their respective origins at rates described by Equation 9.72. All through this time, our viewpoint would be changing because we have our gaze fixed on the complex plane and it is rotating into the paper because of the curvature of the trajectory. In the development of Equation 9.64 and its solution Equation 9.69, the spin of the projectile was neglected. Because of this, these equations are specific to fin- or drag-stabilized projectiles that have relatively small spin rates. References [1,2,5] develop the equation of © 2014 by Taylor & Francis Group, LLC Linearized Aeroballistics 317 α (pitch) K2 g Flight path 2 1 K3 K1 iβ (yaw) FIGURE 9.5 Example of tricyclic arms. motion for spinning projectiles in exactly the same manner. The results essentially incorporate the third angular component known as the roll or spin. The differential equation for a spinning projectile is given by ξ ′′ + ( H − iP)ξ ′ − ( M + iPT )ξ = −iPG (9.76) In this formulation, we can utilize axial symmetry and, thus, define our coefficients as follows: H = CL*α − CD* − M= ( 1 * q + CM * αɺ CM kT2 ) 1 *α CM kT2 T = CL*α + G= I P= P  IT 1 * pα CM kP2 (9.77) (9.78) (9.79) gd V02 (9.80)   pd   V    (9.81) The solution to Equation 9.76 is ξ = K10 exp[λ 1s]exp[i(ψ 10 + ψ 1′s)] + K 20 exp[λ 2s] exp[i(ψ 20 + ψ 2′ s)] + ξ g (9.82) This equation is essentially the same form as Equation 9.69 except for the deletion of the trim arm. It is also noteworthy that we have expanded the slow and fast arm terms © 2014 by Taylor & Francis Group, LLC Ballistics: Theory and Design of Guns and Ammunition 318 α (pitch) ΄ K1 2 ΄1 K2 iβ (yaw) g FIGURE 9.6 Example of tricyclic arms for fin-stabilized projectile. and exponents to display their exponential behavior. The expression is also commonly written as ξ = K1 exp(iψ 1 ) + K 2 exp(iψ 2 ) + ξ g (9.83) where the definitions of Equations 9.70 and 9.71 apply. The λj terms are known as the exponential damping coefficients and the ψ j terms are the precessional and nutation frequencies of the projectile. These are commonly defined as a complex pair where λ1, 2 + iψ 1, 2 = 1 − H + iP ± 4 M + H 2 − P 2 + 2iP(2T − H )   2  (9.84) As a parting note, we need to discuss the behavior of the fast and slow arms and the associated motion that they undergo. For a nonspinning projectile, we shall examine Equation 9.73. In this expression, the sign of M is important. For a nonspinning projectile M is negative. That tells us that the arms turn in opposite directions with K1 being positive (clockwise) and K 2 negative (counter-clockwise). This is depicted in Figure 9.6. Likewise for a spinning projectile, we need to examine the derivative with respect to s of Equation 9.84. In this case, we would find that ψ 1′, 2 = 1( P ± P2 − 4M ) 2 (9.85) Here we shall see in the following section that for stability this must result in a solution that has no imaginary part. So both values of the root will have the same sign, thus the two arms turn in the same direction as shown in Figure 9.7. Initial conditions that are present when the projectile leaves the muzzle of the weapon are important as our starting point for the values of the fast and slow arms. These can even cause drastically different flight behavior when nonlinear coefficients are introduced later. The initial sizes of the fast and slow arms can be expressed as functions of the precession and nutation rates, the damping exponents, and the initial yaw and yaw rates [5] as K10 exp(iψ 10 ) = © 2014 by Taylor & Francis Group, LLC ξ 0′ − ( λ2 + iψ 2′ ) ξ0 λ1 − λ2 + i (ψ 1′ −ψ 2′ ) (9.86) Linearized Aeroballistics 319 α (pitch) K1 K2 ΄ 1 ΄ 2 iβ (yaw) g FIGURE 9.7 Example of tricyclic arms for spin-stabilized projectiles. K 20 exp(iψ 20 ) = ξ 0′ − ( λ1 + iψ 1′ ) ξ 0 λ2 − λ1 + i (ψ 2′ −ψ 1′ ) (9.87) Because the damping exponents are usually an order of magnitude or more smaller than the precession and nutation rates, these equations can be simplified to K10 exp(iψ 10 ) = iξ 0′ + ψ 2′ξ 0 ψ 2′ −ψ 1′ (9.88) K 20 exp(iψ 20 ) = iξ 0′ + ψ 1′ξ 0 ψ 1′ −ψ 2′ (9.89) These equations are important because they allow one to determine the initial amplitudes of the arms given an assumed or measured initial yaw, yaw rate, and muzzle exit conditions for a known projectile geometry. The expressions introduced in this section are the basis for stability criterion to be established next. In the next section, we shall discuss the behavior of these equations and use them to define stability criteria for a projectile. Problem 1 A 155-mm M549A1 projectile has the following properties and initial conditions: CD = 0.3 CLα = 0.13 CMα = 4.28 Clp = −0.024 CMq + CMαɺ = −26 CMpα = 0.876 © 2014 by Taylor & Francis Group, LLC  lbm  ρ = 0.0751  3   ft  I P = 505.5 [lbm-in.2 ] d = 155 [mm] I T = 6610 [lbm-in.2 ]  ft  Vmuzzle = 3000   s m = 96 [ibm] 320 Ballistics: Theory and Design of Guns and Ammunition At an instant in time after launch when p = 220 [Hz] φ = δ = 4°  ft  V = 1764   s Determine 1. The yaw of repose Answer: β R = 0.00172 [rad] 2. The precessional frequency in Hz Answer: dψ 2 = 1.9 [Hz] dt 3. The nutational frequency in Hz Answer: dψ 1 = 14.9 [Hz] dt Problem 2 For the projectile given in Chapter 8, problem 37, determine the precessional and nutational frequencies in Hertz. 9.2 Gyroscopic and Dynamic Stabilities In the previous section, we developed a pair of equations and their solutions using linear aeroballistic coefficients that allow us to examine the motion of a projectile in pitch, yaw, and roll. These equations will now be examined in detail so that we can establish criteria for a stable projectile. In so doing, we will examine some interesting characteristics of motion, which will be displayed as curves in the complex plane. We shall repeat the equations and their solutions here for ease of reference but leave the coefficient definitions in Section 9.1 to preserve space. The governing equations are as follows: For a nonspinning or slowly spinning projectile, V  ξ ′′ + Hξ ′ − Mξ = A + G  0  V  © 2014 by Taylor & Francis Group, LLC 2 (9.90) Linearized Aeroballistics 321 with the solution ξ = K10 exp[λ1s]exp i (ψ 10 + ψ 1′s )  + K 20 exp[λ2s] exp i (ψ 20 + ψ 2′ s )  + K 3 exp[iψ 30 ] + ξ g (9.91) For a spinning projectile, ξ ′′ + ( H − iP)ξ ′ − ( M + iPT )ξ = −iPG (9.92) ξ = K10 exp[λ1s] exp i (ψ 10 + ψ 1′s )  + K 20 exp[λ2s] exp i (ψ 20 + ψ 2′ s )  + ξ g (9.93) with the solution For our general development of stability, we shall focus on Equation 9.92 and its solution (Equation 9.93), since the trim term in Equation 9.91 can be easily dealt with separately. If we examine Equation 9.93, we can readily see that nasty things can happen to us mathematically because of the exponential terms. Since K10 and K 20 are constants (they are the initial magnitudes of the fast and slow arms, respectively), we can focus on the exponential terms that they are multiplied by as a means of determining whether they will grow, shrink, or remain the same. We shall consider the exponential functions of ψ and ψ′ first using the fast arm terms as examples. The term ψ 10 is a constant and will be ignored. This leaves the term ψ 1s, which is multiplied by i in the exponent. If ψ 1′ is purely real, then, when multiplied by i, it becomes purely imaginary in the exponent (because s must be real), the solution is oscillatory and this will cause the fast arm to increase and decrease in amplitude (i.e., oscillate), neither increasing nor decreasing beyond the established limits of oscillation. This would be a gyroscopically stable projectile. If it has an imaginary component, then, when multiplied by i in the exponent, the solution has a real part. This real part will be multiplied by s and continue to grow throughout the flight as s continually increases. This would result in a gyroscopically unstable projectile. The question to answer at this point is “What governs whether the exponents have real or imaginary parts?” This can be answered by examination of a version of Equation 9.84, whereby all aerodynamic forces and moments are ignored except for the largest (pitching) moment. This has been shown [1,5] to result in a governing equation of ξ ′′ − iPξ ′ − Mξ = −iPG (9.94) ξ = K1 exp i (ψ 10 + ψ 1′s )  + K 2 exp i (ψ 20 + ψ 2′ s )  + ξ g (9.95) with the solution resulting in ψ 1′, 2 = ( 1 P ± P2 − 4M 2 ) (9.96) where the subscripts 1 and 2 represent the fast and slow arms, respectively. Using Equation 9.96, we recall that for a gyroscopically stable projectile, ψ′ must be real, therefore for gyroscopic stability, we require that (P 2 − 4 M ) > 0 © 2014 by Taylor & Francis Group, LLC (9.97) 322 Ballistics: Theory and Design of Guns and Ammunition This expression has some interesting implications. If we look back at the definition of our parameter, M in Equation 9.66, we see that it is dependent upon the pitching moment coefficient. This happens to always be negative for a fin-stabilized projectile since the fins impart a restoring moment. Unless there is some unique drag device, this moment is positive in a non-fin-stabilized projectile. Because of this, a fin-stabilized projectile is always gyroscopically stable because P2 must be positive. However, a non-fin-stabilized projectile must have a spin sufficient to make P2 > 4M. We therefore define a statically stable projectile as one in which M < 0. With this definition, a statically stable projectile is always gyroscopically stable. Gyroscopic stability is a necessary but not sufficient condition for a stable projectile. The second condition required is that of dynamic stability. Let us once again examine Equation 9.93, but this time we shall assume that we have a gyroscopically stable projectile. This means that the exponential terms containing ψ′ decay or remain constant, leaving the terms containing λ as potentially destabilizing. We can readily see that, since these are multiplied by the downrange distance, s, they must be negative to assure that the fast and slow arms decay in magnitude. With this, we shall define a dynamically stable projectile as one in which both λ’s are negative throughout the flight. Recall that we calculate λ as the real part of Equation 9.84. For convenience, we shall express them directly as 1 P(2T − H )  λ1, 2 = −  H ∓  2 P2 − 4M  (9.98) It should be noted here that, as is common in ballistics, there are always exceptions to any rule. Some successful projectiles have been fielded where instability occurs for a very short time in a flight or in a range where a certain projectile will never be fired. Of course, it is always best to avoid these situations but sometimes lack of design space makes it unavoidable. In these instances, rational examination of the instability is necessary and should be well documented. We have shown mathematically how we define stability and the parameters that affect stability. Sometimes, it is desirable to quantify how stable a projectile is. We do this through use of a gyroscopic and dynamic stability factors. We define the gyroscopic stability factor as Sg = P2 4M (9.99) Here, with our earlier discussion, Sg > 1 to assure gyroscopic stability. In a similar fashion, we can define a dynamic stability factor as Sd = 2T H (9.100) where for a symmetric projectile to be deemed stable, whether spinning or nonspinning, we require 1 < Sd (2 − Sd ) Sg © 2014 by Taylor & Francis Group, LLC (9.101) Linearized Aeroballistics 323 For a statically stable projectile, we require that 0 < Sd < 2 for dynamic stability. This leads to an interesting condition where one can spin a statically stable projectile too fast, resulting in instability. This condition translated to dimensionless spin rate is given by P< 4M Sd (2 − Sd ) (9.102) for a statically stable projectile. It is interesting to combine Equations 9.96 and 9.98 in various ways writing them in terms of the dimensionless parameters P, M, H, and T. The details of this can be found in Refs. [1,5] with the result P = ψ 1′ + ψ 2′ (9.103) M = ψ 1′ψ 2′ − λ1λ2 (9.104) H = −(λ1 + λ 2) (9.105) PT = − (ψ 1′λ 1 + ψ 2′ λ 2 ) (9.106) If we again examine Equation 9.91 or 9.93, we see that the magnitude of the precessional and nutational arms is highly dependent upon initial conditions. Without going into details (which are described quite well in Ref. [5]), we can express these initial conditions in terms of the complex angle of attack and damping parameters as K10 exp[iψ 10 ] = ξ 0′ − ( λ2 + iψ 2′ ) ξ0 λ1 − λ2 + i (ψ 1′ −ψ 2′ ) (9.107) K 20 exp[iψ 20 ] = ξ 0′ − ( λ1 + iψ 1′ ) ξ 0 λ2 − λ1 + i (ψ 2′ −ψ 1′ ) (9.108) In these equations, ξ0 and ξ 0′ are the initial complex yaw and yaw rates, respectively. These parameters are determined by measurements as the projectile leaves the gun tube or are assumed values. We now have solid criteria by which we can determine whether a projectile will be stable or not. These developments have been made assuming that the projectile aerodynamic coefficients behave in a linear fashion. As such, a projectile is either stable or it is not. This stability, even with our linear model, will change during the flight based on Mach number and angle of attack. We will discuss in a later section how a nonlinearity can help or hurt matters. The true power of these equations is that they can tell us which coefficients need to be altered to affect stability. This can be used in instances where we want to change a physical configuration to make a projectile “drop out of the sky” or design a round such that it damps more quickly and thus can fly with lower drag. Other uses for these equations allow for tweaking the flight characteristics for better flight behavior in general. © 2014 by Taylor & Francis Group, LLC Ballistics: Theory and Design of Guns and Ammunition 324 Problem 3 Up until the late 1960s, many U.S. and foreign ships carried the Bofors 40 mm gun as a general light support weapon. Originally designed as an antiaircraft weapon, this gun served as an antitank weapon if the situation required it and its high rate of fire made it quite successful as an antipersonnel weapon. Assume the properties of the system are given next: 1. Calculate the gyroscopic stability factor at the beginning and at the end of the flight assuming a terminal velocity of 2450 ft/s and the spin rate is 10% lower than the initial value. Answer: At the beginning of flight Sg = 5.814 2. Is the projectile stable throughout the flight? Answer: Yes 3. Assuming that this is the longest time of flight for the projectile, at what spin rate will the projectile become unstable? Answer:  rad  punstable < 1887   s  4. Where will the instability occur? Answer: At the muzzle of the weapon Projectile and weapon information  lbm  ρ = 0.067  3   ft  CMα = 3.10 Clp = − 0.011 d = 40 [mm]  ft  Vmuzzle = 2850   s I P = 1.231 [lbm-in.2 ] I T = 6.263 [lbm-in.2 ] m = 1.985 [ibm] n= 1  rev  30  cal  Please note that this weapon actually has a progressive twist, but when faced with this situation you only need the muzzle velocity and the twist at the muzzle to calculate initial spin. Problem 4 For the projectile described in Problem 35 of Chapter 8, 1. Determine the precessional damping exponent. Answer: λ2 = −0.0000608 2. Determine the nutational damping exponent. Answer: λ1 = −0.001135 © 2014 by Taylor & Francis Group, LLC Linearized Aeroballistics 325 3. With (1) and (2) above which of these modes will damp out? Answer: Both 4. Determine the dynamic stability factor, Sd. Answer: Sd = 0.74 5. Consider a cargo projectile with identical properties to our projectile in Chapter 8, Problem 35. The designer did not secure the cargo well enough so that the cargo fails to spin up completely during gun launch in a worn tube. When this happens, immediately after muzzle exit, the round spins down (and the cargo spins up a little more) so that the projectile finally reaches a spin rate of 100 Hz. The velocity is unaffected. a. Determine the gyroscopic stability factor for each of the two situations. Answer: Sg = 3.192 and Sg = 0.659. b. Will both projectiles fly properly? Why or why not? Answer: No, the second projectile will tumble. Problem 5 A 155 mm HE projectile is fired from a cannon. The muzzle velocity of the projectile is 800 m/s and the twist of the rifling is 1:20. The projectile and filler properties are given next. Assuming the aerodynamic forces and moments are negligible and that the projectile is dynamically stable: 1. The initial spin rate of the complete projectile. Answer:  rad  pmuzzle = 1621.5   s  2. The spin rate of the projectile in flight assuming the fill does not spin up in the bore and both shell and fill come into dynamic equilibrium. Answer:  rad  ptotal = 1259.2   s  3. Determine the gyroscopic stability factors for (1) and (2). Answer: Sg = 16.54 and Sg = 9.97 4. Is the projectile stable in (1) and (2)? Answer: Yes to both. © 2014 by Taylor & Francis Group, LLC Ballistics: Theory and Design of Guns and Ammunition 326 Projectile and weapon information CMα = 1.07 Clp = − 0.012 I Pshell = 431 [lbm-in.2 ] I Pfill = 124 [lbm-in.2 ]  lbm  ρ = 0.067  3   ft  I P total = 555 [lbm-in.2 ] d = 155 [mm] m = 106 [lbm] m Vmuzzle = 800   s n= I Ttotal = 3335 [lbm-in.2 ] 1  rev  20  cal  Problem 6 For the projectile given in Problem 5, determine the nutation and precession frequencies in Hertz. Answer: dψ 1 = 42.05 [Hz] and dt dψ 2 = 0.67 [Hz] dt Problem 7 Assume the projectile in Problem 1 has slipped its rotating band and the spin at the same instant in time is 130 Hz. Is the projectile stable? Answer: No. Problem 8 What is the minimum spin (Hz) required to stabilize the projectile in Problem 1? Answer: pmin = 140 [Hz]. Problem 9 For the projectile given in Chapter 8, Problem 37, determine the minimum spin rate for stability. (Hint: remember when a projectile is least stable.) Problem 10 A right circular cylinder is to be fired horizontally for an impact test. If the cylinder is made of steel (ρ = 0.283 lbm/in.3) and it is 0.5 in. in diameter and 0.75 in. long, determine a. The spin rate required to stabilize the projectile (if it can be stabilized) b. Comment on the answer above—what is dominant in the problem? c. The precessional and nutational frequencies of the projectile at this spin rate in Hz The projectile properties are provided next: CD = 0.4 CLα = 0.18  lbm  ρair = 0.0751  3   ft  CMα = 8.0 d = 0.5 [in] Clp = −0.02 l = 0.75 [in] CMq + CMαɺ = −29  ft  Vmuzzle = 6000   s CMpα = 0.92 © 2014 by Taylor & Francis Group, LLC Linearized Aeroballistics 327 Problem 11 Modifications are made to a 155 mm M483A1 projectile so that it has the following properties and initial conditions: CD = 0.2 CLα = 1.975 CMα = 4.573 Clp = −0.0285 CMq + CMαɺ = −15.2  lbm  ρ = 0.0751  3   ft  I P = 537.1 [lbm-in.2 ] d = 155 [mm] I T = 5753 [lbm-in.2 ]  ft  Vmuzzle = 2900   s m = 103 [lbm] CMpα = 1.20 At an instant in time after launch when p = 100 [Hz] φ = δ = 2°  ft  V = 1000   s determine a. If the projectile is stable b. The precessional frequency in Hz c. The nutational frequency in Hz Problem 12 What is