Gennady Stupakov Gregory Penn Classical Mechanics and...

277
Classical Mechanics and Electromagnetism in Accelerator Physics Gennady Stupakov Gregory Penn Graduate Texts in Physics

Transcript of Gennady Stupakov Gregory Penn Classical Mechanics and...

Page 1: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

Classical Mechanics and Electromagnetism in Accelerator Physics

Gennady StupakovGregory Penn

Graduate Texts in Physics

Page 2: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

Graduate Texts in Physics

Series editors

Kurt H. Becker, Polytechnic School of Engineering, Brooklyn, USAJean-Marc Di Meglio, Université Paris Diderot, Paris, FranceSadri Hassani, Illinois State University, Normal, USABill Munro, NTT Basic Research Laboratories, Atsugi, JapanRichard Needs, University of Cambridge, Cambridge, UKWilliam T. Rhodes, Florida Atlantic University, Boca Raton, USASusan Scott, Australian National University, Acton, AustraliaH. Eugene Stanley, Boston University, Boston, USAMartin Stutzmann, TU München, Garching, GermanyAndreas Wipf, Friedrich-Schiller-Universität Jena, Jena, Germany

Page 3: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

Graduate Texts in Physics

Graduate Texts in Physics publishes core learning/teachingmaterial for graduate- andadvanced-level undergraduate courses on topics of current and emerging fields withinphysics, both pure and applied. These textbooks serve students at the MS- orPhD-level and their instructors as comprehensive sources of principles, definitions,derivations, experiments and applications (as relevant) for their mastery and teaching,respectively. International in scope and relevance, the textbooks correspond to coursesyllabi sufficiently to serve as required reading. Their didactic style, comprehensive-ness and coverage of fundamental material also make them suitable as introductionsor references for scientists entering, or requiring timely knowledge of, a research field.

More information about this series at http://www.springer.com/series/8431

Page 4: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

Gennady Stupakov • Gregory Penn

Classical Mechanicsand Electromagnetismin Accelerator Physics

123

Page 5: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

Gennady StupakovSLAC National Accelerator LaboratoryStanford UniversityMenlo Park, CAUSA

Gregory PennLawrence Berkeley National LaboratoryBerkeley, CAUSA

ISSN 1868-4513 ISSN 1868-4521 (electronic)Graduate Texts in PhysicsISBN 978-3-319-90187-9 ISBN 978-3-319-90188-6 (eBook)https://doi.org/10.1007/978-3-319-90188-6

Library of Congress Control Number: 2018939001

© Springer International Publishing AG, part of Springer Nature 2018This work is subject to copyright. All rights are reserved by the Publisher, whether the whole or partof the material is concerned, specifically the rights of translation, reprinting, reuse of illustrations,recitation, broadcasting, reproduction on microfilms or in any other physical way, and transmissionor information storage and retrieval, electronic adaptation, computer software, or by similar or dissimilarmethodology now known or hereafter developed.The use of general descriptive names, registered names, trademarks, service marks, etc. in thispublication does not imply, even in the absence of a specific statement, that such names are exempt fromthe relevant protective laws and regulations and therefore free for general use.The publisher, the authors and the editors are safe to assume that the advice and information in thisbook are believed to be true and accurate at the date of publication. Neither the publisher nor theauthors or the editors give a warranty, express or implied, with respect to the material contained herein orfor any errors or omissions that may have been made. The publisher remains neutral with regard tojurisdictional claims in published maps and institutional affiliations.

Printed on acid-free paper

This Springer imprint is published by the registered company Springer International Publishing AGpart of Springer NatureThe registered company address is: Gewerbestrasse 11, 6330 Cham, Switzerland

Page 6: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

Preface

This book is a graduate-level textbook covering topics in classical mechanics andelectromagnetic fields that are pertinent for the design and operation of particleaccelerators. It was conceived as an intermediate step between general graduateuniversity courses on these two subjects and specialized textbooks on acceleratorphysics. Our goal was to cover a selected number of subjects that we consider as anessential part of the knowledge that would help a student to better understand beamdynamics, collective effects and electromagnetic radiation of relativistic beams inaccelerators.

The need for this type of course was first recognized by Richard Talman ofCornell University and Helmut Wiedemann of Stanford University, who taughtthis class in 2002 and 2004, respectively, at the US Particle Accelerator School(USPAS). The authors of this book offered this course as part of the USPAS programfrom 2007 to 2016 on a roughly biannual basis. It assumes an undergraduate-levelbackground in mechanics and special relativity, and in electricity and magnetism.Matrix algebra, calculus, and complex variables are used throughout, again at anadvanced undergraduate level.

In the presentation of the material, we made an effort to avoid lengthy deriva-tions which can be found in existing textbooks. Instead, we tried to focus on majorconcepts and to connect those concepts to practical applications for accelerators.Fundamental notions of mechanics played a key role in the invention of the particleaccelerator and continue to inspire new developments. Similarly, the electromag-netic fields produced by relativistic beams are both useful diagnostic signals and akey benefit of accelerator-based “light sources,” while at the same time they imposesignificant constraints on the performance of accelerators.

The book consists of two parts. Part I is devoted to classical mechanics. InChaps. 1–4, we first cover the basics of Lagrangian and Hamiltonian formalism,action-angle variables, and then linear and nonlinear oscillators. Starting fromChap. 5, we introduce specific features of accelerators. In Chaps. 5 and 6, we derivethe Hamiltonian for a circular accelerator and formulate equations of motion, and inChap. 7, we apply the action-angle formalism to the betatron oscillations in anaccelerator. Having developed the basic machinery, we then apply it to the topics of

v

Page 7: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

field errors in Chap. 8 and nonlinear resonances in circular accelerators in Chap. 9.The first part of the book ends with Chap. 10, where we derive the kinetic equationgoverning the dynamics of beams and discuss some of its properties.

Part II of the book focuses on classical electromagnetism and begins with adiscussion of the electromagnetic field of relativistic beams moving in free space inChap. 11, and then, in Chap. 12, we introduce the effect of the material environ-ment, obtaining the fields of a beam propagating inside a round pipe with resistivewalls. In Chap. 13, we review plane electromagnetic waves and Gaussian beams asexamples of the field propagating in free space. In Chap. 14, we expand the dis-cussion to the modes in waveguides and radio-frequency cavities. A large fractionof the second part of the book deals, in Chaps. 15–20, with the radiation processesof relativistic beams in different conditions. We conclude this part with a discussionof laser-driven acceleration of charged particles in Chap. 21 and the radiationdamping effect in Chap. 22. Selected references are included in some chapters formore detailed explanations of technical details and as a starting point for furtherreading.

We would like to thank our colleagues, who have been involved in the teachingof this class, and the organizers of the USPAS program. We are also grateful to thestudents who provided useful feedback in the development of this course. We areespecially thankful to Chris Mayes who read and commented on several chaptersof the book. We appreciate many valuable discussions with Alex Chao andMax Zolotorev on fundamental subjects of accelerator physics.