the minimum spin (Hz) required to stabilize the projectile in Problem 11? Problem 13 For the projectile given in Chapter 8, Problem 25, determine the precessional and nutational frequencies in Hertz. Determine the minimum spin rate for the projectile to be stable. Problem 14 For the projectile given in Chapter 8, Problem 27, determine the precessional and nutational frequencies in Hertz. Determine the minimum spin rate for the projectile to be stable. Problem 15 For the projectile given in Chapter 8, Problem 28, determine the precessional and nutational frequencies in Hertz. Determine the minimum spin rate for the projectile to be stable. Does it differ depending on which side of the aircraft it is fired from? Calculate it for both cases. © 2014 by Taylor & Francis Group, LLC Ballistics: Theory and Design of Guns and Ammunition 328 9.3 Yaw of Repose In Section 9.1, we introduced the yaw of repose for a projectile and defined it in Equation 9.75 using the symbol ξg. The subscript “g” was used to denote that this quantity comes about through the action of gravity on the projectile. In terms of our dimensionless parameters, we can rewrite Equation 9.75 as ξg = PG M + iPT (9.109) A qualitative look at this expression leads to some extremely interesting results. First and foremost is that the spin rate directly affects the yaw. The greater the spin (and therefore the larger the value of P), the greater the yaw of repose is. The second useful item to note is that the more abrupt the trajectory curve is, the greater the yaw of repose is. In fact, if we look at the term G, it is linear in the cosine of angle of attack, ϕ. For details of this form that G takes, the reader is referred to Ref. [5]. Thus, when the projectile approaches maximum ordinate, the yaw of repose should be a maximum given all of the other parameters remain constant. Because of the decay of the other terms, the result is that the yaw of repose is usually a maximum shortly before or after reaching maximum ordinate. The sign of the yaw of repose is important. In our convention, the term P is positive for a right-hand twist. Thus, a positive value of ξg causes the projectile to nose over to the right. Note that there can also be a significant pitch component to this quantity; this is easily seen as the real part of Equation 9.109. If we examine a plot of pitch (α) versus yaw (β) for a British 14-in. projectile in Figure 9.8, we can imagine the yaw of repose as the vector pointing to the right (viewed from the rear) to the center of the precessional path similar to Figure 9.7. We can see that the magnitude as well as the direction of this vector change as the projectile moves downrange. In Figure 9.8, the projectile was analyzed using the PRODAS software and was fired with a muzzle velocity of 2483 ft/s, spin rate of 71 Hz corresponding to a 1:30 twist with an initial pitch α versus β 0.18 0.15 0.12 0.09 α (deg) 0.06 0.03 0.00 –0.03 –0.06 –0.09 –0.12 0.00 0.02 0.04 0.06 0.08 0.10 0.12 0.14 0.16 0.18 0.20 0.22 0.24 0.26 0.28 0.30 β (deg) FIGURE 9.8 Pitching and yawing motion for a British 14-in. Mk.I projectile fired at 2483 ft/s with a 0.1° initial pitch angle. © 2014 by Taylor & Francis Group, LLC Linearized Aeroballistics 329 angle of 0.1°. There was no initial yaw or pitch/yaw rate. This projectile has progressed through only one and one-half yaw cycles (about 1.7 s) when the analysis was stopped to yield a nice clear illustration. 9.4 Roll Resonance Until this point, we have assumed that the projectiles under study have been axially symmetric. This rarely happens in practice because of manufacturing tolerances in a given projectile design. In Chapter 10, we shall discuss the means of handling a slight mass asymmetry. In this section, we shall discuss the implications of a geometric (including slight mass) asymmetry as applied to a fin-stabilized projectile. Fin asymmetries commonly occur when a finned projectile is manufactured or can be the result of damage owing to rough handling. In the field of explosively formed penetrators which are normally drag- or fin-stabilized, inconsistencies can (and usually do) arise due to the explosive formation process. In either case, this effect may be coupled with some mass asymmetry as well. In Equation 9.74, the trim arm was introduced, which would force a statically stable projectile to fly with an angle of attack. It is for this reason that all fin- and drag-stabilized projectiles are designed to roll slightly to increase accuracy. One can see from the way that this equation was written there is no change in the orientation of K30. It was fixed, oriented at the initial angle ψ 30. To begin our assessment of this specific type of asymmetry, we shall start with the governing equation for a spin-stabilized projectile (Equation 9.76) because the roll is going to play a part. We shall alter the RHS to incorporate a forcing term representing the lifting force and moment that is caused by the asymmetry (say, e.g., a bent fin). We shall write this in such a way that the direction of the applied force and moment rotates with the projectile: ξ ′′ + ( H − iP)ξ ′ − ( M + iPT )ξ = −iA3 exp(iψ ) (9.110)  ρ Sd   1  A3 =    2  ( Cm0 + iCn0 ) + (ψ ′ − P) ( CZ0 + iCY0 )  2m   k T  (9.111) where ψ′ = pd , dimensionless turning rate V (9.112) s ∫ ψ = ψ ′ds, dimensionless distance (9.113) 0 This development was put forth in Refs. [1,4,5]. If we look closely at these equations, we see that the forcing function, A3, rotates with the projectile. © 2014 by Taylor & Francis Group, LLC Ballistics: Theory and Design of Guns and Ammunition 330 If we solve Equation 9.110, assuming a solution for the particular part of ξ p = K 3 exp [ i(ψ + ψ 0 )] (9.114) where ψ 0 is some arbitrary angle that contains the plane of the asymmetry, we obtain a general solution for a constant roll rate of ξ = K1 exp[iψ 1 ] + K 2 exp[iψ 2 ] + K 3 exp[i(ψ + ψ 0 )] (9.115) and, after inserting the initial conditions, say, of ψ 0 = 0, we obtain K3 = −iA3 ψ ′2 − Pψ ′ + M − i(ψ ′H − PT ) (9.116) This is the expression for the yaw component caused by a lift force and corresponding moment constrained to rotate at the projectile spin rate. If the spin rate is zero, the orientation of this lift force will be fixed and the projectile will drift more and more in that direction. This is not desirable from an accuracy standpoint so we must have some spin. The denominator in Equation 9.116 is normally dominated by its real part because H and the product PT are small by comparison. However, much like a resonance in a spring-mass system, if the roll frequency ever approaches either one of the precession or nutation frequencies (and remains there for some time), the denominator in Equation 9.114 approaches zero and the yaw becomes very large [1,5]. This usually occurs when the nutational frequency is approached and is called roll resonance or spin-pitch resonance [1]. Since projectiles are usually changing spin rate throughout their flight, this is only a problem if there is a slow change of spin rate when the frequencies are close. Another way of looking at this is to imagine a projectile where this asymmetry is present. Since the asymmetry is at the same frequency as the nutation rate, every time the projectile is at the outer limit of its motion it gets kicked a little further, similar to pushing a child on a swing. This disturbance grows as long as the two motions stay coupled (i.e., at the same frequency); however, if they became out of phase, the problem would correct itself. 1000 600 400 200 0 –200 0 100 200 300 400 500 600 Attitude (deg) Spin rate (rad/s) 800 –400 90 80 70 60 50 40 30 20 10 0 Tumbled 0 (a) Range (ft) (b) 100 200 300 400 500 600 Range (ft) FIGURE 9.9 Explosively formed penetrator experiencing roll resonance. (Courtesy of Eric Volkmann, Alliant Techsystems, Hopkins, MN.) © 2014 by Taylor & Francis Group, LLC Linearized Aeroballistics 331 An example of roll resonance is depicted in Figure 9.9. In this case, an explosively formed penetrator (EFP) was the device under test. Keep in mind that only total angle of attack is measured here so the yawing motion is not constrained to a single plane. We see that as the EFP approached a spin rate of ∼300 rad/s it locked in and flew very far off of the target. Problem 16 Roll resonance of a projectile occurs when the spin rate approaches a “forcing” frequency of the projectile. This occurs more frequently in fin-stabilized projectiles than spin-stabilized projectiles because, in the latter, the spin rate is usually quite high in order to maintain stability, the overturning moment is positive and these forcing functions unless they are intentional—like thrusters are usually due to asymmetries (like bent fins) usually are small. If we examine the projectile of Chapter 8, Problem 37, instead as a fin-stabilized projectile, we can write the equation for the pointing direction as ξ = K1 exp[iψ 1 ] + K 2 exp[iψ 2 ] + K 3 exp[i(ψ + ψ 0 )] where the subscripts 1 and 2 represent the fast and slow modes, respectively. The third term is the forcing function where we can define ψ′ = pd V (1) and, after inserting the initial conditions, say, of ψ 0 = 0, we obtain K3 = −iA3 ψ ′2 − Pψ ′ + M − i(ψ ′H − PT ) (2) Based on what you know about the behavior of imaginary numbers and expressions (1) and (2), determine the spin rate at which catastrophic yaw will occur. Use the aerodynamic properties of the projectile from Chapter 8, Problem 37, but assume that the overturning moment is the negative of what was provided (we are essentially faking a fin-stabilized version). You may assume the velocity stays constant at 750 ft/s. References 1. Murphy, C.H., Free Flight Motion of Symmetric Missiles, Ballistics Research Laboratory Report No. 1216, Aberdeen Proving Ground, MD, 1963. 2. Vaughn, H., A detailed development of the tricyclic theory, Sandia National Laboratories Report No. SC-M-67–2933, Albuquerque, NM, February 1968. 3. McShane, E.J., Kelley, J.L., and Reno, F.V., Exterior Ballistics, University of Denver Press, Denver, CO, 1953. 4. Nicolaides, J.D., On the free flight motion of missiles having slight configurational asymmetries, Ballistics Research Laboratory Report No. 858, Aberdeen Proving Ground, MD, 1953. 5. McCoy, R.L., Modern Exterior Ballistics, Schiffer Military History, Atglen, PA, 1999. © 2014 by Taylor & Francis Group, LLC © 2014 by Taylor & Francis Group, LLC 10 Mass Asymmetries Until this point we have assumed that the projectile has been an axially symmetric body. This allowed us to simplify the equations of motion considerably. Projectiles are rarely axially symmetric. The asymmetry usually comes about through manufacturing tolerances, damage due to rough handling, cargo slippage or, more recently, they are simply designed that way. The purpose of this section is simply to introduce the geometry of mass asymmetries, which will be introduced into the equations of motion for the projectile in later sections. Mass asymmetries come in two categories: static imbalance and dynamic imbalance. In a static imbalance, the center of gravity (CG) of the projectile is not located on the geometric axis of symmetry. The geometric axis of symmetry can be defined by imagining a projectile with the same exterior dimensions as the unbalanced projectile but of uniform density. The symmetry axis would then be centrally located in the body of revolution (i.e., a perfectly axially symmetric body). In a statically imbalanced projectile, this axis would be shifted to pass through the CG but remain parallel to the geometric axis. This is illustrated in Figure 10.1. A dynamically imbalanced projectile also has a CG that is offset from the geometric axis of symmetry. In this case, however, the mass distribution is such that the principal axis of inertia resides as some angle to the geometric axis as well. This is illustrated in Figure 10.2. Whether a projectile is statically or dynamically imbalanced, we shall define the plane in which the CG offset is located relative to some reference plane (we shall arbitrarily use the x–y plane as the reference, which we have defined in earlier sections) using the symbol Φ. This is illustrated in Figure 10.3 as viewed from the rear of the projectile. The effect of these mass asymmetries on projectile flight can dramatically affect accuracy, especially in direct fire systems. Consider a projectile with an imbalance in the gun tube. While in the tube, the projectile is constrained to rotate about the tube geometric axis. If we idealize this situation to say that the tube is perfectly straight, inflexible, and fits the projectile snugly, we can further state that the projectile is constrained to rotate about its own geometric axis. Note that there is a wealth of literature dedicated to the real situation (e.g., [1–10]). 333 © 2014 by Taylor & Francis Group, LLC Ballistics: Theory and Design of Guns and Ammunition 334 CG Principal axis of inertia ε Geometric axis FIGURE 10.1 Statically imbalanced projectile. Principal axis of inertia CG Φ ε Geometric axis FIGURE 10.2 Dynamically imbalanced projectile. Φ CG ε Geometric axis FIGURE 10.3 Center of gravity (CG) offset viewed from rear of projectile. © 2014 by Taylor & Francis Group, LLC Mass Asymmetries 335 References 1. Berger, M.P., Position and form of bands for projectiles, Memoires Militaires et Scientifiques, Publies par le Department de la Marine, Translated by LT C.C. Morrison, Notes on the Construction of Ordnance, Washington, DC, June 10, 1884. 2. Kent, R.H. and Hitchcock, H.P., Comparison of Predicted and Observed Yaw in Front of the Muzzle of a 12′ Gun, Report No. 990 AD-116-140, USA BRL, Aberdeen Proving Ground, MD, July 1956. 3. Kent, R.H. and McShane, E.J., An Elementary Treatment of the Motion of a Spinning Projectile about Its Center of Gravity, Report No. 459 AD-491-943, USA BRL, Aberdeen Proving Ground, MD, April 1944. 4. Heppner, L.D., Setback and Spin for Artillery, Mortar, Recoilless Rifle and Tank Ammunition, Report No. DPS-2611, USA BRL, Aberdeen Proving Ground, MD, January 1968. 5. Gay, H.P. and Elder, A.S., The Lateral Motion of a Tank Gun and Its Effect on the Accuracy of Fire, Report No. 1070 AD-217-657, USA BRL, Aberdeen Proving Ground, MD, March 1959. 6. Kirkendall, R.D., The Yawing Motions of Projectiles in the Bore, Technical Note No. 1739 AD-878327-L, USA BRL, Aberdeen Proving Ground, MD, September 1970. 7. Zaroodny, S.J., On Jump due to Muzzle Disturbances, Report No. 703 AD-805-876, USA BRL, Aberdeen Proving Ground, MD, June 1949. 8. Gay, H.P., On the Motion of a Projectile as It Leaves the Muzzle, Technical Note No. 1425 AD-801974, USA BRL, Aberdeen Proving Ground, MD, August 1961. 9. Sterne, T.E., On Jump due to Bore Clearance, Report No. 491 AD-491-938, USA BRL, Aberdeen Proving Ground, MD, September 1944. 10. Line, L.E., The Erosion of Guns at the Muzzle, NRDC Report No. A-357, OSRD Report No. 6322, National Defense Research Committee, Office of Scientific Research and Development, Washington, DC, November 1945. © 2014 by Taylor & Francis Group, LLC © 2014 by Taylor & Francis Group, LLC 11 Lateral Throwoff Earlier in the text, we stated that projectiles rarely leave the tube with their velocity vectors aligned with the geometric axis of the gun tube. This chapter and Chapter 12 describe this behavior. The result of this behavior is weapon inaccuracy and it must be well understood by the practicing ballistician because, although it is not practical to completely eliminate the behavior, we would like to reduce it to acceptable levels. The first component of this behavior is known as lateral throwoff. It is a dynamic response of the projectile to either a static or a dynamic imbalance and will now be described in detail. If we imagine a projectile with a mass asymmetry as depicted in Figure 10.3, we can imagine the spinning motion as viewed from the rear. If we ignore the axial velocity by simply spinning the projectile at a high rate, say, between two flexible supports on a test stand, we would see a wobble develop as a result of the centrifugal action on the center of mass. All the time the projectile is being spun up in the gun, the tube walls and stiffness of the supporting members prevent this wobble (to the extent the clearances allow) from developing. At the instant, the projectile is free from the constraints of the tube we expect it to become affected by this centrifugal loading. This is lateral throwoff because the effect is to fling the projectile in a direction off the tube centerline. We can use the analogy of a vacuum trajectory to examine the lateral throwoff effect generated by either a static or a dynamic imbalance. Consider the projectile asymmetry from Figure 10.3. If we examine the projectile over a short period of flight, ignoring gravity as well as assuming no drag because of the vacuum assumption, we would see the dynamic forces acting on the projectile as depicted in Figure 11.1. In this figure, the only force acting is the centrifugal force due to spin. This dynamic action will result in the force vector changing direction, though since there is no angular acceleration or deceleration it maintains a constant magnitude. It is worth noting that we have resorted to our complex plane in this example as it is convenient to use in our development. At the instant, in time depicted here, we can break the force into a component in the y-direction and one in the iz-direction. We are not necessarily concerned with the force acting on the CG per se. We want to see where the projectile moves because of this force. To accomplish this, we need to use Newton’s second law. We know that Fr = mar (11.1) This is the centripetal force. The centrifugal force would be equal but opposite in sign. From dynamics [1], we recall that ar = −rp 2 (11.2) In the case we are considering here, we see that r =ε and Φ = pt (11.3) 337 © 2014 by Taylor & Francis Group, LLC Ballistics: Theory and Design of Guns and Ammunition 338 y Fr [i sin(Φ)] Φ Fr [cos(Φ)] Fr iz Center of gravity ε Geometric axis FIGURE 11.1 Dynamic force acting on a statically or dynamically imbalanced projectile. With this, we can write the magnitude of the force as Fr = mε p 2 (11.4) and the centripetal acceleration in the complex plane as a=− Fr [cos( pt) + i sin( pt)] = −ε p 2 [cos( pt) + i sin( pt)] m (11.5) The complex velocity can therefore be expressed as t ∫ V = −ε p 2 [cos( pt) + i sin( pt)]dt (11.6) 0 Evaluating the integral and assuming that as the projectile leaves the muzzle we have an initial orientation of the mass asymmetry of Φ = Φ 0 yields V = −ε p[sin( pt + Φ0 ) − i cos( pt + Φ0 )] = ε p[− sin( pt + Φ0 ) + i cos( pt + Φ0 )] (11.7) To see how much lateral movement has developed, we can integrate again t ∫ r = ε p [− sin( pt + Φ0 ) + i cos( pt + Φ0 )] dt 0 © 2014 by Taylor & Francis Group, LLC (11.8) Lateral Throwoff 339 The evaluation of which yields r = ε [cos( pt + Φ 0) + i sin( pt + Φ 0)] (11.9) As an example, if we were only concerned with motion in the crossrange direction, we could state z = Im{ε [cos( pt + Φ 0) + i sin( pt + Φ 0)]} = ε sin( pt + Φ 0) (11.10) To apply numbers to this example, let us consider a projectile that weighs 100 lbm and is spinning at a rate of 270 Hz. We shall assume the projectile has a CG offset of 0.25 in. If this were the case, the velocity in the z-direction as well as the motion for the first 4 s of flight can be seen in Figures 11.2 and 11.3. Here we have assumed that the CG offset has emerged from the weapon at the 12 o’clock position. 0.2 0.196 Im (Vr(t)) 0.15 0.1 0.05 0 0 1 0 0 2 t 3 4 4 FIGURE 11.2 Velocity in the z-direction of a 100-lbm projectile spinning at 270 Hz with a 0.25 in. CG offset. 0.4 z (ft) 0.3 0.2 0.1 0 0 1 2 3 4 t (s) FIGURE 11.3 Displacement in the z-direction of a 100 lbm projectile spinning at 270 Hz with a 0.25 in. CG offset. © 2014 by Taylor & Francis Group, LLC Ballistics: Theory and Design of Guns and Ammunition 340 The most interesting observation between the figures is that for this arbitrary emergence of the CG offset, we see that the projectile would like to move laterally to the right for a right-hand spin. This is commonly known as drift. Just to put things into perspective, the muzzle velocity consistent with the 270 Hz spin rate is about 2,750 ft/s so the projectile would only have gone about 0.4 ft to the right after it traversed 11,000 ft downrange. We must always bear in mind that this example was an idealized situation. In the case of a real projectile, there are other forces acting which complicate the motion; however, it is instructive to look at simplifications such as this to see the phenomenon at work. We will now move on to examine the dynamic behavior in terms of the equations of motion of a projectile from statically imbalanced and dynamically imbalanced projectiles. We shall see how this affects lateral throwoff. 11.1 Static Imbalance In Figure 10.1, we saw the effect on the principal axis of a static imbalance. Although this rarely happens in production (imbalances are usually of the dynamic type) it can happen and presents an interesting case. We shall follow the analysis procedure documented by McCoy [2] in the development, correcting terms to fit our coordinate system. If we examine the velocity of the center of mass of the projectile as it leaves the gun tube, we see a scene as depicted in Figure 11.4. If the projectile is constrained as it continues down the gun tube, the motion of the CG would resemble a spiral or helix similar to a thread on a bolt, except that the pitch of the helix would continue to increase as the axial velocity increases. We could express this mathematically using a cylindrical set of coordinates with x indicating the axial distance, r indicating the radius of the CG from the Φ Center of gravity VΦ ε Geometric axis FIGURE 11.4 Velocity of a statically imbalanced projectile’s CG. © 2014 by Taylor & Francis Group, LLC Lateral Throwoff 341 centerline, and Φ indicating the angular position from the vertical plane. If we assume the tube is straight, then the axial component of the velocity vector will be constrained along the tube and our unit vector, l, will describe the direction adequately. We shall use the unit vectors er and eΦ to represent the radial and angular positions, respectively. If we use our instantaneous spin rate, p, as defined in Equation 11.3, we can write the tangential component of velocity as VΦ = rpeΦ = ε peΦ (11.11) Vx = V l (11.12) The axial velocity is simply Then the velocity vector could be written in cylindrical coordinates as V = V l + ε pe Φ (11.13) Or, if we like to remain in Cartesian coordinates, we can combine Equation 11.13 with Equation 11.7 to yield V = V l + ε p(− sin( pt + Φ 0)n + i cos( pt + Φ 0)m) (11.14) These Cartesian coordinates are useful when we want to write the velocity vector at the muzzle of the weapon. The lateral throwoff caused by a static imbalance can be described as the tangent of the angle of the projectile CG as it exits. For small angles (usually the case), this is approximately the angle itself in radians. With this, we can define the lateral throwoff at the muzzle owing to a static imbalance as TL = ε p0 [− sin(Φ 0) + i cos(Φ 0)] V0 (11.15) where we have used t = 0 at the muzzle and specified the spin rate and muzzle velocity. We must keep in mind that this is an angular measure for small angles or, more precisely, a tangent of an angle. We can use the relationship i exp(iθ ) = − sin θ + i cos θ (11.16) to write TL = i ε p0 exp(iΦ0 ) V0 (11.17) If the projectile has a rotating band that forces it to spin based on the rifling twist, this expression can be written in terms of the projectile diameter and twist rate as well. This is extremely straightforward and left as an exercise for the reader. © 2014 by Taylor & Francis Group, LLC Ballistics: Theory and Design of Guns and Ammunition 342 11.2 Dynamic Imbalance The diagram of Figure 10.2 represents the most common case of a projectile asymmetry, a dynamic imbalance. In this case, a lateral throwoff effect as described in Section 11.1 will result as the projectile leaves the muzzle of the gun and there will also be significant flight dynamic effects as the projectile moves downrange. Usually, this mass asymmetry is small and can be treated as a small amount of mass removed from or added to a projectile at a point defined by a radial set of coordinates from the CG. We shall use the former approach following the development of Ref. [2]. This is depicted in Figure 11.5. Figure 11.6 depicts how this removed mass is oriented relative to the CG offset in the radial direction. The development put forward in Ref. [2] assumes that the removed mass is much smaller than the overall mass of the projectile. As established earlier, we will use two orthogonal coordinate systems. The first is our i-j-k triad which is oriented along the projectile axis as depicted in Figure 9.1. This coordinate system does not roll with the projectile. We shall also make use of a second non-rolling coordinate system using the l-n-m system depicted in the same figure. In this case, the coordinate system is oriented along the velocity vector. Principal axis of inertia Center of gravity rE ε lE rE Mass removed, mE FIGURE 11.5 Dynamically imbalanced projectile with mass removed. Φ Center of gravity VΦ Geometric axis ε mE rE FIGURE 11.6 Velocity of a dynamically imbalanced projectile’s CG. © 2014 by Taylor & Francis Group, LLC Lateral Throwoff 343 The coordinate systems are related to one another, assuming small yaw angles, through the relationships i = γ l +αn + βm (11.18) j = −α l + n (11.19) k = −β l + m (11.20) γ = cos α cos β ≈ 1 (11.21) where α is the pitch angle β is the yaw angle γ is defined as The angular momentum of the projectile is the vector sum of all of the angular momenta and is closely approximated by  di  H = I P pi + I T  i ×  − mE (rE × v E )  dt  (11.22) Here H is the total angular momentum v E is the velocity of the removed mass This velocity can be broken into two components, one owing to the rotation about the spin axis and the other owing to the yawing motion of the projectile as follows:  di   v E = p(i × rE )+  i ×  × rE  d t     (11.23) Then inserting this relationship into Equation 11.22 and combining terms gives us, after utilization of the vector triple product  di  H = (I p − mErE2 )pi + IT  i ×  − mE  dt   di   di    ×  − p(rE ⋅ i)rE −  rE ×  (rE × i) + (rE ⋅ i)  rE ×   dt   dt     (11.24) If we examine Figure 11.5, we see that (rE ⋅ i) = lE © 2014 by Taylor & Francis Group, LLC (11.25) Ballistics: Theory and Design of Guns and Ammunition 344 And we note that for a spin-stabilized projectile, the yaw rate, di/dt, is much smaller than the spin rate, p, we can eliminate terms in Equation 11.24 to yield  di  H ≈ I P pi + I T  i ×  + mE plErE  dt  (11.26) We can express the mass asymmetry vector r E in terms of the projectile geometric axes as rE = lEi + rE cosΦ j + rE sin Φ k (11.27) This can be expressed in our coordinate system attached to the velocity vector through the relationships in Equations 11.18 through 11.20 as rE = (lEγ − rEα cosΦ − rE β sin Φ )l + (lEα + rE cos Φ )n + (lE β + rE sin Φ )m (11.28) We can simplify this expression somewhat if we use the fact that both α and β are much smaller than γ. In this case, the expression would simplify to rE = (lEγ )l + (lEα + rE cos Φ )n + (lE β + rE sin Φ )m (11.29) We can take the derivative of Equation 11.29 using the fact that the coordinate system is effectively not rotating to write drE = (lEγɺ )l + (lEαɺ − rE p sin Φ )n + (lE βɺ + rE p cos Φ )m dt (11.30) Here we have used the fact that dΦ =p dt (11.31) As in our previous analyses, we shall consider a short period of flight. By doing this, we can neglect all forces and moments except the pitching (overturning) moment. This allows us to equate the rate of change of angular momentum to the applied pitching moment  d 2i  dH di dr * α V 2 ( l × i) ≈ I P p + I T  i × 2  + mE plE E = mCM dt dt dt  dt  (11.32) This expression can be written as a set of three equations in terms of each component as follows:  d2β d 2α  I P pγɺ + I T  α 2 − β 2  + mE plE2γɺ = 0 dt   dt © 2014 by Taylor & Francis Group, LLC (11.33) Lateral Throwoff 345  d 2γ d2β  * αV 2 β I P pαɺ + I T  β 2 − γ + mE plE (lEαɺ − rE p sin Φ ) = −mCM 2  d d t t   (11.34)  d 2α d 2γ  * αV 2α I P pβɺ + I T  γ − α 2  + mE plE (lE βɺ + rE p cosΦ ) = mCM 2 d d t t   (11.35) The details of this are provided in Ref. [2]. If we change the temporal derivatives into spatial derivatives along a dimensionless downrange distance, s, and define the following: I P= P  IT M=   pd   V    (11.36) md 2 *α CM IT (11.37) I E = mErElE (11.38) We can rewrite Equations 11.33 through 11.35 as  m l2  P  1 + E E  γ ′ + αβ ′′ − βα ′′ = 0 IT    m l2  P  1 + E E  α ′ + βγ ′′ − γβ ′′ + Mβ IT   I I −  E 2T  IP (11.39)  2  P sin Φ = 0  (11.40)  m l2  I I  P  1 + E E  β ′ + γα ′′ − αγ ′′ − Mα +  E 2T  P 2 cosΦ = 0 I T   IP   (11.41) Here the primed quantities are differentiated with respect to s. With small yaw as well as classical size assumptions (see Refs. [2,3]), we can neglect several of these terms because they are either products of small numbers or summed with a much larger number. This results in Equation 11.39 vanishing altogether and the other two transforming into I I Pα ′ − β″ + Mβ =  E 2T  IP  2  P sinΦ  (11.42) I I Pβ ′ + α ′′ − Mα = −  E 2T  IP  2  P cosΦ  (11.43) If we now multiply Equation 11.42 by −i and add it to Equation 11.43, we obtain I I (α ′′ + iβ ′′) + P( β ′ − iα ′) − M(α + iβ ) = −  E 2T  IP © 2014 by Taylor & Francis Group, LLC  2  P (cosΦ + i sinΦ )  (11.44) Ballistics: Theory and Design of Guns and Ammunition 346 If we invoke our definition of complex yaw angle, we can write this as I I ξ ′′ − iPξ − Mξ =  E 2T  IP  2  P exp(i Φ )  (11.45) The solution to this differential equation was discussed in Section 9.1. The difference here is that the forcing term on the RHS is somewhat different. If we use a solution written as ξ = K1 exp(iψ 1 ) + K 2 exp(iψ 2 ) + K 4 exp(iΦ ) (11.46) where K1 and K2 are the solutions to the homogeneous part of the equation and K4 is our new term which depends on the spin rate and the mass asymmetry we can solve for the magnitude of the trim arm caused by the asymmetry. If we solve Equation 11.46 for the particular solution, we find that this new trim arm caused by the mass asymmetry is given by K4 = IE  I p2 M  IT − IP +    ITP  (11.47) The third term in the denominator is usually very small so this term has been approximated (see Refs [2,4].) K4 ≈ IE IT − IP (11.48) This trim arm due to a mass asymmetry is usually small. Throughout this development, IP and IT have been used as the moments of inertia even though, in the purest sense, the mass asymmetry removes the axially symmetric properties of the projectile. For most cases, it is sufficient to use these quantities based on an axially symmetric projectile. References 1. Greenwood, D.T., Principles of Dynamics, 2nd edn., Prentice Hall, Englewood Cliffs, NJ, 1988. 2. McCoy, R.L., Modern Exterior Ballistics, Schiffer Military History, Atglen, PA, 1999. 3. Murphy, C.H., Free Flight Motion of Symmetric Missiles, Ballistics Research Laboratory Report No. 1216, Aberdeen Proving Ground, MD, 1963. 4. Murphy, C.H., Yaw Induction by Means of Asymmetric Mass Distributions, Ballistics Research Laboratory Memorandum Report No. 2669, Aberdeen Proving Ground, MD, 1976. © 2014 by Taylor & Francis Group, LLC 12 Swerve Motion Following our procedure of slowly introducing complexity into the description of projectile behavior we shall now develop equations to characterize the remainder of what is known in general as swerve motion. We saw in Chapter 11 that a mass asymmetry can cause projectile motion transverse to the original line of fire even in a vacuum. We stated in that section that a dynamic projectile imbalance was more common than a static imbalance but either can actually occur. Chapter 6 explained many aspects of projectile behavior that arise due to the presence of the air stream. All of the coefficients were functions of the angle of the attack observed by the projectile relative to that air stream. If we examine how a statically or dynamically imbalanced projectile would behave as viewed from above the trajectory curve based on its spin, we would see motion as depicted in Figures 12.1 and 12.2. We must keep in mind that the motion in these figures is greatly exaggerated for ease of viewing. We can imagine, by looking at these figures that the aerodynamic forces would be considerable because even in the case of the statically imbalanced projectile, motion laterally across the trajectory will manifest itself in an angle of attack and therefore affect the flight characteristics. In this section, we shall describe and evaluate the aerodynamic forces that arise from this behavior and include them in our equations of motion for projectile flight. We shall also include the effect of configurational asymmetries such as bent fins or damaged form because these will result in similar behavior even without the mass asymmetry present. In fact, to a varying degree, every projectile has a combination of both form and mass asymmetries present. 12.1 Aerodynamic Jump McCoy [1] has shown that the equation of motion for the point mass solution plus swerving motion is given by ( gd d2 y d2z i + = CLα *ξ − 2 exp 2CD* s 2 2 V0 ds ds ) (12.1) with a solution of  dy y + iz = ( y0 + iz0 ) +   ds  ( )   s 2 gd  exp 2CD* s − 2CD* s − 1  dz  * +i  s + CLα I L − 2  ds 0  2V02  0 2 CD* s   ( ) (12.2) 347 © 2014 by Taylor & Francis Group, LLC Ballistics: Theory and Design of Guns and Ammunition 348 Trajectory FIGURE 12.1 Motion of a statically imbalanced projectile. Trajectory FIGURE 12.2 Motion of a dynamically imbalanced projectile. where s s1 IL = ∫ ∫ ξ ds ds 1 2 (12.3) 0 0 Here we have used s1 and s2 as dummy variables representing integrations with respect to s. Equation 12.2 describes the position of the projectile in a direction perpendicular to the trajectory based on flat fire point mass assumptions. © 2014 by Taylor & Francis Group, LLC Swerve Motion 349 Equation 9.72 was developed as a solution for ξ. Reference [1] has shown that the solution to the double integral of Equation 12.3 can be obtained by substitution of Equation 9.72 into Equation 12.3 resulting in   λ − iψ 1′   λ − iψ 1′  K exp(iψ 20 )  s I L = −  12 K exp(iψ 10 ) +  22 2  20 2  10   λ1 + ψ 1′   λ2 + ψ 2′  + (R11 − iR12 )K10 exp(iψ 10 ) {exp[(λ1 − iψ 1′ )s] − 1} + (R21 − iR22 )K 20 exp(iψ 20 ) {exp[(λ2 − iψ 2′ )s] − 1} + i ( )   PG0 2  exp 2CD* s − 2CD* s − 1  (12.4) s 2   M * C s 2 D   ( ) Here we have used G0 = R11 = R12 = R21 = R22 = gd V02 (12.5) λ12 − ψ 1′2 (λ 2 1 + ψ 1′2 ) (12.6) 2 2λ1ψ 1′ (λ 2 1 + ψ 1′2 ) 2 (12.7) 2 (12.8) λ22 − ψ 2′2 (λ 2 2 (λ + ψ 2′2 2λ2ψ 2′ 2 2 + ψ 2′2 ) ) (12.9) 2 If we make the assumption that λ12, 2 ≪ ψ 1′2, 2 , the aforementioned parameters become R11 ≈ − 1 1 , R12 ≈ 0, R21 ≈ − 2 , and R22 ≈ 0 2 ψ 1′ ψ 2′ (12.10) Inserting these assumptions into Equation 12.4 yields the following result:   1 1 1 I L = i  K10 exp(iψ 10 ) + K 20 exp(iψ 20 )  s − 2 K10 exp(iψ 10 ) {exp[(λ1 + iψ 1′ )s] − 1} ′ ′ ψ2  ψ 1′ ψ1 − ( )   1 PG0 2  exp 2CD* s − 2CD* s − 1  ′ i s K i + i s − 1 + exp( ) exp[( ) ] ψ λ ψ { } 20 2 2 20 2   M ψ 2′2 2CD* s   © 2014 by Taylor & Francis Group, LLC ( ) (12.11) Ballistics: Theory and Design of Guns and Ammunition 350 This result is important because it depicts the three components of swerve motion. The first term on the RHS is called the aerodynamic jump, JA, and it is what we will examine for the remainder of this section. The second two terms are the epicyclic swerve, SE, and will be discussed in Section 12.2. The third term is called drift, DR, and will be discussed in Section 12.3. To keep things simple, we will restate the aerodynamic jump as   1 1 J A = iCL*α  K10 exp(iψ 10 ) + K 20 exp(iψ 20 )  ψ 2′   ψ 1′ (12.12) We should note a few things about Equation 12.12. First, we must keep in mind that in Equation 12.11 this aerodynamic jump term is multiplied by a downrange distance, s, implying that it is actually an angular measure (for small angles). A second observation is that the aerodynamic jump is completely dependent upon the initial conditions of the projectile and how these couple in with the fast and slow arm turning rates. If we insert our approximated initial fast and slow arm amplitudes from Equations 9.78 and 9.79 into Equation 12.12, we obtain  −iξ 0′ − ψ 2′ξ 0′ iξ ′ + ψ 1′ξ 0′  + 0 J A = iCL*α   ψ 1′ (ψ 1′ − ψ 2′ ) ψ 2′ (ψ 1′ − ψ 2′ )  (12.13) This can be rewritten [1] as C J A = kT2  Lα  CMα   (iPξ 0 − ξ 0′ )  (12.14) Downrange This result shows that by knowing the projectile mass properties and launch conditions, we can determine to what angle a projectile will “jump.” We can envision this jump effect as shown in Figure 12.3. While we have said a great deal mathematically about aerodynamic jump, we have not really described the physics behind it. Because of the presence of aerodynamic lift on the projectile, there is a strong influence of angle of attack on the resultant motion. We saw Line of departure of projectile tan–1( JA) Mean motion due to aerodynamic jump Crossrange FIGURE 12.3 Graphical representation of aerodynamic jump. © 2014 by Taylor & Francis Group, LLC Swerve Motion 351 earlier that a projectile, through purely dynamic means, can yaw because of either spin or some geometric asymmetry. When this happens, the aerodynamic forces change, either improving or worsening the situation. This interaction of the aerodynamic forces with the projectile manifests itself in the jump angle as depicted in Figure 12.3. 