Cupertino, USA Gennady StupakovOakland, USA Gregory Penn

vi Preface

Page 8: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

Contents

Part I Classical Mechanics

1 The Basic Formulation of Mechanics: Lagrangianand Hamiltonian Equations of Motion . . . . . . . . . . . . . . . . . . . . . . 31.1 Lagrangian . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 31.2 Lagrangian of a Relativistic Particle in an Electromagnetic

Field . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 71.3 From Lagrangian to Hamiltonian . . . . . . . . . . . . . . . . . . . . . . . 91.4 Hamiltonian of a Charged Particle in an Electromagnetic

Field . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 101.5 The Poisson Bracket . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19

2 Canonical Transformations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 212.1 Canonical Transformations . . . . . . . . . . . . . . . . . . . . . . . . . . . 222.2 Poisson Brackets and Canonical Transformations . . . . . . . . . . . 232.3 Generating Functions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 252.4 Transformations with a Time Dependence and Four

Types of Generating Functions . . . . . . . . . . . . . . . . . . . . . . . . 26

3 Action-Angle Variables and Liouville’s Theorem . . . . . . . . . . . . . . 313.1 Canonical Transformation for a Linear Oscillator . . . . . . . . . . . 313.2 Action-Angle Variables in 1D . . . . . . . . . . . . . . . . . . . . . . . . . 333.3 Hamiltonian Flow in Phase Space and Symplectic Maps . . . . . . 363.4 Liouville’s Theorem . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 383.5 Non-conservative Forces in Hamiltonian Dynamics . . . . . . . . . . 39References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 45

4 Linear and Nonlinear Oscillators . . . . . . . . . . . . . . . . . . . . . . . . . . 474.1 Harmonic Oscillator Without and with Damping . . . . . . . . . . . . 474.2 Resonance in a Damped Oscillator . . . . . . . . . . . . . . . . . . . . . . 49

vii

Page 9: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

4.3 Random Kicks . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 494.4 Parametric Resonance and Slow Variation of the Oscillator

Parameters . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 514.5 Nonlinear Oscillator and Nonlinear Resonance . . . . . . . . . . . . . 53References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 61

5 Coordinate System and Hamiltonian for a CircularAccelerator . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 635.1 Coordinate System . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 635.2 Hamiltonian in Curvilinear Coordinate System . . . . . . . . . . . . . 655.3 Using s as Time Variable . . . . . . . . . . . . . . . . . . . . . . . . . . . . 665.4 Small Amplitude Approximation . . . . . . . . . . . . . . . . . . . . . . . 685.5 Time-Dependent Hamiltonian . . . . . . . . . . . . . . . . . . . . . . . . . 70References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 74

6 Equations of Motion in Accelerators . . . . . . . . . . . . . . . . . . . . . . . . 756.1 Vector Potential for Different Types of Magnets . . . . . . . . . . . . 756.2 Taylor Expansion of the Hamiltonian . . . . . . . . . . . . . . . . . . . . 776.3 Hill’s Equation, Betatron Function and Betatron Phase . . . . . . . 79Reference . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 86

7 Action-Angle Variables for Betatron Oscillations . . . . . . . . . . . . . . 877.1 Action-Angle Variables . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 877.2 Eliminating Phase Oscillations . . . . . . . . . . . . . . . . . . . . . . . . . 907.3 Phase Space Motion at a Given Location . . . . . . . . . . . . . . . . . 91Reference . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 94

8 Magnetic Field and Energy Errors . . . . . . . . . . . . . . . . . . . . . . . . . 958.1 Closed Orbit Distortions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 958.2 Effect of Energy Deviation . . . . . . . . . . . . . . . . . . . . . . . . . . . 988.3 Quadrupole Errors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 99

9 Nonlinear Resonance and Resonance Overlapping . . . . . . . . . . . . . 1079.1 The Third-Order Resonance . . . . . . . . . . . . . . . . . . . . . . . . . . . 1079.2 Standard Model and Resonance Overlapping . . . . . . . . . . . . . . 1129.3 Dynamic Aperture in Accelerators . . . . . . . . . . . . . . . . . . . . . . 115References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 119

10 The Kinetic Equation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12110.1 The Distribution Function in Phase Space and the Kinetic

Equation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12110.2 Integration of the Vlasov Equation Along Trajectories . . . . . . . 12510.3 Action-Angle Variables in the Vlasov Equation . . . . . . . . . . . . 12710.4 Phase Mixing . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12810.5 Damping and Stochastic Motion . . . . . . . . . . . . . . . . . . . . . . . 129

viii Contents

Page 10: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

Part II Electricity and Magnetism

11 Self Field of a Relativistic Beam . . . . . . . . . . . . . . . . . . . . . . . . . . . 13711.1 Relativistic Field of a Particle Moving with Constant

Velocity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13711.2 Interaction of Moving Charges in Free Space . . . . . . . . . . . . . . 14011.3 Field of a Relativistic Bunch of Particles . . . . . . . . . . . . . . . . . 14111.4 Electric Field of a 3D Gaussian Distribution . . . . . . . . . . . . . . . 144References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 150

12 Effect of Environment on Electromagnetic Field of a Beam . . . . . . 15112.1 Skin Effect and the Leontovich Boundary Condition . . . . . . . . . 15112.2 Perfectly Conducting Boundary Conditions . . . . . . . . . . . . . . . 15412.3 Round Pipe with Resistive Walls . . . . . . . . . . . . . . . . . . . . . . . 155Reference . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 161

13 Plane Electromagnetic Waves and Gaussian Beams . . . . . . . . . . . . 16313.1 Plane Electromagnetic Waves . . . . . . . . . . . . . . . . . . . . . . . . . 16313.2 Gaussian Beams . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 165

14 Waveguides and RF Cavities . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17314.1 TM Modes in Cylindrical Waveguides . . . . . . . . . . . . . . . . . . . 17314.2 TE Modes in Cylindrical Waveguides . . . . . . . . . . . . . . . . . . . 17514.3 RF Modes in Cylindrical Resonators . . . . . . . . . . . . . . . . . . . . 17614.4 Electromagnetic Field Pressure . . . . . . . . . . . . . . . . . . . . . . . . 17914.5 Slater’s Formula . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18114.6 Excitation of a Cavity Mode by a Beam . . . . . . . . . . . . . . . . . 182References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 188

15 Radiation and Retarded Potentials . . . . . . . . . . . . . . . . . . . . . . . . . 19115.1 Radiation Field . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19115.2 Retarded Time and Position of a Particle Moving

with Constant Velocity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19315.3 Liénard–Wiechert Potentials . . . . . . . . . . . . . . . . . . . . . . . . . . 19515.4 Radiation in the Ultra-Relativistic Limit . . . . . . . . . . . . . . . . . . 19615.5 Retarded Potentials for an Ensemble of Particles . . . . . . . . . . . . 197

16 Dipole Radiation and Scattering of Electromagnetic Waves . . . . . . 20116.1 Dipole Radiation of a Linear Oscillator . . . . . . . . . . . . . . . . . . 20116.2 Radiation Reaction Force . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20316.3 Thomson Scattering . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20516.4 Light Pressure . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20816.5 Inverse Compton Scattering . . . . . . . . . . . . . . . . . . . . . . . . . . . 209References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 212

Contents ix

Page 11: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

17 Transition and Diffraction Radiation . . . . . . . . . . . . . . . . . . . . . . . 21317.1 Transition Radiation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21317.2 Fourier Transformation of the Radiation Field

and the Radiated Power . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21617.3 Diffraction Radiation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 218Reference . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 220

18 Synchrotron Radiation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22118.1 Synchrotron Radiation Pulse in the Plane of the Orbit . . . . . . . . 22118.2 Radiation Spectrum in the Plane of the Orbit . . . . . . . . . . . . . . 22418.3 Synchrotron Radiation Out of the Orbit Plane . . . . . . . . . . . . . . 22518.4 Integral Characteristics of the Synchrotron Radiation . . . . . . . . 226

19 Undulator Radiation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23119.1 Undulators and Wigglers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23119.2 Undulator Radiation for K � 1 . . . . . . . . . . . . . . . . . . . . . . . . 23219.3 Effects of Finite Length of the Undulator . . . . . . . . . . . . . . . . . 23519.4 Wiggler Radiation for KJ1 . . . . . . . . . . . . . . . . . . . . . . . . . . 236