12.2 Epicyclic Swerve The second two terms in Equation 12.11 describe the epicyclic swerve of a projectile. We can define this parameter specifically [1] as   1 1 SE = −CL* α  2 K 20 exp(iψ 20 ) {exp[(λ1 + iψ 1′ )s] − 1} + 2 K 20 exp(iψ 20 ) {exp[(λ2 + iψ 2′ )s] − 1}  ′ ′ ψ ψ 2   1 (12.15) McCoy [1] has shown that this equation can be put into a more useful form through use of the relation C CL*α = kT2  Lα  CMα  ψ 1′ψ 2′  (12.16) If we insert Equation 12.16 into Equation 12.15, we obtain  C  ψ ′ ψ′  SE = −kT2  L α   2 K10 exp(iψ10 ) {exp[(λ1 + iψ 1′ )s] − 1} + 1 K 20 exp(iψ 20 ) {exp[(λ2 + iψ 2′ )s] − 1} ′ ′ ψ ψ C M 2 1 α    (12.17) Following McCoy, we shall examine two special cases of this equation. The first is where we have a projectile that is non-spinning (statically stable) and the second is a spin-stabilized projectile with a good gyroscopic stability (measured at muzzle exit) of at least 1.5. For the non-spinning projectile, the following conditions apply: M < 0, P = 0, and ψ 2′ = − ψ 1′ (12.18) If we insert these conditions into Equation 12.17, we get C  SEnon-spin = kT2  Lα  ( K10 exp(iψ 10 ) {exp[(λ1 + iψ 1′ )s] − 1} + K 20 exp(iψ 20 ) {exp[(λ2 + iψ 2′ )s] − 1})  CMα  (12.19) Now we can invoke the fact that the spin is equal to zero and insert Equation 9.59, in which we shall neglect the trim and yaw of repose, into Equation 12.19 to yield C  S Enon-spin = kT2  Lα  (ξ − ξ 0 )  CMα  © 2014 by Taylor & Francis Group, LLC (12.20) Ballistics: Theory and Design of Guns and Ammunition 352 This relationship will produce a motion in exactly the same manner as the aerodynamic jump developed in Section 12.1. It essentially couples the yawing motion of the projectile to the swerving motion. Both will thus damp together and the more yaw, the greater the epicyclic swerve. If we examine the spin-stabilized projectile, we can write M > 0 and ψ 1′2 ≫ ψ 2′2 (12.21) With the aforementioned mathematical statements, McCoy [1] has stated that an excellent approximation of Equation 12.15 for a spinning projectile is  C ψ ′  SEspin = −kT2  Lα   1  ( K 20 exp(iψ20 ) {exp[(λ2 + iψ 2′ )s] − 1})  CMα   ψ 2′  (12.22) An interesting comparison may be drawn between the epicyclic swerving behavior of a spinning projectile and a non-spinning projectile. If we compare Equations 12.22 and 12.20, we see that in the latter the yawing motion and swerve are locked together and operate in the same fixed plane. This is because the lift generated by the motion never rotates. In a spin-stabilized projectile, the lift vector is always rotating, thus the center of mass of the projectile will move in a helical manner around the flight path. Furthermore, the motion will be locked to the rate of turning of the slow arm and will damp or increase as the slow arm does. 12.3 Drift The last term in Equation 12.11 describes the drift of a projectile. We can define this parameter specifically [1] as DR = i ( )   PG0 2  exp 2CD* s − 2CD* s − 1  s 2   M 2CD* s   ( ) (12.23) If we expand the term in brackets in a power series, we can rewrite this equation as  PG0 DR = i   M 2  1  2 2 *  s 1 + CD s + CD* s + ⋯ 3   3  ( ) ( ) (12.24) Examination of the drift equation in this form has some advantages. First, we can see that if a projectile has no spin, P = 0 and there is no drift. If we look at a fin- or drag-stabilized projectile where M < 0, we see that the projectile will drift in the direction opposite to © 2014 by Taylor & Francis Group, LLC Swerve Motion 353 the spin. That is, a left-hand spin will produce a right-hand drift and vice versa. In a statically unstable (spin-stabilized) projectile where M > 0, we see that the projectile will drift in the same direction as the spin. It must be noted that this drift is very small compared to the other swerve components as well as Coriolis drift. In fact, to even measure it, some researchers [1] have fired two projectiles simultaneously out of side-by-side gun barrels with both left- and right-hand twist to remove Coriolis and wind drift components which would affect each equally. The interested reader should consult Ref. [1] for further information on this topic. Reference 1. McCoy, R.L., Modern Exterior Ballistics, Schiffer Military History, Atglen, PA, 1999. © 2014 by Taylor & Francis Group, LLC © 2014 by Taylor & Francis Group, LLC 13 Nonlinear Aeroballistics Until this point, we have concerned ourselves with linear behavior of the aerodynamic coefficients only. This is very convenient for direct fire projectiles and projectiles that fly with very little yaw. It had the benefit of allowing us to make a black or white decision with regard to projectile stability as well—the projectile was either stable or not. In real systems, several of the coefficients are only linear over a small range of angles of attack. This can be either helpful or hurtful to a particular design. Limit-cycle motion is motion that develops over time in a projectile, whereby the projectiles angle of attack grows until a certain (sometimes rather large) angle is achieved. As the angle of attack increases (or some other parameter such as the air density changes), the coefficients change so that the projectile will actually become stable at some large angle of attack. At first, this may seem like it is a desirable quality in a projectile; however, range is sacrificed due to the larger drag generally associated with this large yaw. Some systems have been fielded unwittingly in this condition and it was only after a large number of firings in the field that this was determined to be an issue. This nonlinear behavior arises out of the interaction between the air and the surfaces of the projectile. It is a rather complicated mechanism that can arise (many times in a discontinuous manner) from boundary layer separation, fin masking, vortex shedding, etc. All of which are fluid dynamic phenomenon. This is and continues to be a challenging area of aeroballistic research, where experimental, theoretical, and computational techniques are pushed to the limit of their usefulness. The next two sections will look at this behavior to some degree of detail; however, because of space constraints, the reader is encouraged to consult the literature for more detailed mathematical and theoretical treatment. 13.1 Nonlinear Forces and Moments In general, we can divide nonlinear forces and moments into two categories: geometric and aerodynamic nonlinearities. The geometric nonlinearities arise from the cosine terms in the equations of motion that were eliminated when we assumed a small yaw angle. This small angle assumption is generally valid for most projectiles in flight. If a projectile is flying with large yaw, the cosine terms must be retained and the resulting equations are more difficult to solve. Since this behavior is usually designed out of projectiles, we shall focus on the second type of nonlinearity, the aerodynamic nonlinearity. The aerodynamic nonlinearity can exist even at angles of attack that are consistent with the small yaw assumption. They arise due to the fluid–mechanic interaction of the air with the solid projectile body. This interaction can consist of phenomena such as vortex shedding, separation, shock interactions, etc. 355 © 2014 by Taylor & Francis Group, LLC Ballistics: Theory and Design of Guns and Ammunition 356 The most dominant force acting on the body is the drag force. In all of our previous discussions, we have stated that the forces that arise due to other sources are small, and that is still true for the case of nonlinearities; however, the moments caused by these other forces cannot be neglected. We can define a nonlinear drag coefficient as CD = CD0 + CDδ 2δ 2 + ⋯ (13.1) In this equation, the first term on the RHS is the zero-yaw-drag coefficient and the second term is the yaw-drag coefficient. More coefficients can be added but typically the expression is truncated at the yaw-drag term. There are essentially two common ways of determining the yaw-drag coefficient: experimentally or computationally. Experimental evaluation is more common although recent advances in computational fluid dynamics (CFD) [1] have shown that it is possible to extract coefficients directly from analyses. In either case, the overall drag coefficient at multiple angles of attack is determined from either a direct force measurement (in the case of a wind tunnel or CFD model) or the velocity decay (in a free flight test), and the results are plotted as CD versus angle of attack. The slope of the resulting line (hopefully, it is a line) is then the yaw-drag coefficient and the y-intercept is the zero-yaw-drag coefficient. In the case of a free flight firing where the projectile is dragging down continuously, Murphy [2] and McCoy [3] suggest an averaging scheme that has been successfully demonstrated based on a great deal of experience. The aforementioned technique is known as a quasi-linear approach because it defines a linear function that is a solution to a nonlinear equation. The same approach is used to determine the nonlinear moments, which are generally assumed to have the same form as Equation 13.1. In general, both the zero-yaw-drag coefficient and the yaw-drag coefficient are positive values. In the case of the pitching or overturning moment of a spin-stabilized projectile, the zero-yaw overturning moment coefficient is positive while the cubic overturning moment coefficient is negative [3]. This condition can have some interesting effects on stability as summarized by McCoy [3]. The overall equation of motion that includes all of the nonlinear terms that is equivalent to our linear equation (Equation 9.76) with the gravitational term neglected is ξ ′′ + ( H 0 + H 2δ 2 − iP)ξ ′ − [ M0 + M2δ 2 + iP(T0 + T2δ 2 )]ξ = 0 (13.2) We can define our coefficients as follows: H0 = ρ Sd  1  CLα 0 − CD0 − 2 (CMq + CMαɺ )0   2m  kT  (13.3) H2 = ρ Sd  1  CLα 2 − CD2 − 2 (CMq + CMαɺ )2  2m  kT  (13.4) M0 = ρ Sd 1 CMα0 2m kT2 (13.5) M2 = ρ Sd 1 CMα2 2m kT2 (13.6) © 2014 by Taylor & Francis Group, LLC Nonlinear Aeroballistics 357 T0 = ρ Sd  1  CLα0 + 2 CMpα0   2m  kP  (13.7) T2 = ρ Sd  1  CLα 2 + 2 CMpα2   2m  kP  (13.8) I P= P  IT   pd   V    (13.9) The solution to Equation 13.2 is ξ = K10 exp[λ1s] exp[i(ψ 10 + ψ 1′s)] + K 20 exp[λ2s] exp[i(ψ 20 + ψ 2′ s)] (13.10) where again, we are reminded that the gravitational term has been neglected. McCoy [3] has written expressions for the coefficients in terms of the damping exponents and turning rates as follows: δ e12 = K12 + 2K 22 (13.11) 2 δ e2 = K 22 + 2K12 (13.12)  K 2 + K 22  ψ 1′ + ψ 2′ = P + M2  1 ≈P  ψ 1′ −ψ 2′  (13.13) ψ 1′δ e22 −ψ 2′δ e21 ψ 1′ −ψ 2′ (13.14) ψ 1′ψ 2′ = M0 + M2δ e2 (13.15) δ e2 = λ1 = ) ( ) ( ) − H 0ψ 1′ + P T0 − T2δ e21 − H 2 ψ 1′ K12 + K 22 + ψ 1′K 22    ψ 1′ −ψ 2′ ( ) − H 0ψ 2′ + P T0 − T2δ e22 − H 2 ψ 2′ K12 + K 22 + ψ 1′K12    λ2 = ψ 1′ −ψ 2′ ( (13.16) (13.17) In terms of some of these parameters, McCoy [3] has derived a form for the nonlinear lift coefficient as  ψ ′2δ 2 K eiψ 1 + ψ 1′2δ e22K 2eiψ 2  CLα = CLα 0 + CLα 2  2 e12 1 iψ 1  iψ 2 2  ψ 2′ K1e + ψ 1′ K 2e  © 2014 by Taylor & Francis Group, LLC (13.18) Ballistics: Theory and Design of Guns and Ammunition 358 As is readily apparent, these expressions are significantly more complex than their linear cousins. Because of this, they are generally solved using numerical schemes. The interested reader is referred to Refs. [4–12] for a more detailed treatment as well as examples of this behavior. Problem 1 If the projectile in Problem 1 of Chapter 9 happens to be flying at a limit-cycle yaw of 4° with a spin rate of 130 Hz and velocity 1764 ft/s. What would the nonlinear pitching moment have to be for the projectile to be marginally stable? Hints: 1. Assume all of the other coefficients are linear. 2. Recall the definition of the nonlinear pitching moment (you have the linear part in Problem 1 of Chapter 9). Answer: CMα 2 = −119.079 13.2 Bilinear and Trilinear Moments We have discussed nonlinear forces and moments and their implications in the previous section. At this point, we shall turn our attention to nonlinear moments in which the cubic behavior itself can be described by a bilinear or trilinear curve. This is evident when the cubic coefficient is plotted versus yaw angle. A bilinear coefficient would have two different linear slopes, while a trilinear moment would have three. This is quite useful since many experimental data can be fitted using these curves. In particular, we shall examine the Magnus moment and its implications because this is the dominant moment in spinstabilized projectile flight behavior [3]. If we are examining projectile flight data, it is often tempting to fit a higher order polynomial curve to deal with the nonlinearity. This is usually not advisable since the abrupt changes in behavior at certain angles of attack are caused by fluid–solid interactions such as boundary layer separation, vortex shedding, etc. To describe the behavior of projectiles with nonlinear Magnus moment coefficients, we shall use two examples: one with a linear cubic Magnus moment and one with a bilinear Magnus moment. We are interested in two things: first, the effect of initial conditions on projectile stability and second, limit-cycle motion. In the excellent treatment by McCoy [3], for illustrative purposes, the author suggested assuming a linear pitch damping moment with a cubic Magnus moment coefficient. This will force H2 to be zero and allow Equations 13.16 and 13.17 to be written as λ1 = λ2 = © 2014 by Taylor & Francis Group, LLC ( − H 0ψ 1′ + P T0 + T2δ e21 ) ψ 1′ −ψ 2′ ( − H 0ψ 2′ + P T0 + T2δ e22 ψ 1′ −ψ 2′ ) (13.19) (13.20) Nonlinear Aeroballistics 359 We can put these equations into the form λ1 = λ10 + λ12 δ e21 (13.21) λ2 = λ20 + λ22 δ e22 (13.22) where, we can define λ10 = − H 0ψ 1′ + PT0 ψ 1′ −ψ 2′ (13.23) λ20 = − H 0ψ 2′ + PT0 ψ 1′ −ψ 2′ (13.24) PT2 ψ 1′ −ψ 2′ (13.25) λ12 = −λ22 = With these expressions, we can draw plots of damping coefficients versus yaw angle in a manner similar to the coefficients. At this juncture, we need to recall that these damping exponents will decrease the yaw of their particular mode if they are negative, and increase the yaw if they are positive. Thus, negative values are stabilizing and positive values are destabilizing. As a simple example, let us look at a projectile that has a linear cubic Magnus moment. In analyzing this projectile, we create two plots of damping coefficient versus yaw. These are depicted as in Figures 13.1 and 13.2. In Figure 13.1, we can see the fast mode damping coefficient is negative for all yaw angles of interest (if the projectile is flying at an angle above 11°, we probably have a problem). 0.0006 0.0004 λ1 0.0002 0.01 0.02 0.03 0.04 0.05 δ 2e1 –0.0002 –0.0004 λ10 –0.0006 FIGURE 13.1 Plot of fast mode damping coefficient versus yaw for linear cubic fast mode. © 2014 by Taylor & Francis Group, LLC Ballistics: Theory and Design of Guns and Ammunition 360 0.0006 λ2 0.0004 0.0002 0.01 0.02 0.03 0.04 0.05 δ 2e2 –0.0002 –0.0004 λ20 –0.0006 FIGURE 13.2 Plot of slow mode damping coefficient versus yaw for linear cubic slow mode. Thus, the fast mode will always damp for this projectile. Examination of Figure 13.1 reveals that as long as the projectiles yaw angle is below 5.73°, the slow arm will damp to zero (recall that the yaw angle is equal to sin(δ)); above this angle, it will grow without bound. Although this angle is fairly large for a projectile, there have been instances documented where a slowly launched missile was stable when fired from one side of a fast warship, but unstable when launched from the other [2,8]. The instability was caused by the vector addition of the ships own speed with the launch velocity. Figures 13.3 and 13.4 show the fast and slow damping exponents for a projectile with bilinear cubic Magnus moment behavior. This is an interesting example because it 0.0006 λ1 0.0004 0.0002 0.01 0.02 0.03 0.04 0.05 δ 2e1 –0.0002 –0.0004 λ10 –0.0006 FIGURE 13.3 Plot of fast mode damping coefficient versus yaw for bilinear cubic fast mode. © 2014 by Taylor & Francis Group, LLC Nonlinear Aeroballistics 361 0.0006 0.0004 λ20 λ2 0.0002 0.01 0.02 0.03 0.04 0.05 δ 2e2 –0.0002 –0.0004 –0.0006 FIGURE 13.4 Plot of slow mode damping coefficient versus yaw for bilinear cubic slow mode. illustrates how a projectile can enter into limit-cycle motion. Limit-cycle motion is motion in which the projectile cones in a predictable manner about the velocity vector. If we examine Figure 13.3, we see that, similar to our earlier case, the fast arm damping coefficient is everywhere negative. Because of this, the fast mode will always damp to zero. The interesting part of the story is shown in Figure 13.4. Here, we see that for small angles, the projectiles slow arm will continue to grow because the damping exponent is positive. Once the amplitude of the motion grows beyond 5.74°, the sign of the coefficient changes driving the motion back to zero. However, the motion cannot be driven all the way back to zero because as soon as the angle decreases below 5.74°, the now positive damping coefficient will again cause it to increase. The end result will be a projectile that cones about the velocity vector at a 5.74° angle. These examples assumed that the velocity of the projectile has had no affect on the exponents. We must always keep in mind that there are many interrelated phenomena that affect these coefficients—the real world is a complicated place. This discussion should provide you with a feel for the physics of the projectile behavior. References 1. DeSpirito, J. and Heavey, K.R., CFD computation of magnus moment and roll damping moment of a spinning projectile, AIAA Paper No. 2004-4713, American Institute of Aeronautics and Astronautics, New York, August 2004. 2. Murphy, C.H., Free flight motion of symmetric missiles, Ballistics Research Laboratory Report No. 1216, Aberdeen Proving Ground, MD, 1963. 3. McCoy, R.L., Modern Exterior Ballistics, Schiffer Military History, Atglen, PA, 1999. 4. Murphy, C.H., Data reduction for the free flight spark ranges, Report No. 900, USA BRL, Aberdeen Proving Ground, MD, February 1954. 5. Murphy, C.H., Limit cycles for non-spinning statically stable symmetric missiles, Report No. 1071, USA BRL, Aberdeen Proving Ground, MD, March 1959. © 2014 by Taylor & Francis Group, LLC 362 Ballistics: Theory and Design of Guns and Ammunition 6. Murphy, C.H., The measurement of non-linear forces and moments by means of free flight tests, Report No. 974, USA BRL, Aberdeen Proving Ground, MD, February 1956. 7. Murphy, C.H., Advances in the dynamic analysis of range data, Memorandum Report No. 1270, USA BRL, Aberdeen Proving Ground, MD, May 1960. 8. Platus, D.H., Dynamic instability of finned missiles caused by unequal effectiveness of windward and leeward fins, AIAA Paper No. 70–206, American Institute of Aeronautics and Astronautics, New York, January 1970. 9. Tobak, M., Schiff, L.B., and Peterson, V.L., Aerodynamics of bodies of revolution in coning motion, AIAA Journal, 7(1), January 1969, 95–99. 10. Seginer, A. and Rosenwasser, I., Magnus effect on spinning transonic finned missiles, AIAA Paper No. 83–2146, American Institute of Aeronautics and Astronautics, New York, August 1983. 11. Platou, A.S., Magnus characteristics of finned and nonfinned projectiles, AIAA Journal, 3(1), American Institute of Aeronautics and Astronautics, New York, January 1965. 12. Cohen, C.J., Clare, T.A., and Stevens, F.L., Analysis of the non-linear rolling motion of finned missiles, AIAA Paper No. 72-980, American Institute of Aeronautics and Astronautics, New York, September 1972. © 2014 by Taylor & Francis Group, LLC Part III Terminal Ballistics © 2014 by Taylor & Francis Group, LLC © 2014 by Taylor & Francis Group, LLC 14 Introductory Concepts Terminal ballistics is the regime that the projectile enters at the conclusion of its flight. It has been delivered into its flight by the interior ballistician, pursued and guided through its flight by the exterior ballistician, and now at its target becomes the responsibility of the terminal ballistician. The basic objective of firing the projectile is to defeat some type of target and we will study the widely varying phenomena of terminal effects that are the tools of the terminal ballistician. These end effects are dependent on the design and mission of the projectile. The most common of the missions are as follows: fragmentation of the projectile body by its cargo of high explosives; penetration or perforation of the target by the application of kinetic or chemical energy; blast at the target area delivered by the chemical energy of the explosive cargo; and the dispersal of the cargo for lethal or other missions, e.g., smoke, illumination, propaganda dispersal, etc. Since most terminal ballistic phenomena involve the generation and effects of stress waves in solids, we will spend some time examining the details of this field. We must gain some knowledge of terminal ballistic terminology to be able to study the theories of kinetic energy penetration of solid targets; detonation, deflagration, and burning of energetic materials; the fundamentals of shaped charges; fragmentation theories; blast effects; and lethality with the study of wound ballistics. We shall begin by introducing some concepts that we shall use throughout our study of this field. In examination of penetration theories, we need to consider the following items: What constitutes defeat of the target? What is the source of the data for which we have to create a theory? Does the theory track with respect to momentum balance or energy balance? How many empirically derived constants are there in the model (this tells us how universal the theory will be)? What simplifications and assumptions were made? Penetration is defined as an event during which a projectile creates a discontinuity in the original surface of the target. Perforation requires that, after projectile or its remnants are removed, light may be seen through the target. Since penetration is a somewhat stochastic event, we need to define some statistical parameters. V10 is the velocity at which a given projectile will defeat a given target 10% of the time. V50 is the velocity at which a given projectile will defeat a given target 50% of the time, and V90 is the velocity at which a given projectile will defeat a given target 90% of the time. These quantities are depicted in Figure 14.1. The 50% penetration velocity is commonly used as both experimental measurement as well as a production check. The following procedure illustrates its usage in an experiment. The reader should refer to Figure 14.2 to illustrate the meaning. First, we should estimate V50 through a calculation. Once this is accomplished, we fire a projectile with a Vs as close to V50 as we can achieve. Let us say, the velocity of this experimental firing is a bit over our estimate (at 1 in Figure 14.2). Assuming shot 1 only partially penetrated, we increase the velocity considerably, and let us say that we achieve complete penetration at 2 in the figure. We now assume that V50 is midway between 1 and 2. We now would attempt to fire at the velocity halfway between 1 and 2 (at 3) and, say, we get complete penetration. We would 365 © 2014 by Taylor & Francis Group, LLC Ballistics: Theory and Design of Guns and Ammunition 366 Probability of success 0.90 0.50 0.10 V V10 V50 V90 FIGURE 14.1 Statistical velocities defined. Vs Complete penetration Complete penetration Partial penetration 2 6 Complete penetration 3 1 5 Partial penetration Initial estimate Partial penetration 4 FIGURE 14.2 Illustration of the V50 experimental procedure. next lower the velocity to get a partial penetration, say at 4, then we would increase it to get a complete penetration (but let us say, we get only a partial penetration at 5). We would then have to increase the next shot velocity to 6. We would continue the aforementioned procedure, commonly known as an up and down test, until we obtained three complete penetrations and three partial penetrations with the difference between the highest and the lowest velocities in the set less than 200 ft/s. At that point, we would calculate the experimental V50 from V50 ∑ = 6 i=1 6 Vi (14.1) The limit velocity, Vi (sometimes called the ballistic limit when referring to the armor), is the velocity below which a given projectile will not defeat a given target. The technique for determining it was invented by the U.S. Army Ballistics Research Laboratory (BRL), Aberdeen, Maryland. The object is to fire a few projectiles that achieve complete penetration, measuring the residual velocity through the use of flash x-rays, and then generate a © 2014 by Taylor & Francis Group, LLC Introductory Concepts 367 Vr Vl Vs FIGURE 14.3 Limit velocity illustrated. curve as shown in Figure 14.3. We then plot the residual velocity after penetration versus the striking velocity. Usually, there will be a lower limit that develops below which the armor is not penetrated or the projectile gets stuck in the armor. From experimental evidence, we know that the following factors affect the limit velocity: material hardness, yaw at impact, projectile density, projectile nose shape, and length to diameter ratio of the projectile. For the material hardness, in general, the harder the target, the higher V50 becomes; while the harder the penetrator, the lower V50 becomes and there is more residual penetrator. With respect to yaw at impact, the more yaw, the greater chance for breakup or ricochet and the higher V50 becomes. With projectile density, we find that the more dense the projectile is, the lower V50 becomes. A blunter nose translates, in general, to a higher V50. If the target is overmatched significantly, however, the nose shape has negligible effect. The length to diameter ratio can go either way and a great deal depends on the obliquity of impact. We will now introduce some concepts, which we shall use in our examination of penetration events. © 2014 by Taylor & Francis Group, LLC © 2014 by Taylor & Francis Group, LLC 15 Penetration Theories Now that we have a firm grounding in some penetration concepts such as limit velocity, we can proceed to discuss various penetration theories. We shall discuss, in some detail, penetration mechanisms in a variety of materials, all of which, to different degrees, serve to protect some vital target. Because these materials behave very differently from one another, they must be treated separately. It is this large difference in behavior, as well as mechanical properties, which makes the selection of a material for ballistic protection an important one. We shall move successively through metals, concrete, soil, ceramic, and composite armors so that the reader gets a feel for how they behave. In all instances, the day-to-day analysis techniques of these materials are progressing, especially in the areas of numerical methods. 15.1 Penetration and Perforation of Metals Metals are by and large the most common target of medium to large caliber projectiles. Although small caliber ammunition is generally used against soft targets, there are times when even they are called upon to penetrate metal objects. This section will discuss several models of penetration into two of the most common metals: steel and aluminum. While these formulas are not exactly perfect for other metals, usually a metal will behave like one or the other. Projectiles may impact metallic targets under a wide range of velocities. The nature of the target material is such that different velocities must be handled using somewhat different techniques. At very low velocities (<250 m/s), the penetration is usually coupled to the overall structural dynamics of the target. Responses are on the order of 1 ms. As the impact velocity increases (500–2000 m/s), the local behavior of the target (and sometimes penetrator) material dominates the problem. This local zone is approximately 2–3 projectile diameters from the center of impact. With further increases in velocity (2000–3000 m/s), the high pressures involved allow the materials to be modeled as fluids in the early stages of impact. At impact speeds greater than 12,000 m/s, energy exchange occurs at such a high rate that some of the colliding material will vaporize. This energy exchange must be accounted for. We will not treat this last case as it is beyond the normal scope of military applications. A typical sequence of events that occur during a projectile impact is developed here [1]. Given that a projectile strikes a target, compressive waves propagate into both the projectile and the target. Relief waves propagate inward from the lateral free surfaces of the penetrator, cross at the centerline, and generate a high tensile stress. If the impact were normal, we would have a two-dimensional stress state. If the impact were oblique, bending stresses will be generated in the penetrator. When the compressive wave reached the 369 © 2014 by Taylor & Francis Group, LLC Ballistics: Theory and Design of Guns and Ammunition 370 free surface of the target, it would rebound as a tensile wave. The target may fracture at this point as will be seen in Section 16.3. The projectile may change direction if it perforates (usually toward the normal of the target surface). Because of the differences in target behavior based on the proximity of the distal surface, we must categorize targets into four broad groups. A semi-infinite target is one where there is no influence of distal boundary on penetration. A thick target is one in which the boundary influences penetration after the projectile is some distance into the target. An intermediate thickness target is a target where the boundaries exert influence throughout the impact. Finally, a thin target is one in which stress or deformation gradients are negligible throughout the thickness. There are several methods by which a target will fail when subjected to an impact. The major variables are the target and penetrator material properties, the impact velocity, the projectile shape (especially the ogive), the geometry of the target supporting structure, and the dimensions of the projectile and target. The failure modes of the target are depicted in Figure 15.1. They will now be described. Spalling is very common and is the result of wave reflection from the rear face of the plate. It is common for materials stronger in compression than in tension. Scabbing is similar to spalling, but the fracture results predominantly from large plate deformation, which begins a crack at a local inhomogeneity. These failure mechanisms will be expounded upon in Section 16.3. Brittle fracture occurs usually in weak and lower density targets. Radial cracking is common in ceramic type materials where the tensile strength is lower than the compressive strength, but it does occur in some steel armor. Plugging occurs in materials that are fairly ductile and usually when the projectile impact velocity is very close to the ballistic limit. Petaling occurs when the radial and circumferential stresses are high and the projectile impact velocity is close to the ballistic limit. Perforation Non-perforation Plugging Piercing (ductile) Petaling Scabbing Scab Spalling (brittle) Spall ring FIGURE 15.1 Target failure modes. © 2014 by Taylor & Francis Group, LLC Star crack Penetration Theories 371 Because of the very high loading rates and correspondingly high temperatures, we need to describe some phenomena that occur during penetration events. Terms such as these occur throughout the literature, so it is good to understand what they mean. The concept of adiabatic shearing is encountered in impacts where a plug has been formed. On initial impact, a local ring of intense shear is generated. Since this occurs very quickly (∼μs), the target does not have sufficient time to build up any motion. Locally intense heat is generated. Because of the time scale and a large deformation rate, the heat cannot be conducted away. Since the material properties are weaker at this high temperature, the material tends to yield readily and flow plastically. The process then feeds on itself. Finally, a plug is formed and breaks free. If the minimum perforation velocity is exceeded by more than about 5%–10%, the plug will usually break up. Blunt noses on projectiles tend to increase the propensity to fail a target by adiabatic shear. Hydrodynamic erosion is an important concept in terminal ballistics. Metal cutting tools such as water jets or soft metal penetrators and shaped charge jets can defeat a target by hydrodynamic erosion. During hydrodynamic erosion, the penetrator material forces the target material aside in a manner similar to a punch being pushed into the target material except that the hole will be larger. This phenomenon usually occurs at impact velocities over 1000 m/s. Deposition of the penetrator material on the walls of the hole is an indication that this failure mechanism played a part in the penetration. The hydrodynamic transition velocity is the velocity below which the projectile and target act as essentially elastic bodies and above which both target and projectile can be treated as fluids. This concept is illustrated by the penetration sequence of Brooks [1]. For all penetration velocities, the target material is accelerated radially away from the axis of penetration. At low velocities, elastic strain keeps the target material in contact with the penetrator. At high velocities, the material is thrown away from the projectile, so that the hole becomes bigger than the projectile diameter. The radial acceleration of the material is greatest at the tip of the projectile. At the hydrodynamic transition velocity, the tip of the penetrator deforms laterally. The projectile tip becomes spherically blunted and forms a stable shape that penetrates the target for the remainder of the event. The transition velocity varies inversely with the tip radius. Hydrodynamic transition velocity is possibly related to the rate of rod erosion and plastic wave propagation. Shear banding is a form of adiabatic shearing in which layers of material in a like state of shear tend to form. There are discontinuities in stress and strain instead of a gradual increase in shear strain near the disturbed region. Uranium and tungsten tend to display this phenomenon. Normal material models used in finite element codes do not show this effect. A model that includes thermal softening is required. The analytical models in use today to solve these types of problems can be organized into three broad categories: empirical or quasi-analytical, approximate analytical, and numerical. In empirical or quasi-analytical models, algebraic equations are developed from large amounts of experimental data. These models are generally curve fits (results based). They usually do not incorporate physics and tend to be configuration dependent. An approximate analytical model attempts to examine the physics of a particular aspect of the penetration process or failure mechanism such as petaling, plugging, etc. The mathematics becomes tractable because we must make simplifying assumptions. They are usually limited to particular situations. Numerical models usually attempt to solve the full equations of continuum mechanics using finite difference or finite element techniques. This is the most general method. The problem with numerical models is that good material models are required and this can be expensive. © 2014 by Taylor & Francis Group, LLC 372 Ballistics: Theory and Design of Guns and Ammunition Most analytical models can only consider one damage mechanism (like plugging or fracture) or conservation law before they become mathematically intractable. Some allow as many as two mechanisms. The approach is to make simplifying assumptions. Typical assumptions are to assume localized influence where the projectile is only influenced by a small region of the target, to ignore rigid body motions, and to ignore thermal, friction, shock heating, and any material behavioral changes due to these mechanisms, that the target is initially stress free, etc. One important thing to recognize is that a complicated model does not necessarily yield a more accurate answer. Perforation of finite thickness plates in which plugging is the predominant penetration mode is divided into three stages. In the first stage, locally, the material ahead of the projectile is compressed and the mass is added to the projectile (i.e., the projectile decelerates somewhat and the added mass accelerates). In the second stage, more material is accelerated but shearing is occurring on the surface area of the plug. In the third stage, the plug has completely sheared out and both the plug as well as the projectile move with the same velocity. If this model is used for an oblique impact, one must use the line-of-sight thickness. At velocities from 1200 to 5000 m/s, the model used usually involves hydrodynamic erosion of the projectile tip as the first stage as well. This can be followed by both plugging and further tip erosion. In the third stage, we usually consider the projectile to be completely eroded and the plug is ejected from the armor [1]. Some models account for the flexibility of the target. This is usually required as the impact velocities approach the limit velocity. In this case, a significant amount of energy is consumed in both elastically and plastically bending the target plate. We shall examine the underlying assumptions in a few penetration theories before moving on to detailed examination of the theories themselves. Theories that are derived from a momentum balance are typically used for thin plates. These theories can be used with minor modifications when the target petals. They usually require that the projectile remains intact. Theories that are derived from an energy balance are typically used for thick and moderately thick plates. With moderately thick targets, plugging can occur. Thick plates are usually defeated by a piercing phenomenon that also has distinct phases. The first phase is a radial displacement of the target material. Sometimes, there