20 Formation Length of Radiation and Coherent Effects . . . . . . . . . . . 24120.1 Longitudinal Formation Length . . . . . . . . . . . . . . . . . . . . . . . . 24120.2 Transverse Formation Length . . . . . . . . . . . . . . . . . . . . . . . . . 24420.3 Coherent Radiation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 24520.4 Effect of the Transverse Size of the Beam . . . . . . . . . . . . . . . . 247

21 Topics in Laser-Driven Acceleration . . . . . . . . . . . . . . . . . . . . . . . . 25121.1 The Lawson–Woodward Theorem . . . . . . . . . . . . . . . . . . . . . . 25121.2 Laser Acceleration in Space with Material Boundaries . . . . . . . 25321.3 Inverse FEL Acceleration . . . . . . . . . . . . . . . . . . . . . . . . . . . . 255References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 258

22 Radiation Damping Effects . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 25922.1 Radiation Damping in Equations of Motion . . . . . . . . . . . . . . . 25922.2 Synchrotron Damping of Betatron Oscillations . . . . . . . . . . . . . 26122.3 Vlasov Equation and Robinson’s Theorem . . . . . . . . . . . . . . . . 265Reference . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 267

Appendix A: Maxwell’s Equations, Equations of Motion,and Energy Balance in an Electromagnetic Field . . . . . . . . 269

Appendix B: Lorentz Transformations and the RelativisticDoppler Effect . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 273

Index . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 279

x Contents

Page 12: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

Part IClassical Mechanics

Page 13: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

Chapter 1The Basic Formulation of Mechanics:Lagrangian and Hamiltonian Equationsof Motion

The Lagrangian and Hamiltonian formalisms are among the most powerful ways toanalyze dynamic systems. In this chapter we will introduce Lagrange’s equations ofmotion and discuss the transition from Lagrange’s to Hamilton’s equations. We writedown the Lagrangian and Hamiltonian for a charged particle in an electromagneticfield, and introduce the Poisson bracket.

For more general information about classical mechanics, the authors recommendtextbooks by Goldstein, Safko and Poole [1] and by Landau and Lifshiftz [2].

1.1 Lagrangian

There are many different types of dynamic systems. There can be imposed trajec-tories, or motion with constraints, or the basic F = ma equations of motion. Inaccelerator systems particles move freely, and while control systems such as feed-back are almost always used they have long time scales and are usually analyzed ona separate level. Thus, most accelerator systems involve applying forces to particlesthrough prescribed fields, with the possibility of interactions among particles beingimportant as well.

Systems with simple expressions for the acceleration can sometimes be solved fordirectly given the physical laws at work. For example, a simple harmonic oscillatorsatisfying the equation

x + ω20x = 0 (1.1)

has straightforward solutions of the form a cos(ω0t + φ), where a characterizesthe amplitude of the motion and φ is a phase term that describes the timing.

© Springer International Publishing AG, part of Springer Nature 2018G. Stupakov and G. Penn, Classical Mechanics and Electromagnetismin Accelerator Physics, Graduate Texts in Physics,https://doi.org/10.1007/978-3-319-90188-6_1

3

Page 14: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

4 1 The Basic Formulation of Mechanics: Lagrangian and Hamiltonian…

Even slight generalizations, however, can be trickier to solve for and involve at thevery least mathematical objects that are more esoteric than trigonometric functions.A common nonlinear example is the pendulum equation,

θ + ω20 sin θ = 0 , (1.2)

where for a physical pendulum ω20 = g/ l, with l being the length of the pendulum

and g being acceleration due to gravity. We can still solve the pendulum equationexactly if we use energy conservation. Multiplying Eq. (1.2) by θ gives

1

2

d

dtθ2 − ω2

0d

dtcos θ = 0 , (1.3)

from which it follows that the quantity

E = 1

2ω20

θ2 − cos θ = const (1.4)

is conserved. We call E the energy of the system; each orbit is characterized by itsown energy. For a given energy E we have

θ = ±ω0

√2(E + cos θ) , (1.5)

and one more integration of this equation gives an explicit formula for the trajectoryθ(t) (see details in Sect. 4.5).

How does one write equations of motion for more complicated mechanical sys-tems, like the double pendulum shown in Fig. 1.1? The equations we would like touse are not easy to relate to the constraints representing, in this case, a joint. TheLagrangian formalism allows for easy formulation of such systems.

The first step in the Lagrangian formulation consists of choosing generalizedcoordinates, q1, q2, . . . , qn , which uniquely define a snapshot or configuration ofthe system at a particular time (these coordinates are said to define the configu-ration space). The number n is the number of degrees of freedom of our system.Each mechanical system possesses a Lagrangian function (or Lagrangian for short),which depends on the coordinates q1, q2, . . . , qn , velocities q1, q2, . . . , qn (withqi = dqi/dt), and time t—for brevity, we will write the Lagrangian as L(qi , qi , t)instead of L(q1, q2, . . . , qn, q1, q2, . . . , qn, t).

The Lagrangian has the following property: the integral

S =∫ t2

t1

L(qi , qi , t)dt , (1.6)

called the action, reaches an extremum along the true trajectory of the system whenvaried with fixed end points. The illustration in Fig. 1.2 gives an indication of whatthis condition means. This property can be used directly to find trajectories of a

Page 15: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

1.1 Lagrangian 5

Fig. 1.1 A double pendulum

Fig. 1.2 Several trajectorieshave the same initialcoordinates q(1)

i at t1 and

final coordinates q(2)i at t2.

The integral of theLagrangian reaches anextremum along the truetrajectory

system by numerically minimizing the action S. It is however not very practical, inpart because the varied trajectory is specified by its initial, qi (t1), and final, qi (t2),positions. In applications we would prefer to specify a trajectory by its initial positionand velocity instead.

For mechanical systems, the Lagrangian is equal to the difference between thekinetic and the potential energies. The kinetic energy represents the energy fromthe particle motion alone, while the potential energy U defines a force acting onthe particle through F = −dU/dx . For example, for a single pendulum with theequation of motion given by Eq. (1.2), with the angle θ chosen as a generalizedcoordinate q, the Lagrangian is

L(θ, θ) = m

2l2θ2 + gml cos θ . (1.7)

The most convenient approach to the problem of obtaining equations of motion fora given Lagrangian is based on the variational calculus. By direct minimization ofthe action integral, requiring

δ

∫ t2

t1

L(qi , qi , t)dt = 0 , (1.8)

Page 16: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

6 1 The Basic Formulation of Mechanics: Lagrangian and Hamiltonian…

one can get equations of motion in the following form:

∂L

∂qi− d

dt

∂L

∂qi= 0 , i = 1, . . . , n . (1.9)

These are ordinary differential equations which are easier to solve than trying todirectly minimize S.

Let us derive Eq. (1.9). Assume that qi (t) is a true orbit and the values qi (t1)and qi (t2) are fixed. Let δqi (t) be a deviation from this orbit; it has the propertyδqi (t1) = δqi (t2) = 0. Compute the variation of the action:

δ

∫ t2

t1

L(qi , qi , t)dt

=∫ t2

t1

L(qi + δqi , qi + δqi , t)dt −∫ t2

t1

L(qi , qi , t)dt

=∫ t2

t1

n∑

i=1

(∂L

∂qiδqi + ∂L

∂qiδqi

)dt

=∫ t2

t1

n∑

i=1

(∂L

∂qi− d

dt

∂L

∂qi

)δqidt , (1.10)

where in the last step we integrated the term with δqi by parts. Since qi (t) is a trueorbit, the action reaches an extremum on it, and the variation of the action should beof second order, i.e., ∝ δq2

i . This means that the linear variation that we have foundabove vanishes for arbitrary δqi , hence Eq. (1.9) must be satisfied.