is plugging at this stage. This stage is followed by plastic flow and yielding of the target. The target material may well be able to be treated like a fluid during this phase. Many empirically based predictive relationships are based on energy approaches. A particularly popular model takes the form of E = kd mt n (15.1) m+n ≈ 3 (15.2) where we have In these equations E is the perforation energy d is the projectile diameter t is the plate thickness k is an empirically derived constant (see Figure 15.2) © 2014 by Taylor & Francis Group, LLC Penetration Theories 373 θ t d FIGURE 15.2 Projectile impact problem illustrated. If we let m = 1.5 and n = 1.4, we get the famous DeMarre formula for normal impact. If we would like to include an angle of obliquity in the aforementioned formula, it is common practice to use E = kd mt nsec pθ (15.3) Here p is an experimental parameter based on the projectile–armor combination and θ is the angle of obliquity measured from the normal to the plate. Sometimes, the armor fabrication process will affect the penetration. In this case, there is a function called the figure of merit (FOM) where the perforation velocity of the armor is compared to that of mild steel. FOM = Vl Vlmild steel (15.4) Note that in Equation 15.4, the velocity used does not necessarily have to be the limit velocity. Another useful relationship commonly employed by the projectile designer is Eperf = 1 2 mVperf 2 (15.5) Inserting Equation 15.3 into Equation 15.5 yields 2 Vperf = 2k d mt n sec p θ m (15.6) Now taking the square root and assimilating terms, we get Vperf = k © 2014 by Taylor & Francis Group, LLC d mt n sec j(θ ) θ m (15.7) 374 Ballistics: Theory and Design of Guns and Ammunition In 1886, DeMarre developed a famous formula for the penetration of a plate given a normal impact: mV 2 t1.4 α = d3 d1.5 (15.8) where m is the penetrator mass V is the impact velocity d is the diameter of the projectile t is the plate thickness with α being an empirically derived constant As a word of caution, many of these formulas are dangerous because of the units in the empirically derived constant, it is commonplace to see CGS units in these formulas as well. Over time, many have modified the DeMarre formula and used it in this form mV 2 t =α  d3 d β (15.9) Here β is an empirically derived constant as well. In the aforementioned form, the DeMarre formula is used when considering a normal impact. Some researchers have extended its use to include an oblique impact and it would then take the following form: mV 2  tg(θ )  =α  d3  d  β (15.10) where g(θ) is a function of the angle of obliquity and is most often taken as sec θ. We sometimes define the specific limit energy (SLE) as mVl2 ≡ SLE d3 (15.11) Hans Bethe, a physicist at Cornell University in 1941, determined that for piercing type problems (i.e., thick plate perforation where a hole is laterally or radially widened by the penetrator), the constant, β, should be equal to 1, thus yielding mVl2 ∼ td 2 (15.12) Around the same time (1942), Zener and Holloman from Watertown Arsenal came up with a formula for use when plugging or petaling is the predominant penetration mode. They stated that in this case, β should equal to 2, thus yielding mVl2 ∼ t 2d (15.13) In 1943, Curtis and Taub attempted to modify the DeMarre formula to account for a mode change during the penetration event. In a thick plate, the mode changes at some point © 2014 by Taylor & Francis Group, LLC Penetration Theories 375 t t΄ d FIGURE 15.3 Section of a target plate that defines t and t′. from a piercing to a plugging at the rear surface. This results in a decrease in energy consumed per unit path length, so the DeMarre formula had to be further modified to mVl2 t  =α  +γ  d3 d   (15.14) Here α and γ are constants and γ < 0. If we define t′ as depicted in Figure 15.3, then γ is a quadratic function of t′. Also t′ ∼ d and is the distance after the mode changes. S. Jacobson, working at the Picatinny Arsenal in New Jersey further refined the concept that there is a different energy relationship for each of the two modes. For plugging, this is Eplug = force ⋅ distance ≈ π dt Ys ⋅ t (15.15) where Ys is the shear yield strength of the material. For the piercing mode, we have E piercing = Yflow ⋅ V ≈ π d2 tYflow 4 (15.16) where V is the volume of the plug Yflow is the flow or plastic yield stress of the target material We can rewrite Equations 15.15 and 15.16 as 2 t Eplug = k plug d 3   Ys d (15.17) t Epiercing = k piercing d 3   Yflow d (15.18) If we graph both expressions, we obtain a plot as illustrated in Figure 15.4. To obtain t/dcrit, we solve Equations 15.17 and 15.18 where E plug = E piercing (15.19) Ys ≈ 0.6Yflow (15.20) using the relations that k piercing = © 2014 by Taylor & Francis Group, LLC π 4 and k plug = π (15.21) Ballistics: Theory and Design of Guns and Ammunition 376 E Plugging mode Piercing mode t/d t/dcrit FIGURE 15.4 Energy in penetration modes based on the model of Jacobson. Then combining Equations 15.17 and 15.18, we get 2 π 3 t  t t d   Yflow = π d 3   0.6Yflow →   = 0.42 4 d  d crit d (15.22) This value of t/d is the point where the mode of penetration changes from plugging to piercing. Thus, against targets whose thickness is such that an attack by a penetrator whose t/d ratio is greater than 0.42, we can expect that the penetration mode will be piercing, otherwise plugging is to be expected. Lambert and Zukas proposed a model in 1982 while working at the BRL to cover more general cases of penetration. If we examine Equation 15.14, we can see that as the plate thickness goes to zero, the residual velocity should approach the striking velocity and the limit velocity should approach zero. Expressed mathematically, we require that lim Vl → 0 t →0 (15.23) However, if we look at Equation 15.14, we note that if Vl = 0 and t = 0 it requires the product γα to equal zero, which is not physically possible. Therefore, the Lambert model replaces γ by [exp (−t/d) − 1] as t mVl2  t  = α  + exp  −  − 1 3 d  d  d (15.24) Vl = 0 at t = 0 (15.25) Vl = ∞ at t = ∞ (15.26) This forces and © 2014 by Taylor & Francis Group, LLC Penetration Theories 377 Since the penetrator volume is proportional to d2l and since there should be a dependence on this volume in the specific limit energy, we want to keep the dimension of diameter cubed in Equation 15.24, thus we shall write 2  l  d 3 → d 3 − cl c = d l   d c −1 l = d3   d c (15.27) where c is a constant. We can then incorporate this into Equation 15.24 as c mVl2  l   t  t  =   α  + exp  −  − 1 3 d  d   d  d (15.28) Next we will include obliquity effects by adding in the angle of obliquity, θ through replacement of t by t seck θ. In this case, if k = 1, we have the true path length through the armor plate (line-of-sight thickness). We shall define z= t sec k θ d (15.29) We can now rewrite Equation 15.29 as c mVl2  t    l  t = α    sec k θ + exp  − sec k θ  − 1 3 d  d    d  d (15.30) If we solve Equation 15.30 for the limit velocity, we obtain c 3  t  d  l  t Vl = α    sec k θ + exp  − sec k θ  − 1  d  m  d  d (15.31) The Lambert model was used to examine the firing of 200 long-rods into rolled homogeneous armor (RHA). The test conditions were as follows: 0.5 ≤ m [g] ≤ 3630 0.6 ≤ t [cm] ≤ 15 0.2 ≤ d [cm] ≤ 0.5 0° ≤ θ ≤ 60° 4≤ l ≤ 30 d  g  7.8 ≤ ρ  3  ≤ 19.0  cm  A least-squares fit of the results yielded the following: α = (4000)2, c = 0.3, and k = 0.75. If we insert these into Equation 15.31, we get l Vl =   d 0.15 ( 4000) © 2014 by Taylor & Francis Group, LLC d3  t  t   m sec 0.75 θ + exp  − sec 0.75 θ  − 1   m  d d   s (15.32) Ballistics: Theory and Design of Guns and Ammunition 378 Please note the CGS units. The authors suggest that the model is applicable where t/d > 1.5. Also we must note that nose geometry has a significant influence for t/d < 1.0. RHA or good quality steel is the target (the specific properties are unimportant). One measure of lethal effects once a projectile has perforated the target material is the residual velocity. Vr is the symbol for the residual velocity of the penetrator. That is the velocity that the penetrator moves with once it perforates the target. Mathematically, it is defined in the Lambert model as 0, 0 ≤ Vs ≤ Vl     Vr =   1 p  a (Vsp − Vlp ) , Vs > Vl  (15.33) If we assume that Vs is large so that the absorption of momentum by the target is negligible, then the momentum balance can be written in terms of identifiable penetrator mass and velocity (mr and Vr), and the large quantity of unidentifiable target and penetrator ejecta with each particle mi having a particular velocity, Vi. Thus, the momentum balance is n mrVr + ∑ m V → m V as V → ∞ i i s s (15.34) s i =1 Even though Equation 15.34 is mathematically satisfying, in practice, it is usually difficult to measure the mass and velocity of all of the fragments, so most of the miVi will remain unknown. We shall now consider a general case of impact as illustrated in Figure 15.5. Here we shall let m′ be the mass of the ejecta. We can then write π 3 d z 4 (15.35) t sec 0.75 θ d (15.36) m′ = ρ where z= t m θ Vs m΄ FIGURE 15.5 General case of projectile impact. © 2014 by Taylor & Francis Group, LLC d Penetration Theories 379 therefore m′ = ρ π 3 t π  d  sec 0.75 θ  = ρ d 2t sec 0.75 θ 4 d 4  (15.37) If we now assume that n ∑ m V = hm′V i i (15.38) r i =1 This is equivalent to stating that m′ is the mass of material pushed ahead of the penetrator, m′ is ejected with speed Vr (plugging theory), and the total momentum of the ejecta jumble is proportional to m′Vr. We can also write, in the limiting case, that the residual momentum approaches the initial momentum or, mathematically Mr →1 M (15.39) If we substitute Equation 15.38 into Equation 15.34, we get mVr + hm′Vr → msVs as Vs → ∞ (15.40) which can be rearranged to yield Vr  ms  →  as Vs → ∞ Vs  mr + hm′  (15.41) We know that if penetration occurred, Equation 15.33 applies, so we have ( Vr = a Vsp − Vlp ) 1 p (15.42) We can divide Equation 15.42 by Vs to get 1   V p  p Vr = a 1 −  l   Vs   Vs   (15.43) which means that as Vs approaches infinity, the second term in the parentheses approaches zero or Vr → a as Vs → ∞ Vs This is illustrated in Figure 15.6. © 2014 by Taylor & Francis Group, LLC (15.44) Ballistics: Theory and Design of Guns and Ammunition 380 ms Vr mr + hm΄ Vr Vs Vs Vl FIGURE 15.6 Asymptote on limit velocity. If we look at Equations 15.44 and 15.41, we see that  ms  a=   mr + hm′  (15.45) Furthermore, we can assume in the plugging mode that the penetrators’ mass does not change significantly during penetration, so we get ms = mr = m. We can then write Equation 15.45 as  m  a=   m + hm′  (15.46) There is empirical evidence that suggests that h ≈ 1/3, so we can write  m  a =   1  m + 3 m′  (15.47) If we assume that the penetrator remains intact throughout the perforation event, we can write (15.48) KEimpact = KElimit + KEresidual This can also be expressed as ( Vr2 ∼ Vs2 − Vl2 → Vr ∼ Vs2 − Vl2 ) 1 2 (15.49) which, if written as ( Vr = a Vs2 − Vl2 ) 1 2 (15.50) would say that p = 2. If we looked at momentum, we would get Vs ~ Vl + Vr © 2014 by Taylor & Francis Group, LLC (15.51) Penetration Theories 381 which could be written as Vs = a(Vl + Vr ) (15.52) Equation 15.52 implies that for a momentum balance, p = 1. Thus, it is clear that the value for p should fall between 1 and 2. Lambert accounted for this by choosing t sec 0.75 θ d (15.53) t π → ∞ and/or θ → d 2 (15.54) p → 2 as t → 0 (15.55) p = 2+z = 2+ where both p and z grow monotonically as also Lambert also found that a better empirical fit was obtained if he let p = 2+ z t = 2+ sec 0.75 θ 3 3d (15.56) A numerical model for penetration was proposed by A. Tate to determine penetration of metals [2]. The base equation for this model is 1 1 ρ p ( Vi − u)2 + Yp = ρ t u2 + Rt 2 2 (15.57) where ρp is the density of the projectile material ρt is the density of the target material Vi is the impact velocity u is the instantaneous projectile velocity Yp and Rt are the ballistic resistances of the projectile and target, respectively, defined as Yp = 1.7σ p (15.58) 2 E   Rt = σ t  + ln  0.57 t   3 σ t    (15.59) where σ P is the yield strength of the projectile material σt is the yield strength of the target material Et is the modulus of elasticity of the target material © 2014 by Taylor & Francis Group, LLC Ballistics: Theory and Design of Guns and Ammunition 382 The way the Tate model is used is to integrate Equation 15.57 numerically until the velocity goes to zero or perforation occurs. When the projectile stops, a second integration determined the depth of penetration. When perforation occurs, the value of u is the residual velocity. Tate states that the accuracy of this method is within 20%. One of the model’s downsides is that it does not handle oblique impacts but it can at least be altered by the line-of-sight thickness. If a penetrator hits a target at a great enough angle, it may ricochet. The ricochet process can be described as follows. During impact, both the projectile and the target are compressed elastically. When this energy is released, it will change the projectile’s motion. Deformations because of resisting force of the target will change the direction of the penetrator. Rotating moments are generated by internal forces acting within the projectile. In general, thin plates do not allow ricochet except at extreme angles of attack. Tate has produced a ricochet formula for the critical ricochet angle (oblique impacts at angles greater than this will ricochet): 1 2 ρ pV 2  L2 + d 2    ρ p  2  tan β >   1+   3 YP  Ld    ρ t     3 (15.60) Here Yp is a characteristic strength usually taken as the Hugoniot elastic limit (described in Section 16.3), the subscripts “p” and “t” are projectile and target, respectively, and L, d, and V are the length, diameter, and velocity of the penetrator, respectively. As vehicles become lighter weight, aluminum is being used more and more as armor. It is therefore necessary to determine the penetration capabilities of projectiles into aluminum. Aluminum behaves a little differently than steel during penetration by ogival projectiles in its tendency to be pierced rather than to develop plugs. A penetrator is usually of significantly greater density than the target in most cases. One significant difference is the evidence of a layer of aluminum with an altered microstructure on the penetrated surface. This indicates a melt layer that is believed to assist in penetration. A simple model of projectile penetration into aluminum was put forward by Forrestal et al. in 1992 [3]. A distinct advantage of this model is its simplicity. A possible disadvantage is that the empirical nature is not universal. Even though the study was performed specifically with 7075-T651 targets, it yields a fairly good representation of aluminum penetration. The model assumes normal impact of the projectile and that the projectile is rigid. This may, at first, seem to be a restrictive assumption, but the method provides reasonable estimates for slightly yawed projectiles if the angle is below about 5° and possibly further. We first define the caliber-radius-head as s d (15.61) d 4ψ − 1 2 (15.62) ψ = where d is the diameter of the projectile ψ is the caliber-radius-head s is the ogive radius We can also define a nose length as l= This geometry is illustrated in Figure 15.7. © 2014 by Taylor & Francis Group, LLC Penetration Theories 383 L l d s θ FIGURE 15.7 Ogival penetrator for the model of Forrestal et al. We shall say that the resistance force of the aluminum target on the penetrator in this case will have two components: one normal to the surface (normal stresses) and one tangential to the surface (shear stresses and friction). If we lump the shear stress in with the stress owing to friction and furthermore assume that the tangential stress is proportional to normal stress, we can write σ t = µσ n (15.63) where σt is the tangential stress σn is the normal stress μ is the proportionality constant (coefficient of sliding friction) Forrestal et al. [4] developed a formula for the axial force on an ogival nose: π 2 Fz = 2πs ∫ θ0  d   s−  sin θ −  2     s      σ (V , θ ) dθ + θ µ θ (cos sin    n z  (15.64) where d   s− 2 θ 0 = sin   s −1     (15.65) Here Vz is the instantaneous velocity during penetration. The stress function σn(Vz, θ) is assumed to be similar to that of a spherically symmetric expanding cavity (defined later). If we let V be the constant velocity at which the tip of the projectile radially expands the hole, then we can write the radial stress at the cavity surface as  ρt  σr = A + B  V  Y  Y  where σr is the radial stress Y is the material yield stress ρt is the target density A and B are constants defined as © 2014 by Taylor & Francis Group, LLC 2 (15.66) Ballistics: Theory and Design of Guns and Ammunition 384 A= n 2   2E   1 +   I 3   3Y   (15.67) 3 2 (15.68) B= where ( ) 1− 32YE I= ∫ 0 (− ln x)n dx 1− x (15.69) In these expressions, E is Young’s modulus and n is the strain hardening exponent (assumes power-law strain hardening). For an assumed incompressible 7075-T651 aluminum, Forrestal et al. [3] provide I = 3.896 and A = 4.609. Empirically, curve-fitting the stress–strain curves (thus including compressibility) for 7075-T651 yielded slightly different results with A = 4.418 and B = 1.068. To approximate the normal stress on the ogive, we can replace the spherically symmetric velocity, V in Equation 15.66 with Vz cos θ, then we have  ρt  σ n (Vz , θ ) = A + B  Vz , cos θ  Y  Y  2 (15.70) If we insert Equation 15.70 into Equation 15.64, we obtain π 2 Fz = 2π sY ∫ θ0      s − d2   ρt 2 2   sin θ −    (cos θ + µ sin θ   A + B Vz cos θ  dθ s Y        (15.71) Now we integrate to obtain Fz = π d2  ρ V2  Y α + β t z  4 Y   (15.72) where   π  α = A 1 + 4 µψ 2  − θ 0  − µ (2ψ − 1) 4ψ − 1  2     2  8ψ − 1  µ (2ψ − 1)(6ψ + 4ψ − 1) 4ψ − 1  2π − µψ θ β = B + − 0    2 24ψ 2 2   24ψ  (15.73) (15.74) Now that we have an expression for force as a function of velocity, we need to come up with how this varies during penetration. © 2014 by Taylor & Francis Group, LLC Penetration Theories 385 We can write Newton’s second law as dVz dt − Fz = m (15.75) We can convert this time integral to a distance integral and rewrite it as follows: dVz dz − Fz = mVz (15.76) One can write the mass of our projectile in terms of the parameters we have already described. The mass of the cylindrical section of the projectile is mcylinder = ρ p π d2 L 4 (15.77) We can write the mass of the ogive as mogive = ρ p π d3 k 8 (15.78) where  4ψ − 1  4ψ 1   k =  4ψ 2 − +  4ψ − 1 − 4ψ 2 (2ψ − 1)sin −1   3 3   2ψ  (15.79) Now the total mass of the projectile is m = mcylinder + mogive = ρ p π d2  kd  L+  4  2  (15.80) If we insert Equations 15.80 and 15.72 into Equation 15.76, we get, after some rearrangement kd  Vz  dVz −dz = ρ p  L +  + 2 Y α βρ tVz2   (15.81) This can be integrated as P 0 kd  Vz  dVz − dz = ρ p  L +  + 2 Y α βρ tVz2   V 0 ∫ ∫ (15.82) 0 The result of this integration is P= 1  ρp   kd    β   ρ tV02      L +  ln 1 +     2β  ρt   2    α   Y   where P is the final penetration depth V0 is the impact velocity © 2014 by Taylor & Francis Group, LLC (15.83) Ballistics: Theory and Design of Guns and Ammunition 386 If the penetration depth, P, is greater than the target thickness, perforation will occur. When this is the case, it is useful to be able to calculate the residual velocity of the penetrator, which we do by integrating Equation 15.82 with different limits of integration: Vr T kd  Vz  dVz − dz = ρ p  L +  2  α Y + βρ tVz2  V 0 ∫ ∫ (15.84) 0 where T is the target thickness. Performing the integration yields    αY 2βρ tT  α Y Vr =  + V02  exp  − − kd   βρ t  ρ p ( L + 2 )  βρ t (15.85) This model has proven to be fairly accurate (within 15%) once the coefficients have been tuned. It is fairly sensitive to the friction coefficient, μ, incorporated in both α and β, which Forrestal et al. [4] suggest should be between 0 and 0.06. Problem 1 A German 280 mm armor-piercing projectile weighs 666 lbm and is about 34 in. in length. It strikes a British warship in the 1/2 in. thick vertical side plating at an angle of 12° from horizontal along the path depicted below. The initial impact velocity is 2000 ft/s. Determine the residual velocity of the shell after passing through each compartment and how far through the ship it will go (i.e., in which compartment will it stop). Assume the density of the armor plate to be ρ = 0.283 lbm/in.3 7.00-in. thick 0.25-in. thick 0.50-in. thick Path of shell 1.25-in. thick 12° 4.00-in. thick (Assume normal to shell path) Answer: The projectile is arrested by the 1.25 in. deck. Problem 2 An explosively formed penetrator impacts a 4 in. thick RHA plate at a velocity of 1500 m/s. The penetrator parameters are given later. Determine if the penetrator will perforate the target using the Lambert/Zukas model given 1. A normal impact m Answer: Vl = 1299   yes s 2. An impact at 30° obliquity m Answer: Vl = 1389   yes s © 2014 by Taylor & Francis Group, LLC Penetration Theories 387 Penetrator information: l = 95 [mm] m = 1.25 [lbm] m d = 22 [mm] Vs = 1500   s Problem 3 A German 7.5 cm Gr 34A1 projectile is fired at a 2 in. thick armor plate at a 30° obliquity. The impact velocity is 400 m/s. The penetrator parameters are given later. 1. Determine whether the penetration mode will be plugging or piercing through use of the Jacobson model for a normal impact. Answer: Piercing 2. Determine if the penetrator will perforate the armor though use of the Lambert model. Answer: No 3. Comment on the validity of the model. Penetrator information: l = 39 [cm] m = 1.25 [kg] m d = 7.5 [cm] Vs = 400   s Problem 4 A Japanese 20 mm projectile with the properties given later impacts the 1/2 in. thick aluminum armor plate on a U.S. plane’s rear gun mount at 30° obliquity. If the projectile and the armor have the following properties: Determine how deep the projectile will penetrate into the armor (assume μ = 0.03). Answer: P = 53.1 [mm] = 2.09 [in.] 1. If the projectile perforates the armor, determine its residual velocity. m Answer: Vr = 423   s Estimated penetrator information: s = 40 [mm] m = 128 [g] m d = 20 [mm] Vs = 500   s  lbm  ρ p = 0.283  3   in.  L = 60 [mm] Estimated armor information:  lbm  A = 4.418Y = 39, 000 [psi] ρ t = 0.098  3   in.  B = 1.068 © 2014 by Taylor & Francis Group, LLC Ballistics: Theory and Design of Guns and Ammunition 388 Problem 5 You are asked to design a “slug butt” for a test gun. The requirement for thus device is to stop a 106 lbm steel projectile moving at 1000 m/s. You have available a pile of 2″ thick Rolled Homogeneous Armor (RHA) plates that you can stack together. Note that stacking plates together results in slightly worse performance than a solid plate—but we can account for this by adding some extra material to provide “margin.” Normally slug butts are constructed at an angle to deflect ricochets downward and improve ballistic performance but this decreases the “target” area. Given the information provided earlier and later, determine the number of plates you need to prevent a penetration and determine the angle of the slug butt—do not use an angle greater than 40° as the “target area” will be too small. List all assumptions and comment on your design. Projectile information: l = 40 [in] m = 106 [lbm] d = 155 [mm] m Vs = 1000   s  lbm  ρ p = 0.283  3   in.  Armor information:  lbm  ρs = 0.283  3  available in 2 in. thick plates  in.  Problem 6 A German 305 mm armor piercing projectile weighs 894 lbm and is about 35.2 in. in length. It strikes a British warship in the 3-1/4 in. thick turret crown at an angle of 20° from horizontal along the path depicted later. The initial impact velocity is 1800 ft/s. Determine the residual velocity of the shell after passing through each compartment and how far through the ship it will go (i.e., in which compartment will it stop). Assume the density of the armor plate to be ρ = 0.283 lbm/in.3. 3.25˝ thick 15° 9.00˝ thick 1.50˝ thick 20° 8.00˝ thick Path of shell 1.00˝ thick 6.00˝ thick © 2014 by Taylor & Francis Group, LLC Penetration Theories 389 Problem 7 A German 280 mm armor piercing projectile weighs 666 lbm and is about 34 in. in length. It strikes a British warship in the 0.43″ thick vertical side plating at an angle of 10° from horizontal along the path depicted later. The initial impact velocity is 1900 ft/s. Determine the residual velocity of the shell after passing through each compartment and how far through the ship it will go (i.e., in which compartment will it stop) Assume the density of the armor plate to be ρ = 0.283 lbm/in3. 3.25˝ thick 15° 9.00˝ thick 0.43˝ thick 1.50˝ thick 8.00˝ thick 10° Path of shell 1.00˝ thick 6.00˝ thick Problem 8 A British Short Magazine Lee Enfield (SMLE) is fired at a sniper plate across no-man’s land in WWI (because of the static nature of the fighting, snipers in the opposing lines fired through small holes in thick metal plates to minimize exposure). The projectile has a mass of 175 grains and the projectile is 1.1 in. long and made of lead (ignoring the copper jacket). The diameter is 0.310 in. The range to the target is 200 yards so we can assume an impact velocity of 1512 ft/s. The angle of impact is 12.85 min from the normal. Using the Lambert–Zukas model, determine the thickness of armor plate that the projectile will penetrate (i.e., obtain V50). Assume the density of the armor plate to be ρt = 0.283 lbm/in3. Assume the density of lead to be ρp = 0.407 lbm/in.3. Problem 9 Using the same information in Problem 8, determine how deep a projectile will penetrate into a 1 in. thick steel sniper plate assuming the impact is normal this time and the impact velocity is 1800 ft/s. Use the Tate formula. The additional target and bullet properties are as follows:  lbf  σ p = 10, 000  2   in.   lbf   lbf  σ t = 36,000  2  Et = 29,000,000  2   in.   in.  The Tate formula shows the energy balance between the projectile and the target. When solved for the penetration depth, the following equation results ρp P= Yp © 2014 by Taylor & Francis Group, LLC Vi ∫ u(v)l(v)dv vc Ballistics: Theory and Design of Guns and Ammunition 390 In this formula, u(v) is the instantaneous velocity of the base of the hole, v is the instantaneous velocity of the projectile and l(v) is the instantaneous projectile length. The limits of integration are between the impact velocity, Vi, and a cutoff velocity, vc, that depends on whether the projectile is stronger than the target or not. For this case, we can use vc = 2 Rt − Yp ρp for Rt > Yp The other terms are as follows: ) (  1  u(v) =  v − γ v2 + A γ = 2  − γ 1   A= ρt ρp 2(Rt − Tp )(1 − γ 2 ) ρt l(v)  v + v 2 + A =  Vi + Vi + A L      Rt − Yp γ Yp )( (  γρ p  v v2 + A − γ v2 − V V 2 + A − γ V 2 exp  i i i 2  2(1 − γ )Yp  )  where L is the initial length of the projectile. Comment on the results. Problem 10 At the end of World War I, the German navy surrendered to the British at Scapa Flow (a large anchorage in northern Scotland). When it appeared that surrender negotiations were breaking down, the German sailors opened the sea cocks (valves in the bottom of the ships) and sank most of their ships within sight of astounded British onlookers. In the 1920s, the British raised what ships they could and used them as targets to assess the penetration and bursting characteristic of their heavy shell [5]. One such test was against the side armor of S.M.S. Baden, the largest warship built by Germany during the war. We would like to examine two impacts against the armor of this vessel. In each case, the projectile data are given next: Projectile information [6]: l = 66 [in.] d = 15 [in.] s = 90 [in.] m = 1938 [lbm]  lbm  ρ p = 0.283  3   in.   ft  Vs = 1550   s Armor information:  lbm  ρs = 0.283  3   in.  © 2014 by Taylor & Francis Group, LLC Penetration Theories 391 For each of the following situations, determine the residual velocity of the projectile (if any): (a) Gun deck ( b) B-Turret roof θ = 16.5 [degrees] θ = 75.25 [degrees] t = 1.187 [in.] t = 4 [in.]  lbm  ρ t = 0.283  3   in.   lbm  ρ t = 0.283  3   in.  15.2 Penetration and Perforation of Concrete Concrete penetrating munitions have always been important in the military arsenal. Bunkers, buildings, and walls are used as cover by an enemy and it is required to perforate the structure and deliver some type of lethal or nonlethal effect behind the obstruction. Concrete comes in a variety of forms that have variable strengths, reinforcement geometry, and material properties owing to curing. Each of these forms behaves somewhat differently when impacted by a projectile. There is some evidence that once the impact velocity of a projectile is great enough, one can ignore reinforcement and only the concrete strength becomes important. As a consequence of the high compressive strength of concrete relative to its tensile strength, it tends to spall readily. A relatively simple model of projectile penetration into concrete was put forward by Forrestal et al. in 1994 [7]. This model has an advantage in its simplicity. But a slight disadvantage is that its empirical nature makes its global applicability somewhat limited. We shall use this model as a fairly good representation of concrete penetration physics. The model assumes normal impact of the projectile and that the projectile is rigid. This may seem to be restrictive assumptions; however, the method provides reasonable estimates for slightly yawed projectiles if the angle is below about 5° based on this author’s own work. The point of departure is the determination of the force on the nose of the projectile that is defined in a manner similar to a fluid mechanics analysis as F= π d2 (τ 0 A + NBρV 2 ) 4 (15.86) 8ψ − 1 24ψ 2 (15.87) With N defined as N= In these equations, the projectile properties are as follows (see Figure 15.7): d is the diameter of the projectile; ψ is the caliber-radius-head, defined in Equation 15.88; V is the projectile velocity (assuming rigid body motion); and s (used in Equation 15.88) is the ogive radius. © 2014 by Taylor & Francis Group, LLC Ballistics: Theory and Design of Guns and Ammunition 392 The caliber-radius-head is defined as ψ = s d (15.88) The target properties used in Equation 15.86 are as follows: ρ is the density of the target, the product τ 0A is a shear strength parameter obtained from a triaxial strength test, and B is a compressive strength parameter. In this model, the parameters are set as B=1 (15.89) τ 0 A = Sf c′ (15.90) Here S is a dimensionless empirical constant that depends upon the unconfined compressive strength f c′. If we define the instantaneous depth of penetration as z, we find that for z > 2d, we can write F= π d2 ( Sfc′ + NρV 2 ) , z > 2d 4 (15.91) This equation is valid for deep penetration depths. For depths less than two projectile diameters, the penetration process is affected by surface cratering. Beyond two projectile diameters, the hole caused by the projectile will be approximately equal to the projectile diameter. This is known as the tunnel region. We shall define the penetration depth as P. Below two projectile diameters, the damage to the concrete will, in general, be a conical taper called the crater. This is illustrated in Figure 15.8. In the surface crater region, the force on the projectile nose is proportional to the penetration depth or, mathematically F = cz , 0 < z < 2d (15.92) Here c is a constant, which we will soon define. Crater region Tunnel region Direction of penetration d 2d z FIGURE 15.8 Illustration of a concrete penetration. © 2014 by Taylor & Francis Group, LLC Penetration Theories 393 If we begin with Newton’s second law, we see that F = ma = m d2z dt 2 (15.93) Here m is the mass of the projectile. Since we know the force acting on the projectile will tend to slow it down, we can equate Equations 15.92 and 15.93: d2z = −cz dt 2 (15.94) d2z = −ω 2 z dt 2 (15.95) m We can rewrite Equation 15.94 as where we have defined ω2 = c m (15.96) If we assume a solution of the form z = A1sin ωt (15.97) dz = A1ω cos ωt dt (15.98) d2z = − A1ω 2 sin ωt dt 2 (15.99) we can write Our initial conditions are such that at t = 0, dz/dt = Vs, where Vs is our striking velocity, so Vs = A1ω → A1 = Vs ω (15.100) Then, we have for z < 2d z= © 2014 by Taylor & Francis Group, LLC Vs sin ωt ω (15.101) dz = Vs cos ωt dt (15.102) d2z = −ωVs sin ωt dt 2 (15.103) Ballistics: Theory and Design of Guns and Ammunition 394 We now use a compatibility condition that at z = 2d, both Equations 15.103 and 15.91 must yield the same answer. We shall call the time it takes the projectile to reach 2d, t1 and the velocity at that point will be V1, thus at z = 2d we have F t=t1 = πd 2 Sf c′ + NρV12 , z = 2d 4 ( ) d2z = −ωVs sin ωt1 , z = 2d dt 2 (15.104) (15.105) Since F = ma, we can combine the aforementioned equations to write mωVs sin ωt1 = πd 2 Sf c′ + NρV12 , z = 2d 4 ( ) (15.106) Also at t = t1, Equations 15.101 and 15.102 can be written as Vs sin ωt1 ω (15.107) V1 = Vscos ωt1 (15.108) 2d = We now rearrange Equation 15.108 to Vs = ω 2d cos ωt1 (15.109) Now insert Equation 15.109 into Equation 15.106 giving us mω 2 2 d = πd 2 Sf c′ + NρV12 4 ( ) (15.110) And if we make use of Equation 15.96, we can obtain c as c= πd Sfc′ + N ρV12 8 ( ) (15.111) We now need to find V1, which we do by squaring Equations 15.107 and 15.108 and adding them, resulting in V12 + 4cd 2 = Vs2sin 2 ωt1 + Vs2cos 2 ωt1 m (15.112) Making use of a trigonometric identity and rearranging brings us to c= © 2014 by Taylor & Francis Group, LLC ( m Vs2 − V12 4d 2 ) (15.113) Penetration Theories 395 If we now equate Equations 15.111 and 15.113, we get 2 mVs2 − πd 3 Sfc′ 2 m + πd 3 N ρ V12 = (15.114) Once we have V1 and c, the determination of the time t1 is found simply through use of Equation 15.108: t1 = 1 V  cos −1  1  = ω  Vs  m V  cos −1  1  c  Vs  (15.115) To summarize the analysis procedure for the crater region, we must first find V1 through use of Equation 15.114, then we find c through use of Equation 15.113, and finally, we find t1 through use of Equation 15.115. If V goes to zero before time, t1 is reached, the projectile never penetrates deeper than the crater region and our analysis would be complete. The depth of penetration in this case would be found from Equation 15.102: V = 0 = Vs cos ωt (15.116) This would occur when ωt = π π m ฀t= ฀sin ωt = 1 2 2 c (15.117) If we insert this result into Equation 15.101, we obtain the achieved depth of penetration, P: P = Vs m c (15.118) The striking velocity that would make this true would be determined from Equation 15.114 with V1 set equal to zero. So for a projectile to stop before creating a tunnel, the velocity is given by VsNotunnel ≤ 2πd 3Sf c′ 2m (15.119) If the projectile penetrates beyond 2 diameters into the concrete, it will enter the so-called tunnel region. When the projectile continues into the tunnel region, there is a change in the governing equation as discussed earlier. To determine the depth of penetration, we begin by combining Equations 15.91 and 15.93 to obtain m d 2 z πd 2 Sf c′ + NρV 2 , 2d < z < P = dt 2 4 ( ) (15.120) We can transform our independent variable from time to distance and we can write mV © 2014 by Taylor & Francis Group, LLC dV πd 2 Sf c′ + NρV 2 , 2d < z < P = dz 4 ( ) (15.121) Ballistics: Theory and Design of Guns and Ammunition 396 If we rewrite Equation 15.121 as follows: dV πd 2  Sf c′  = + Nρ V  dz 4m  V  (15.122) Now we integrate it from V1 to zero and 2d to P, so we can write 0 0  P dV    dz dV =  dz  V1  2d  ∫ ∫ V1  P πd 2 Sf ′   c   + NρV  dz  dV   4m  V    2d  ∫ ∫ (15.123) which results in P=  NρV12  2m ln  1+  − 2d , 2d < P 2 πd Nρ Sf c′   (15.124) If we have determined through use of Equations 15.113 through 15.115 that a projectile will penetrate beyond the tunnel region, we can write a procedure to determine the depth of penetration as follows. First, calculate V1, c, and t1 as described earlier for the crater region. Then calculate P from Equation 15.124. If this is greater than the concrete thickness, the projectile will perforate. If not, the projectile will penetrate to depth P. It would be good to see if a spall thickness is created (as will be described in Section 16.3) by the impact and if this is the case, we could add the spall thickness to P and perforation may still result (though with low residual velocity). Since the dimensionless parameter S is obtained or verified experimentally, it would be nice to know how close we came to our estimate by direct calculation. If one had an experiment where a given projectile penetrated to depth P, we can back calculate S as follows. We start with Equation 15.124 and rearrange thusly S= 1 NρV12 f c′   πd 2 Nρ   exp (P − 2d)  − 1 2m     (15.125) In an experiment, we usually are given the striking velocity, so we want to replace V1 in this equation with Vs, so we use Equation 15.114: S= ( ) N ρ 2 mVs2 − πd 3 Sfc′ fc′ (2 m + πd N ρ ) 3 1  πd 2 N ρ    − 1 exp ( P − 2 d) 2 m     (15.126) which can be simplified to  π d 3 Sfc′  N ρVs2  1 − 2 mVs2   1 S= 3  πd N ρ    πd 2 N ρ   fc′  1 + − 1 exp ( P − 2 d)   2 m    2 m    © 2014 by Taylor & Francis Group, LLC (15.127) Penetration Theories 397 Dimensionless empirical constant versus unconfined compressive strength 25 S (dimensionless) 20 15 y = 86.431x –0.5158 10 y = 93.48x –0.5603 y = 103.71x –0.6142 5 0 0 20 40 60 80 100 120 f ΄c(MPa) FIGURE 15.9 Determination of dimensionless parameter S for Forrestal et al. [8] concrete penetration model. With this equation, one can find S if you know the striking velocity and the concrete strength. Forrestal et al. [7] have calibrated this equation with several experiments. A reproduction of their chart is shown in Figure 15.9 with the addition of upper and lower bounds based on their data. The equation used to determine S given the unconfined compressive strength f c′ is S = 93.48 f c′−0.5603 (15.128) Here recall that f c′ is in MPa and S is dimensionless. Bounding equations are shown in Figure 15.9. These equations were obtained through use of a curve fit routine. Problem 11 A 0.50 caliber projectile is fired at an extremely thick concrete wall of 2100 psi unconfined compressive strength and density of 0.084 lbm/in.3 It strikes with no obliquity and a 2000 ft/s velocity. How far does it penetrate? Answer: P = 20.4 [cm] Projectile information: s = 63.50 [mm] m = 662 [grains]  ft  d = 12.70 [mm] Vs = 2000   s © 2014 by Taylor & Francis Group, LLC 398 Ballistics: Theory and Design of Guns and Ammunition Problem 12 Areal density is a measure of the mass of one square inch (or sometimes a square foot or meter) of a material used as armor. Given a threat projectile with the properties provided later, find the armor solution with the lowest areal density that prevents perforation if your choices are steel with density 0.283 lbm/in.3, concrete of 2500 psi unconfined compressive strength and density of 0.084 lbm/in.3, and aluminum with the following properties: A = 4.418 B = 1.068 Y = 39, 000 [psi]  lbm  ρ t = 0.098  3   in.  