The Lagrangian for a given system is not unique. There exist many Lagrangiansfor the same physical system that lead to identical equations of motion. This is easyto see from Eq. (1.8): adding to L any function g(qi , t) that is a total time derivativeof an arbitrary function f (qi , t) of coordinate and time,

g(qi , t) = d f (qi , t)

dt≡ ∂ f

∂t+

n∑

i=1

qi∂ f

∂qi, (1.11)

does not change the Eq. (1.9).There are several advantages of using a Lagrangian as a starting point for the

formulation of the equations of motion. First, we have complete freedom to choosegeneralized coordinates qi . Second, the Lagrangian formalism is closely related topowerful variational principles, as indicated by Eq. (1.8). Finally, there is a connectionbetween the symmetries of the Lagrangian and the conservation laws for the system.A simple example of such a connection is given by the case when L does not dependon qi : as follows from Eq. (1.9), in this case the quantity ∂L/∂qi is conserved.

Page 17: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

1.2 Lagrangian of a Relativistic Particle in an Electromagnetic Field 7

1.2 Lagrangian of a Relativistic Particlein an Electromagnetic Field

For particle accelerators, the primary interest is in the motion of relativistic chargedparticles in an electromagnetic field. The Lagrangian for such a particle is formulatedin terms of the vector potential A and the scalar potential φ; the electric E and mag-netic B fields are given by E = −∇φ − ∂A/∂t and B = ∇ × A. The Lagrangianhas the following form:

L(r,υ, t) = −mc2√

1 − υ2/c2 + eυ · A(r, t) − eφ(r, t) (1.12)

= −mc2

γ+ eυ · A(r, t) − eφ(r, t) ,

where e is the electric charge of the particle, β = υ/c, and γ = (1 − β2)−1/2 is theLorentz factor. In a Cartesian coordinate system, r = (x, y, z), and the Lagrangian isgiven as a function L(x, y, z, x, y, z, t), where, of course, x = υx , y = υy , z = υz .

As an example of applying the Lagrangian formalism, let us study in detail themotion of a particle in a uniform magnetic field using the above Lagrangian. Weassume that the field is directed along the z-axis:

B = (0, 0, B0) , (1.13)

with the corresponding vector potential

A = (−B0y, 0, 0) . (1.14)

This gives for the Lagrangian,

L = −mc2√

1 − υ2/c2 − eB0υx y . (1.15)

Let us first consider the z direction of motion along the magnetic field. Because Ldoes not depend on z, from Eq. (1.9) we obtain d(∂L/∂vz)/dt = 0 and it followsthat

dγυz

dt= 0 . (1.16)

For the equation of motion in the x direction we have ∂L/∂x − d(∂L/∂υx )/dt = 0.Noting that L does not depend on x and that ∂L/∂υx = mc2γυx/c2 − eB0y, weobtain

−mdγυx

dt+ eB0υy = 0 . (1.17)

Page 18: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

8 1 The Basic Formulation of Mechanics: Lagrangian and Hamiltonian…

The Lagrange equation in the y direction similarly gives

−mdγυy

dt− eB0υx = 0 . (1.18)

These three equations can also be written in the familiar vectorial form

d pdt

= eυ × B , (1.19)

where the momentum p = mγυ. Because the magnetic force on the right-hand sideof this equation is perpendicular to the velocity, it does not produce work on theparticle, and the kinetic energy is conserved, γ = const. This can also be provenstraightforwardly from Eqs. (1.16)–(1.18): multiplying each of the equations by υz ,υx and υy , respectively, and adding them gives

υzdγυz

dt+ υy

dγυy

dt+ υx

dγυx

dt= 0 , (1.20)

which can also be written as υd(γυ)/dt = 0, with v =√

υ2x + υ2

y + υ2z . Because γ

is a function of υ, from this equation it follows that dγ/dt = 0. Knowing that γ isconstant, we can rewrite Eqs. (1.16)–(1.18) as

υz = 0, υx = ωHυy, υy = −ωHυx , (1.21)

where we have introduced the cyclotron frequency ωH ,

ωH = eB0

γm. (1.22)

Integrating Eq. (1.21) we find

υx = υ0 cos(ωH t + φ0), υy = −υ0 sin(ωH t + φ0) , (1.23)

where υ0 and φ0 are arbitrary transverse velocity and phase. With one more integra-tion we arrive at

x = υ0

ωHsin(ωH t + φ0) + x0 , y = υ0

ωHcos(ωH t + φ0) + y0 , (1.24)

with arbitrary x0 and y0. Equation (1.24) represents a circular orbit with the centerat x0, y0, and the radius

R = υ0

ωH= p

eB0. (1.25)

For a nonzero, constant υz the full 3-dimensional orbit would be a helix.

Page 19: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

1.3 From Lagrangian to Hamiltonian 9

1.3 From Lagrangian to Hamiltonian

The Hamiltonian approach has considerable advantages over the Lagrangian one.We will find that in the Hamiltonian approach it is simpler to change how a systemis characterized in order to suit our needs, and the quantities which come out of thisapproach have clearer physical meanings as well. These advantages are especiallyuseful in accelerator physics. A transition from the Lagrangian to the Hamiltoniandescription is made in three steps. First, we define the generalized momenta pi :

pi (qk, qk, t) ≡ ∂L(qk, qk, t)

∂qi, i = 1, . . . , n . (1.26)

Second, from the n equations pi = pi (qk, qk, t), i = 1, . . . , n we express all thevariables qi in terms of q1, q2, . . . , qn , p1, p2, . . . , pn and t ,

qi = qi (qk, pk, t) , i = 1, . . . , n . (1.27)

Finally, we construct the Hamiltonian function H as

H =n∑

i=1

pi qi − L(qk, qk, t) , (1.28)

and express all qi on the right-hand side through qi , pi and t using Eq. (1.27).This gives us the Hamiltonian as a function of variables qi , pi and t , namelyH(q1, q2, . . . , qn, p1, p2, . . . , pn, t). With the function H defined in this way, weclaim that the equations of motion of our system become:

pi = −∂H

∂qi, qi = ∂H

∂ pi. (1.29)

The partial derivatives ∂/∂qi are here understood to be the derivatives taken whileholding all pi variables constant, in contrast to the derivatives ∂/∂qi in Eq. (1.9)where the differentiation is carried out with constant qi . The variables pi and qi arecalled canonically conjugate pairs of variables.

Let us now prove Eq. (1.29) starting with the calculation of the partial derivative−∂H/∂qi ,

−∂H

∂qi= − ∂

∂qi

(n∑

k=1

pkqk − L

)

=n∑

k=1

(−pk

∂qk∂qi

+ ∂L

∂qk

∂qk∂qi

)+ ∂L

∂qi. (1.30)

Page 20: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

10 1 The Basic Formulation of Mechanics: Lagrangian and Hamiltonian…

Noting that, according to the definition Eq. (1.26), the momentum pk is equal to∂L/∂qk , we conclude that the sum in the parentheses on the right-hand side cancels,and we are left with ∂L/∂qi . According to the Lagrangian equations of motion ofEq. (1.9), we then have

−∂H

∂qi= ∂L

∂qi= d

dt

∂L

∂qi= dpi

dt. (1.31)

This proves the first of Eq. (1.29). The second one is proven analogously.As a simple exercise, let us derive the Hamiltonian for the pendulum. Starting from

the Lagrangian of Eq. (1.7) we use Eq. (1.26) to find the generalized momentum pcorresponding to the angular variable θ: p = ml2θ. The Hamiltonian is then obtaineddirectly from (1.28):

H(θ, p) = p2

2ml2− ω2

0ml2 cos θ , (1.32)

with ω0 = √g/ l.

1.4 Hamiltonian of a Charged Particlein an Electromagnetic Field

Starting from the Lagrangian (1.12),

L(r,υ, t) = −mc2√

1 − υ2/c2 + eυ · A(r, t) − eφ(r, t) ,

we first need to find the canonical conjugate momentum for which we will use thevectorial notation π = (πx ,πy,πz), combining the three conjugate variables to theCartesian coordinates r = (x, y, z):

π = ∂L

∂υ= m

υ√

1 − υ2/c2+ eA = mγυ + eA . (1.33)

Here the derivative with respect to the vector υ is understood as a collection of threederivatives, (∂/∂υx , ∂/∂υy, ∂/∂υz). Note that the conjugate momentum π differsfrom the kinetic momentum mγυ of the particle. Note also that as follows from theprevious equation, γβ = (π − eA)/mc, and hence

γ2β2 = (π − eA)2

m2c2. (1.34)

We will need this relation in what follows.

Page 21: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

1.4 Hamiltonian of a Charged Particle in an Electromagnetic Field 11

Now let us derive the Hamiltonian H = υ · π − L ,

H = υ · π + mc2√

1 − υ2/c2 − eυ · A + eφ

= mγυ2 + mc2

γ+ eφ

= mγc2 + eφ , (1.35)

where we have used Eq. (1.33) and the relation γ2 = 1 + γ2β2. Remarkably, theHamiltonian is the sum of the particle energy γmc2 and the potential energy associ-ated with the electrostatic potential φ. It seems that the vector potential A does notenter this expression. This conclusion, however, is misleading — the vector potentialis implicitly present in Eq. (1.35). To see this, remember that we need to express H interms of the conjugate coordinates r and momenta π. Again using γ2 = 1 + γ2β2,this time in combination with Eq. (1.34), we express γ2 through π and A:

γ2 = 1 + (π − eA)2

m2c2. (1.36)

Taking the square root of this expression and substituting it for γ in the Hamiltonianof Eq. (1.35) gives

H(r,π, t) =√

(mc2)2 + c2(π − eA(r, t))2 + eφ(r, t) . (1.37)

In this form, the Hamiltonian is expressed in terms of the proper conjugate variablesand explicitly contains both the scalar and vector potentials.

1.5 The Poisson Bracket

Let f (qi , pi , t) be a function of the conjugate coordinates, momenta and time, andtake any Hamiltonian trajectory qi (t) and pi (t). Then f becomes a function of timet only: f (qi (t), pi (t), t). What is the derivative of this function with respect to time?Using the chain rule for computing the derivative of the composition of functions wefind

d f

dt= ∂ f

∂t+

i

(∂ f

∂qiqi + ∂ f

∂ pipi

). (1.38)

Often d f/dt is referred to as the convective, or Lagrangian, derivative with respectto time. Substituting Eq. (1.29) into this equation gives

Page 22: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

12 1 The Basic Formulation of Mechanics: Lagrangian and Hamiltonian…

d f

dtconv= ∂ f

∂t+

i

(∂ f

∂qi

∂H

∂ pi− ∂ f

∂ pi

∂H

∂qi

)

= ∂ f

∂t+ { f, H}, (1.39)

where we have introduced the Poisson bracket of two functions f and g,

{ f, g} ≡∑

i

(∂ f

∂qi

∂g

∂ pi− ∂ f

∂ pi

∂g

∂qi

), (1.40)

and specifically indicated with the subscript the convective time derivative of f .Poisson brackets have many remarkable properties, which we will use in the next

chapter. They are anticommutative: for two functions f (qi , pi , t) and g(qi , pi , t),changing the order of f and g in the Poisson bracket flips the sign,

{g, f } = −{ f, g}. (1.41)

From this property it immediately follows that

{ f, f } = 0. (1.42)

If d f/dt = 0, then f is an integral of motion, meaning that its value remains constantalong the orbit. Note that if f does not explicitly depend on time, f = f (qi , pi ), itis an integral of motion if and only if { f, H} = 0, as follows from Eq. (1.39). Fromthis observation, we immediately conclude that a Hamiltonian that does not dependexplicitly on time is an integral of motion, because the Poisson bracket of H withitself is always equal to zero.

Considering each coordinate qk and momentum pk as a function that only dependson itself, we can easily apply to them the partial derivative operators∂/∂qi and∂/∂ pi .The result is: ∂qk/∂qi = ∂ pk/∂ pi = δik and ∂qk/∂ pi = ∂ pk/∂qi = 0. It is a simpleexercise to substitute these relations into Eq. (1.40) and to obtain:

{qi , qk} = {pi , pk} = 0, {qi , pk} = δik . (1.43)

Worked Examples

Problem 1.1 Consider a pendulum of length l and mass m, supported by a pivot thatoscillates in the vertical direction with frequency �, Y (t) = a sin(�t), where a isthe amplitude of oscillations. Obtain the Lagrangian and derive equations of motionfor the pendulum.

Solution: The pendulum position is given by x = L sin θ and y = Y − L cos θ.The potential energy is then V = mgy = mg[Y (t) − L cos θ], and the kinetic ener-gy is

Page 23: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

1.5 The Poisson Bracket 13

T = 1

2m(x2 + y2)

= 1

2m

[(L cos θ)2θ2 + (Y + θL sin θ)2]

= 1

2m

[L2θ2 + a2�2 cos2(�t) + 2θLa� cos(�t) sin θ

].

We then find the Lagrangian

L = 1

2m

[L2θ2 + a2�2 cos2(�t) + 2θLa� cos(�t) sin θ

]− mg [a sin(�t) − L cos θ] ,

and equations of motion from

∂L

∂θ= −mgL sin θ + θLma� cos(�t) cos θ,

d

dt

∂L

∂θ= 1

2m

[2L2θ − 2a�2 sin(�t)L sin θ + 2L θa� cos(�t) cos θ

].

Combining, we get

L2θ + L[g − a�2 sin(�t)] sin θ = 0,

so the gravitational term is modified by the acceleration, g → g − a�2 sin(�t).

Problem 1.2 Analyze a particle’s motion in a rotating frame using the Lagrangianapproach.

Solution: For a free particle, L = T so

L = 1

2mυ2 = 1

2m

(x2 + y2 + z2

).

For the rotating frame we introduce new coordinates x ′, y′, and z′ defined by

x = x ′ cos ωt − y′ sin ωt,

y = x ′ sin ωt + y′ cos ωt,

z = z′,

giving

x = x ′ cos ωt − x ′ω sin ωt − y′ sin ωt − y′ω cos ωt ,

y = x ′ sin ωt + x ′ω cos ωt + y′ cos ωt − y′ω sin ωt ,

z = z′ .

Page 24: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

14 1 The Basic Formulation of Mechanics: Lagrangian and Hamiltonian…

So we find

x2 + y2 = (x ′ cos ωt − x ′ω sin ωt − y′ sin ωt − y′ω cos ωt

)2

+ (x ′ sin ωt + x ′ω cos ωt + y′ cos ωt − y′ω sin ωt

)2

= x ′2 + y′2 + 2ω(yx ′ − x y′) + ω2(x ′2 + y′2) ,

and in the rotating frame we have the Lagrangian

L = 1

2m(x ′2 + y′2 + z′2) + mω(yx ′ − x y′) + 1

2mω2(x ′2 + y′2) .

In addition to the kinetic energy term we have a Coriolis force term (second term)and a repulsive potential (final term, often identified as the centrifugal force).

Problem 1.3 Derive equations of motion

d pdt

= eE + eυ × B ,

from the Lagrangian (1.12). Give the appropriate definition of the relativistic mo-mentum, p.

Solution: The Langrangian for a charged particle in an electromagnetic field is

L(r,υ, t) = −mc2√

1 − υ2/c2 + eυ · A(r, t) − eφ(r, t) .