Assume the projectile strikes with 15° of obliquity. Penetrator information: s = 35 [mm] d = 12 [mm] L = 15 [mm] m = 0.03 [kg] m Vs = 600   s  lbm  ρ p = 0.283  3   in.  Problem 13 We would like to compare defenses against the projectile in Problem 10. Assuming this projectile impacts a steel plate at zero degrees obliquity a. The thickness of the armor required to prevent penetration. b. The thickness of 1500 psi unconfined compressive strength concrete required to do the same assuming that the concrete does not spall—assume a density of 0.084 lbm/in.3 for the concrete—note that this particular shell has a secant ogive so as an estimate for the purposes of this problem—divide the resultant ogive length by 2—please note that this is simply a guess and not based on physics. c. Comment on the validity of the answer to part b. © 2014 by Taylor & Francis Group, LLC Penetration Theories 399 15.3 Penetration and Perforation of Soils In recent times, the penetration of soils has gained importance in the terminal ballistic field. Enemy strong points have been encountered below a soil layer. Land mines need to be defeated below various types of soils as well. It is therefore necessary to determine the penetration capabilities of projectiles into soils with the intention of defeating a buried target. As a reasonable approach to determine soil penetration, we shall use the method of Forrestal and Luk [8]. While other approaches exist, this rather simple procedure is excellent for introducing the physics of the problem. Soils vary widely in their behavior under penetration loadings. Because the behavior is somewhat complicated, more parameters are needed to describe a soil than a material such as a metal. The first thing we have to realize is that soil can be in a state where the density is less than its locked density. The locked density is where the soil behaves like a solid or fluid in compression (i.e., its states are defined by a hydrostat). We therefore need to introduce two densities: p0, its initial density and ρ*, its locked density. We also need to define η*, its locked volumetric strain. Here we define η* as η* = 1 − ρ0 ρ* (15.129) Two typical models used for soils come directly from our failure theories of structures. They are the Tresca (maximum shear stress) theory and the Mohr–Coulomb theory of failure. Both of these were introduced in Section 4.2. Here we shall use a combination of the two. A Mohr–Coulomb yield criteria with a Tresca flow rule. For the Tresca criterion, once a shear stress failure level is achieved, the material strength is not increased with increasing load. With the Mohr–Coulomb criterion, the yield stress in the material increases with compressive load. The combination of the two allows the material to resist more load as compression is applied up to a point, then further increase in the compressive loading will not affect the material strength. Similar to the aluminum penetration model, we again define the caliber-radius-head as s d (15.130) d 4ψ − 1 2 (15.131) ψ = where d is the diameter of the projectile ψ is the caliber-radius-head s is the ogive radius We again define nose length as l= The method considers the resistance force of the soil on the penetrator to have two components: a normal force (normal stresses) and a tangential force (shear stresses and friction). © 2014 by Taylor & Francis Group, LLC Ballistics: Theory and Design of Guns and Ammunition 400 If we lump the shear stress in with the stress owing to friction and furthermore assume that the tangential stress is proportional to normal stress, we can again write σ t = µσ n (15.132) where σt is the tangential stress σn is the normal stress μ is the proportionality constant (a coefficient of sliding friction) Forrestal et al. [4] developed a formula for the axial force on an ogival nose that we introduced in Section 15.1 and we again use here π /2 Fz = 2π s     s − d2     (cos θ + µ sin θ ) σ n (Vz ,θ )dθ s   ∫ sin θ −  θ0 (15.133) where  s − d2 θ 0 = sin −1   s  −1  2ψ − 1    = sin    2ψ  (15.134) Here Vz is the instantaneous velocity during penetration. The stress function σn (Vz, θ) is assumed to be similar to that of a spherically symmetric expanding cavity. At this point, we are going to depart from the mathematics to look at the penetration event in a qualitative manner. Let us assume that we are at some axial location in the ogive of the projectile and we are looking in the direction of penetration at time, t. What we would see is illustrated in Figure 15.10. The projectile would be opening a cavity at a rate that we shall call Vt. The plastic zone would be expanding at some rate ct. Here c is the speed of the plastic wave (dependent upon the Hugoniot jump conditions to be discussed Elastic region c1 c Plastic region Undisturbed region Vt Projectile FIGURE 15.10 Elastic and plastic compression zones at a section of an ogive penetrating into soil looking in the direction of penetration. © 2014 by Taylor & Francis Group, LLC Penetration Theories 401 in Section 16.1). The elastic zone would be expanding at a rate c1t. Here c1 is the speed of the dilatational wave in the material. It can be shown that V and c are related (as per the shock theory that will follow in Chapter 16) and we can define a parameter, γ as 1 3 3 τ  V  γ = =  1 + c  − (1 − η *) c  2E   (15.135) where E is Young’s modulus. With the aforementioned physics, Forrestal and Luk [6] derived material response models for each of the three failure models we have discussed earlier. For the detailed derivation, the interested reader is referred to that paper. The basic idea was to have a general function for the force acting on the projectile nose that we can integrate using Newton’s second law to obtain the velocity and penetration distance as a function of time. If we insert expressions that relate the radial expansion velocity of the cavity, V, to the projectile penetration velocity, Vz, we can put the expression for the retarding force in this form Fz = α s + β s Vz2 (15.136) where   πd 2 π  τ c A 1+ 4 µψ 2  − θ 0  − µ (2ψ − 1) (4ψ − 1)  4 2     (15.137) 2  8ψ − 1 πd 2  µ (2ψ −1)(6ψ + 4ψ − 1) 4ψ − 1  2π − + − µψ θ ρ 0B  0    2 4 24ψ 2 2   24ψ  (15.138) αs = βs = Here the coefficients A and B are dependent upon the material model used for the soil. Recall that the definition for the Tresca criterion implies that once a material reaches its state of maximum shear stress, it begins to deform plastically and cannot support any more load. For a soil that behaves in a Tresca type manner, we have 3    τc   − (1 − η *)   1+  2 2E    A = 1 − ln   3   3 τc     1+   2E      (15.139) 1 3  3 τc  3τ c 3τ c   η ( *) − − + 1 1     +η *  1 −  2E  3   2E  E  + B= 2 − 4 2(1 − η *) τ   3  3 21+ c  τc  1 1 η ( *) − − + 2E      2E    2 © 2014 by Taylor & Francis Group, LLC 3  τc     31+   2E   1 +   (1 − η *)      (15.140) Ballistics: Theory and Design of Guns and Ammunition 402 Recall that the definition for the Mohr–Coulomb criterion implies that as the compressive forces increase, it becomes harder to have the material fail in shear. For a soil that behaves in a Mohr–Coulomb type manner, we have τc   1  1 + 2E  A=   α γ  B= 2α − 1 λ (15.141) 3 (1 − η *)(1 − 2α )(2 − α ) τc   1  1 + 2E  + 2  γ  γ    2α   2   3 3 3 τ τ γ [ ( η *)( α ) + 3 γ ] 3 2 1 − 2 −   c   +η *  1 − c  −  4 E E 2   τ    (1 − η *)(1 − 2α )(2 − α )  1 + c    2E    (15.142) Also note that the Tresca criterion behaves the same as the Mohr–Coulomb criteria with λ = 0. We define α= 3λ 3 + 2λ (15.143) Because of a singularity in the governing equations, there is a special set of equations for the Mohr–Coulomb criterion when we have λ = 3/4. In this case, τc    1+ 2E  4 A = 2 −  γ  3   2 τ c   3τ c 3τ c     η * 1 + − + 1      2E   E 2E   2  −2 ln γ  + − B= (1 − η *) γ3 3  τ   3 ln  1 + c    1 2E     3 −   ( *) η 1 − τc  1 +   2E    (15.144) (15.145) When a material behaves according to the model that combines both Mohr–Coulomb and Tresca behaviors, things become slightly more complicated. The parameters A and B will be dependent upon the rate of loading. One must keep in mind that this failure criterion implies that up to some stress level, the material will have improved resistance to compressive loading because of the internal friction of the grains and after a limit load is reached (τm), the material simply yields regardless of load. Thus, we can consider three velocity regimes: V < Vmin, where the yielding is completely Mohr–Coulomb behavior; Vmin < V < Vmax, where the yielding closest to the projectile is by Tresca criterion and the yielding near the elastic–plastic interface is according to the Mohr–Coulomb criterion; and V > Vmax, © 2014 by Taylor & Francis Group, LLC Penetration Theories 403 where the entire yield region is according to the Tresca model. We shall consider each of these cases. If V < Vmin, we stated that the yielding is completely according to the Mohr–Coulomb model. Thus, Equations 15.141 through 15.145 apply. The equation required to determine Vmin is Vmin 2α  τc    1+  τ c τ m  2E    − =  αρ0B  τ c  γ       (15.146) Recall that τm is the stress level at which the material behaves according to the Tresca model. We shall discuss how we determine V shortly. If Vmin < V < Vmax, the zone of yielding material has two subzones: a zone next to the projectile that behaves according to the Tresca model and a zone next to the elastic region that behaves according to the Mohr–Coulomb model. We shall first write the equation for Vmax. Vmax =  1  τm    3τ c 3τ c    α  τ c  ρ0  +η *  1 −   2E     E τ cγ 2 2  1 2 − λ − 3   (15.147) If we define a coordinate, ξ, that varies from 0 at the projectile surface to 1 at the elastic– plastic interface, we can determine a coordinate, ξm, where the yield behavior changes from Tresca to Mohr–Coulomb. Unfortunately, this crossover point has to be solved numerically with the equation that follows: τc   1+  E 2  2α (1 − η *)ξ m3 + γ 3  + 2α 3   γ 3  2(1 − η *)(2 − α )ξ m3 + 3γ 3   αρ 0V 2  + 4  2  τ cγ  (1 − η *)(1 − 2α )(2 − α ) (1 − η *)ξ m3 + γ 3  3    τ   α ρ 0V 2  1 + c  2 E  2α τ cγ 2 (1 − η *)ξ m3 + γ 3  2α 3   3 3 2    γ  2(1 − η *)(2 − α ) + 3γ   τ m 3τ   3τ c  +η *  1 − c  − =0  4 − 2E   τc   τc   E (1 − η *)(1 − 2α )(2 − α )  1 +   2E    (15.148) Keep in mind here that we know all of the information (including V) and we are solving for ξm. A good math code will generally solve this equation quickly. Once we have ξm, then A and B are given at the projectile surface (Tresca) by A= 1  τm α  τ c © 2014 by Taylor & Francis Group, LLC 3  ξm    1 2  τm    η ln 1 + ( *) − + − 1     λ 3τ    c    γ   (15.149) Ballistics: Theory and Design of Guns and Ammunition 404   3   ξm    3 + 4(1 − η *)    γ   1   B= 3 − 4  2(1 − η *)  3 3    ξ   1 + (1 − η *)  m       γ     (15.150) These equations account for the fact that the yielding is Mohr–Coulomb outside of ξ = ξm. If V > Vmax, the yielding is completely according to the Tresca model. Thus, A and B are given by A= 2 τ − 2 m 3  τc γ     ln    τ    1+ c    2E    2 3τ c 3τ   +η *  1 − c  γ 3 2E  E  B= − + 4 2 γ 2(1 − η *) τ   21+ c  2E   (15.151) 3  τc    + 3 1     2E   1 +   (1 − η *)  (15.152) To approximate the normal stress on the ogive, we can replace the spherically symmetric velocity, V, in our previous equations with Vz cos θ. We can write an equation for the normal stress function on the ogive [6] as σ n (Vz , θ ) = τ c A + ρ0B[Vzcos θ ] 2 (15.153) We can write Newton’s second law as − Fz = m dVz dt (15.154) We can then convert this time integral to a distance integral as before to yield − Fz = mVz dVz dz (15.155) If we substitute Equation 15.136 into the aforementioned equation and integrate, we get an equation for the acceleration, velocity, and depth of penetration, respectively, as a function of time: a=−   cos  tan −1    2 © 2014 by Taylor & Francis Group, LLC αs m βs  t α s βs  V0  −  α s  m  (15.156) Penetration Theories 405 Vz =   βs  t α s βs  αs tan  tan −1  V0  −  βs m   αs      βs  t α s βs  cos  tan −1  V0  − m  m   αs  ln  z=  βs  βs   V0   cos  tan −1     α s            (15.157) (15.158) If we determine the distance where the velocity of the projectile slows to zero, we obtain the depth of penetration as P=  β V2  m ln  1 + s 0  2β s  αs  (15.159) where P is the final penetration depth V0 is the impact velocity So now that we have developed penetration formulas for soils, what do we do with them? The use of models such as this one, as nice as it is, usually carries with it some practical issues. A detailed model like this requires detailed material properties that, in practice, one rarely has. It usually will require a test or two to calibrate it. Forrestal and Luk [8] suggest using a value of 0.13 for η*. The authors claim the model is relatively insensitive to it. The model was derived for normal penetration, but the authors claim good results up to impact yaw angles of 30°. In this case, they used the line-of-sight penetration depth. As one might expect, the accuracy of this particular model varies significantly with the properties of the soil. Rocks, roots, and soil layers further complicate everything. Nevertheless, the model is an excellent tool and describes the physics of soil penetrations well. This is a highly active area of current research. Problem 14 A 0.50 caliber projectile is fired at a soil berm with properties established later. How far does it penetrate? Answer: P = 84.9 [cm] Projectile information: s = 63.50 [mm] m = 662 [grains] d =12.70 [mm]  ft  Vs = 2000   s © 2014 by Taylor & Francis Group, LLC Ballistics: Theory and Design of Guns and Ammunition 406 Soil information (assume Mohr–Coulomb behavior): Initial density  kg  ρ0 = 1860  3  m  µ = 0.1 Locked density  kg  ρ * = 2125  3  m  λ = 0.33  lbf  τ c = 1500  2   in.   lbf  E = 2×107  2   in.   lbf  τ m = 2500  2   in.  15.4 Penetration and Perforation of Ceramics The desire to decrease the weight of vehicles coupled with constant improvements in manufacture has increased interest in the use of ceramics as armor. The design of an armored vehicle using ceramics requires an understanding of their behavior under impact loads. The advantages of ceramic armor are its relatively low density, high hardness, and high compressive strength. The disadvantages are that ceramics are usually brittle, have low tensile strength that when coupled with high compressive strength can be a problem from a spallation standpoint, they allow the protection to be degraded in a multihit situation, and they are somewhat expensive. Their complex structural behavior makes them difficult to model although this is only a disadvantage to the designers. The response of a ceramic to penetration is unique amongst all of the other materials discussed in this text. The material behaves differently depending on the radial confinement and whether it is backed or not. For reasons such as spallation, they are usually backed by a fiber reinforced composite, plastic, elastomer, or metal plate. If a ceramic is not backed, it will most likely spall when subjected to a high-shock load. This spallation can be analyzed by the techniques we will discuss in Chapter 16 on shock theory. Although not exhaustive, this is a list of common ceramics currently either in use or being studied for armor applications: Boron carbide (B4C) Silicon carbide (SiC) Titanium di-boride (TiB2) Aluminum nitride (AlN) Alumina (Al2O3) Historically, terra cotta (ceramic) armor has been found in Chinese tombs dating from 400 BC. Before First World War, the practice of placing coal bunkers around magazines to take advantage of comminution, a phenomena that we shall discuss shortly. If a ceramic is backed, one can take advantage of its high compressive strength to resist penetration. This will cause the tip of the penetrator to deform. Large stresses then build © 2014 by Taylor & Francis Group, LLC Penetration Theories 407 up in the penetrator. If the striking velocity is low enough, the penetrator will break up or ricochet. This process is called interface defeat or infinite dwell. If the penetrator survives the initial impact, the ceramic begins to fail. This process is complicated, which is why it is difficult to model, but it is key to understanding the behavior and utilizing the ceramic to the maximum extent possible. The ceramic penetration process has been documented by Cheeseman [9]. After an initial dwell and several reflections of the shocks and rarefactions, the following events occur and will either continue to perforation or stop when the penetration is arrested. Initially, tensile cracks appear near the penetrator forming circular rings. These cracks propagate along the principal stress planes that are usually 25°–75° from the surface normal. Once the cracks reach the distal boundary, they coalesce into conical form. At this point, if the ceramic was not backed, a plug would be ejected and the material would be perforated. If the plate is backed, then at the time when the conoid is formed, the stress is redistributed circumferentially and radial cracks appear. After the appearance of radial cracks, lateral cracking in the plane of the impact surface forms. This process is illustrated in Figure 15.11. With backing material present that holds the ceramic plug in place, the material has nowhere to go so micro-cracking begins. This pulverizes the ceramic material. This is known as the comminuted zone. The process of comminution and the sand-like character of the comminuted material erode the penetrator at a rapid rate. The powdered material continually gets in the way of the penetrator. This material flows radially outward and rearward. A similar effect occurs during shaped charge jet penetration into sand bags. The penetration of a ceramic armor is highly dependent upon the boundary conditions. It is known that confinement increases the penetrati