To find the equations of motion we need the following quantities:

∂L

∂r= e∇ [υ · A(r, t)] − e∇ [φ(r, t)] ,

d

dt

∂L

∂υ= d

dt

(

mc2 υ

c2√

1 − υ2/c2+ eA(r, t)

)

= d

dt(γmυ) + e

(∂A∂t

+ (υ · ∇)A)

,

where we used d/dt = ∂/∂t + υ · ∇. Combining to find the equations of motion,we find

d

dt(γmυ) = e∇ (υ · A) − e∇φ − e

(∂A∂t

+ (υ · ∇)A)

=(

−e∇φ − e∂A∂t

)+ e [(υ · ∇)A + ∇ (υ · A)]

= eE + e (υ × (∇ × A))

= eE + eυ × B ,

Page 25: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

1.5 The Poisson Bracket 15

Fig. 1.3 The coordinatesystem ξ, s, z. The referencecircle, shown in blue, has aradius equal to R. Thecoordinate ξ of a particle isdefined as the differencebetween its polar radius rand the circle radius R

where we used the identity a × (∇ × b) = ∇(a · b) − (a · ∇)b. The mechanical mo-mentum is p = γmυ.

Problem 1.4 Write the Lagrangian of a relativistic charge e moving in a uniformmagnetic field B0 with axis z directed along the magnetic field. Use the coordinatesystem (ξ, s, z) shown in Fig. 1.3. Derive the equations of motion.

Solution: In the system of coordinates r = (ξ, s, z), the expression for the mag-netic field in terms of the vector potential A = (Aξ, As, Az) is

B = ∇ × A

=[

1

1 + ξ/R

∂Az

∂s− ∂As

∂z,

∂Aξ

∂z− ∂Az

∂ξ,

1

1 + ξ/R

(∂(1 + ξ/R)As

∂ξ− ∂Aξ

∂s

)].

We choose Aξ = −B0s (1 + ξ/R), with Az = As = 0; a direct calculation showsthat this vector potential corresponds to a uniform magnetic field B = (0, 0, B0). Wefind that the Lagrangian is

L = −mc2

√√√√1 − 1

c2

[

ξ2 +(

1 + ξ

R

)2

s2 + z2

]

− eξB0s

(1 + ξ

R

).

The derivatives of the Lagrangian are:

∂L

∂ξ= mγs2(1 + ξ/R)/R − eB0ξs/R ,

∂L

∂s= −eB0ξ(1 + ξ/R) ,

∂L

∂z= 0 ,

∂L

∂ξ= mγξ − eB0s(1 + ξ/R) ,

∂L

∂s= mγs (1 + ξ/R)2 ,

∂L

∂ z= mγ z ,

where we simplified using γ = 1/√

1 − v2/c2. We then use Eq. (1.9) to give theequations of motion:

mγξ = (1 + ξ/R)(eB0s + γms2/R

),

mγ(1 + ξ/R)2s = −2mγ(1 + ξ/R)ξs/R − eB0(1 + ξ/R)ξ ,

z = 0 .

Page 26: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

16 1 The Basic Formulation of Mechanics: Lagrangian and Hamiltonian…

Note one particular solution of these equations: ξ = const, s = −eB0R/γm, z =const. This solution corresponds to circular motion about the origin with constantorbit radius R + ξ. Taking into account that the velocity is |s|(1 + ξ/R) it is easy tosee that the radius of the orbit satisfies Eq. (1.25).

Problem 1.5 The magnetic field as given in Eq. (1.13), B = (0, 0, B0), can be rep-resented not only by the vector potential (1.14) but also by A = 1

2 (−B0y, B0x, 0).Show that the equations of motion are the same as for the vector potential (1.14).

Solution: With A = 12 (−B0y, B0x, 0), we have

L = −mc2√

1 − (x2 + y2 + z2)/c2 + eB0 (x y/2 − yx/2) .

We calculate:

∂L

∂x= −1

2B0ey ,

∂L

∂y= 1

2B0ex ,

∂L

∂z= 0 ,

∂L

∂ x= 1

2B0ey + mγ x ,

∂L

∂ y= −1

2B0ex + mγ y ,

∂L

∂ z= mγ z .

We then use Eq. (1.9) to give the equations of motion,

mγ x = −B0ey ,

mγ y = B0ex ,

z = 0 ,

which are equivalent to Eqs. (1.16)–(1.18).

Problem 1.6 Find conjugate momenta in cylindrical coordinates of a charged par-ticle moving in an electromagnetic field.

Solution: The Lagrangian is given by

L(r,υ, t) = −mc2√

1 − υ2/c2 + eυ · A(r, t) − eφ(r, t) .

For cylindrical coordinates ρ, φ, z, the velocity components are υρ = ρ, υφ = ρφ,υz = z, the total velocity is given by υ2 = z2 + ρ2 + ρ2φ2 and we find

L(r,υ, t) = −mc2

1 − z2 + ρ2 + ρ2φ2

c2+ e

(ρAρ + ρφAφ + z Az

) − eφ(r, t) .

Page 27: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

1.5 The Poisson Bracket 17

The conjugate momenta for the cylindrical coordinates ρ, φ, and z are found by

πρ = ∂L

∂ρ= mγρ + eAρ ,

πφ = ∂L

∂φ= mγρ2φ + eρAφ ,

πz = ∂L

∂ z= mγ z + eAz .

Problem 1.7 The angular momentum M of a particle is defined as M = r × p.Find the Poisson brackets {Mi , xk}, {Mi , pk} and {Mi , Mk}, where the indices i andk take the values x , y and z.

Solution: We write

M = r × p = (pz y − pyz, px z − pzx, pyx − px y

),

or more conveniently

Mi =3∑

a=1

3∑

b=1

εiabxa pb ,

where εiab is the Levi-Civita symbol equal to +1 (−1) when (i, a, b) is an even (odd)permutation of (1, 2, 3), and 0 if any index is repeated. We calculate the Poissonbracket (1.40) using

∂xa∂xi

= ∂ pa∂ pi

= δai ,

where δai is the Kronecker delta.To reduce the number of times needed to relabel indices, we will start by calcu-

lating the brackets of Mi with x j and p j :

{Mi , x j } = −∂Mi

∂ p j= −

3∑

a=1

εia j xa =3∑

a=1

εi ja xa ,

{Mi , p j } = ∂Mi

∂x j=

3∑

b=1

εi jb pb .

Page 28: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

18 1 The Basic Formulation of Mechanics: Lagrangian and Hamiltonian…

For {Mi , Mk} we have

{Mi , Mk} =3∑

j=1

(∂Mi

∂x j

∂Mk

∂ p j− ∂Mi

∂ p j

∂Mk

∂x j

)

=3∑

j=1

3∑

a=1

3∑

b=1

εi jaεk jb(xa pb − xb pa)

= xi pk − xk pi ,

using the identity3∑

j=1

εi jaεk jb = δikδab − δibδak .

The term including δab drops out because xa pb − xb pa must cancel whenevera = b. This result is equivalent to

{Mi , Mk} =3∑

j=1

εik j M j ,

in other words it is the angular momentum component along the cross product ofthe directions corresponding to i and k.

Problem 1.8 Simplify L and H in the nonrelativistic limit υ � c.

Solution: In the nonrelativistic limit, υ � c, we can simplify the Lagrangian

L(r,υ, t) = −mc2√

1 − υ2/c2 + eυ · A(r, t) − eφ(r, t)

≈ −mc2

(1 − υ2

2c2

)+ eυ · A(r, t) − eφ(r, t) ,

and the Hamiltonian

H = mγc2 + eφ

≈ mc2

(1 + (π − eA)2

2m2c2

)+ eφ ,

where we used γ = √1 + (π − eA)2/m2c2 ≈ 1 + (π − eA)2/2m2c2, for γ

close to 1.

Page 29: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

References 19

References

1. H. Goldstein, J. Safko, C. Poole, Classical Mechanics, 3rd edn. (Wiley, New York, 1998)2. L.D. Landau, E.M. Lifshitz, Mechanics, vol. 1, Course of Theoretical Physics (Elsevier

Butterworth-Heinemann, Burlington, 1976). (translated from Russian)

Page 30: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

Chapter 2Canonical Transformations

One of the benefits of the Lagrangian approach to mechanical systems is that wecan choose the generalized coordinates as we please. We have seen that once weselect a set of coordinates qi we can define the generalized momenta pi according toEq. (1.26) and form a Hamiltonian (1.28). We could also have chosen a different setof generalized coordinates Qi = Qi (qk, t), expressed the Lagrangian as a functionof Qi , used Eqs. (1.26) and (1.28), and obtained a different set of momenta Pi and adifferent Hamiltonian H ′(Qi , Pi , t). Although mathematically different, these tworepresentations are physically equivalent — they describe the same dynamics ofour physical system. Understanding the freedom that we have in the choice of theconjugate variables for aHamiltonian is important: a judicious choice of the variablescould allow us to simplify the description of the system dynamics.

The above procedure shows us that we can rewrite our equations of motion for anycoordinate system (even a moving one). But the Hamiltonian approach allows evenmore general changes in our choice of variables, as explained below. Let us assumethat we have a set of canonical variables qi , pi and the corresponding HamiltonianH(qi , pi , t), and then make a transformation to new variables

Qi = Qi (qk, pk, t) , Pi = Pi (qk, pk, t) , i = 1 . . . n. (2.1)

Can we find a new Hamiltonian H ′(Qi , Pi , t) such that the dynamics as expressedin the new variables is also Hamiltonian? What are the requirements on the transfor-mation (2.1) for such a Hamiltonian to exist?

These questions lead us to the notion of the canonical transformation.

© Springer International Publishing AG, part of Springer Nature 2018G. Stupakov and G. Penn, Classical Mechanics and Electromagnetismin Accelerator Physics, Graduate Texts in Physics,https://doi.org/10.1007/978-3-319-90188-6_2

21

Page 31: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

22 2 Canonical Transformations

2.1 Canonical Transformations

We first consider a time-independent Hamiltonian H(qi , pi ), and later generalize theresult for the case when H is a function of time. Instead of qi , pi we would like touse a new set of independent variables Qi with the transformation from the old tonew variables given by the following 2n equations,

Qi = Qi (qk, pk) , Pi = Pi (qk, pk) , i = 1 . . . n, (2.2)

see Fig. 2.1. An inverse transformation from Qi , Pi to qi , pi is written as

qi = qi (Qk, Pk) , pi = pi (Qk, Pk) , i = 1 . . . n. (2.3)

It is obtained by considering Eq. (2.2) as 2n equations for the old variables andsolving them for qi , pi .

Substituting (2.3) in H(qi , pi ), we can express our Hamiltonian in terms of thenew variables:

H ′(Qk, Pk) = H(qi (Qk, Pk), pi (Qk, Pk)), (2.4)

where we denote the new function by H ′. Let us assume that we have solved theHamiltonian equations of motion (1.29) and found a trajectory qi (t), pi (t). Thistrajectory is mapped, through the transformation (2.2), to an orbit in the phase spaceQi , Pi :

Qi (t) = Qi (qk(t), pk(t)) , Pi (t) = Pi (qk(t), pk(t)) . (2.5)

We would like the trajectory defined by the functions Qi (t) and Pi (t) to be a Hamil-tonian orbit, which means that it has to satisfy the equations

dPidt

= −∂H ′(Qk(t), Pk(t))

∂Qi,

dQi

dt= ∂H ′(Qk(t), Pk(t))

∂Pi. (2.6)

The right-hand side of these equations is understood as follows: we first take a partialderivative of H ′ with respect to Qi or Pi , holding all other Q and P variables constant,

Fig. 2.1 Transformationfrom the old variables qi , pito the new variables Qi , Pi :a point in the old phase spacemaps to a point in the newspace, and an old orbit istransformed into a new one

Page 32: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

2.1 Canonical Transformations 23

and then substitute for the arguments their values on the orbit Qk(t), Pk(t). If thetransformation (2.2) is such that Eq. (2.6) is satisfied for every Hamiltonian H , thenit is called a canonical transformation.

Two simple examples of the canonical transformations are easily established byinspection:

Qi = pi , Pi = −qi . (2.7)

Qi = −pi , Pi = qi . (2.8)

In both cases, the transformation trivially renames the coordinates to momenta, andvice versa, with the proper change of signs. This example clearly shows that theconjugate variables in the Hamiltonian dynamics play an equal role, in contrast tothe Lagrangian formalism where the coordinates qi are fundamentally different fromthe velocities qi .

2.2 Poisson Brackets and Canonical Transformations

We will now show how to test that a given transformation (2.2) is canonical. Thetest is based on the invariance of the Poisson bracket with respect to the canonicaltransformations.

Let us assume that we have two functions of canonical variables, f (qi , pi ) andg(qi , pi ), and calculate their Poisson bracket:

{ f, g}q,p =∑

i

(∂ f

∂qi

∂g

∂ pi− ∂ f

∂ pi

∂g

∂qi

)≡ J (q, p) , (2.9)

where, on the left-hand side, we now indicate the variables with respect to which thePoisson bracket is calculated. Using the inverse transformation (2.3) we can expressour functions in terms of the new variables Qi and Pi ; the resulting new functionsare denoted by f ′(Qi , Pi ) and g′(Qi , Pi ). Let us also calculate the Poisson bracketof the new functions with respect to the new variables:

{ f ′, g′}Q,P =∑

i

(∂ f ′

∂Qi

∂g′

∂Pi− ∂ f ′

∂Pi

∂g′

∂Qi

)≡ J ′(Q, P) . (2.10)

We will now show that if qi , pi → Qi , Pi is a canonical transformation, then rewrit-ing J ′(Q, P) by expressing the new variables in terms of the old ones gives J (q, p):

J ′(Qi (qk, pk), Pi (qk, pk)) = J (qk, pk) . (2.11)

To prove this statement we recall the relation between the Poisson bracket and thevariation of a function along a trajectory, Eq. (1.39). Because the canonicity property

Page 33: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

24 2 Canonical Transformations

is valid for all possible Hamiltonians, let us consider g(qi , pi ) as a Hamiltonian of afictitious system. Denote by qi (t), pi (t) an orbit that satisfies this Hamiltonian. Thenaccording to (1.39) the total time derivative d f/dt along such an orbit is given by

d f

dt= { f, g}q,p. (2.12)

Whenwe change to the new variables Qi and Pi in f , the same time dependence f (t)is now given by f (t) = f ′(Qi (t), Pi (t)), where Qi (t), Pi (t) is the orbit qi (t), pi (t)transformed into the newphase space. For the transformation to be canonical, the orbitQi (t), Pi (t)must satisfy Hamiltonian equations of motion with the newHamiltoniang′(Qi , Pi ) = g(qi (Qi , Pi ), pi (Qi , Pi )). Hence

d f

dt= { f ′, g′}Q,P (2.13)

and with (2.12) we conclude that

{ f, g}q,p = { f ′, g′}Q,P , (2.14)

which proves Eq. (2.11) for arbitrary functions f and g.1

In practice, to establish that a transformation is canonical, we do not need to verifythe equality (2.14) for all possible functions f and g; it is enough to make sure thatit holds for a special set of 3n2 pairs of functions.2 The first n2 pairs are obtained bychoosing two arbitrary indices i and k and setting f = Qi (ql, pl) and g = Qk(ql , pl).Using the fact that the Poisson brackets of the old coordinates are identically zero,Eq. (1.43), we conclude that the Poisson brackets of the new coordinates also haveto vanish,

{Qi , Qk}q,p = {qi , qk}q,p = 0 . (2.15)

The second n2 pairs are obtained by choosing f = Pi (ql , pl) and g = Pk(ql , pl) andrequiring

{Pi , Pk}q,p = {pi , pk}q,p = 0 . (2.16)

Finally, choosing f = Qi (ql , pl) and g = Pk(ql , pl) we obtain the third set ofequations,

{Qi , Pk}q,p = {qi , pk}q,p = δik . (2.17)

1Strictly speaking we have only proved Eq. (2.14) for the points in the phase space along oneparticular trajectory qi (t), pi (t). To establish Eq. (2.14) in the whole domain of the transformationqi , pi → Qi , Pi we need to consider a set of orbits in this domain with different initial conditions.2Actually, due to the symmetry of the Poisson bracket with respect to transpositions, the numberof independent, nontrivial equations is equal to 2n2 − n.

Page 34: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

2.2 Poisson Brackets and Canonical Transformations 25

To summarize, we have proven that if Eq. (2.2) represents a canonical transformation,then for any pair of indices i and k the new variables satisfy Eqs. (2.15)–(2.17). Theseare the necessary conditions for a transformation to be canonical. It turns out (butwe do not prove it here), that they also constitute a sufficient condition, that is if theyare satisfied for all pairs i and k, the transformation is canonical.

2.3 Generating Functions

Poisson brackets are useful for establishing that a given transformation is canonical.They do not, however, provide a tool with which one can create canonical transfor-mations. A technique that allows one to do that is based on the approach that usesso-called generating functions.

We will give a complete formulation of the method of generation functions in thenext section. Here, we consider a particular case of a time-independent generatingfunction of the so-called first type, F1. Such a generating function depends on 2nvariables: n old coordinates qi and n new coordinates Qi :

F1(qi , Qi ) . (2.18)

Having chosen an arbitrary function F1, one can generate a transformation of vari-ables (2.1) using the following set of equations:

pk = ∂F1(qi , Qi )

∂qk, Pk = −∂F1(qi , Qi )

∂Qk, k = 1 . . . n . (2.19)

The first n relations for pk represent equations for n unknown Qi . Solving theseequations one finds n functions Qi (qk, pk), i = 1 . . . n. Substituting these functionsto the right-hand side of the second n relations for Pk gives n functions Pk(qi , pi ) interms of the old variables. It turns out that the transformation of variables so obtainedis canonical.

While this statement is valid for arbitrary n, we will only prove it here for a simplecase of one degree of freedom, n = 1. In this case, we have two conjugate variables,q and p, and the canonical transformation (2.2) is defined by the following twoequations,

Q = Q(q, p) , P = P(q, p) . (2.20)

The generating function F1(q, Q) is a function of two variables, and Eq. (2.19) re-duces to

p = ∂F1(q, Q)

∂q, P = −∂F1(q, Q)

∂Q. (2.21)

Page 35: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

26 2 Canonical Transformations

We need to show that from Eq. (2.21) follow Eqs. (2.15)–(2.17). Since in one dimen-sion i = k = 1, Eqs. (2.15) and (2.16) are trivially satisfied, {Q, Q} = {P, P} = 0,and we only need to prove that

{Q, P}q,p = ∂Q

∂q

∂P

∂ p− ∂Q

∂ p

∂P

∂q= 1 . (2.22)

From the second of Eq. (2.21) we find

∂P

∂q= − ∂2F1

∂q∂Q− ∂2F1

∂Q2

∂Q

∂q,

∂P

∂ p= −∂2F1

∂Q2

∂Q

∂ p. (2.23)

Substituting these equations into the Poisson bracket in (2.22) we obtain

{Q, P}q,p = ∂Q

∂q

∂P

∂ p− ∂Q

∂ p

∂P

∂q= ∂Q

∂ p

∂2F1

∂q∂Q. (2.24)

The derivative ∂Q/∂ p in this expression can be found when we differentiate the firstof Eq. (2.21) with respect to p:

1 = ∂2F1

∂Q∂q

∂Q

∂ p. (2.25)

The right-hand sides of the previous two equations are identical, leading to the desiredrelation

{Q, P}q,p = 1 . (2.26)

2.4 Transformations with a Time Dependence and FourTypes of Generating Functions

Functions of the type (2.18) are not the only ones that generate canonical trans-formations. As it turns out, there are four different kinds of functions that can beused for this purpose. Moreover, all of these generating functions can be applied totime-dependent Hamiltonians, H(qi , pi , t), and produce time-dependent canonicaltransformations as well. In this section, we will consider this most general case thatincludes a time dependence.

A time-dependent canonical transformation adds the time variable to the relationsbetween the old and new variables:

Qi = Qi (qk, pk, t) , Pi = Pi (qk, pk, t) , i = 1 . . . n. (2.27)

Page 36: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

2.4 Transformations with a Time Dependence … 27

They map the orbits from the old phase space to the new one as was discussedearlier, and the question is whether the orbits in the new phase space satisfy Hamil-tonian equations. However, we will no longer insist on the particular relation (2.4)between the old and the newHamiltonians, and allow for a broader class of functionsH ′(Qk, Pk, t).

The first type of the generating functions, generalized to include a time depen-dence, takes the form F1(qi , Qi , t). The relations between the old and the new vari-ables are still given by Eq. (2.19), but the new Hamiltonian differs from the old oneby the time derivative ∂F1/∂t :

pi = ∂F1

∂qi, Pi = − ∂F1

∂Qi, (2.28a)

H ′ = H + ∂F1

∂t. (2.28b)

We remind the reader that the relations (2.28a) should be considered as a set ofequations through which one finds the transformation (2.27), and the old variables onthe right-hand side of (2.28b) are expressed through the new ones. As a consequence,while one can try any generating function to obtain a valid canonical transformation,it is not always straightforward to choose a generating function that yields a specificchange in variables.

The second type of generating function depends on the old coordinates and newmomenta, F2(qi , Pi , t). The equations for the new variables and the newHamiltonianare given by the following relations:

pi = ∂F2

∂qi, Qi = ∂F2

∂Pi, (2.29a)

H ′ = H + ∂F2

∂t. (2.29b)

The third type is defined by a function F3(pi , Qi , t):

qi = −∂F3

∂ pi, Pi = − ∂F3

∂Qi, (2.30a)

H ′ = H + ∂F3

∂t, (2.30b)

and the fourth type is generated by F4(pi , Pi , t):

qi = −∂F4

∂ pi, Qi = ∂F4

∂Pi, (2.31a)

H ′ = H + ∂F4

∂t. (2.31b)

Note that all of the generating functions depend on n old variables and n new ones.

Page 37: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

28 2 Canonical Transformations

Even for a time-dependent transformation the Poisson bracket can still be used fortesting if a given transformation is canonical; the variable t in this case is consideredas a parameter in calculations of partial derivatives.

Worked Examples

Problem 2.1 In a later chapter, we will have need of a transformation that is notcanonical. Show that the transformation Pi = λpi , Qi = qi , H ′ = λH , where λ isa constant parameter, preserves the Hamiltonian structure of equations.

Solution: The fact that the equations of motion in new variables are Hamiltonianis established by direct calculation:

dPidt

= dλpidt

= −λ∂H

∂qi= −∂H ′

∂Qi,

dQi

dt= dqi

dt= ∂H

∂ pi= ∂H ′

∂Pi.

Note that this transformation is not canonical, because

{Qi , Pk} = λ{qi , pk} = λδik .

Problem 2.2 Using the Poisson bracket, prove that the transformations Eqs. (2.7)and (2.8) are canonical.

Solution: We start with the transformations

Qi = pi , Pi = −qi .

Wecan then check {Qi , Qk} = {pi , pk} = 0, {Pi , Pk} = {−qi ,−qk} = {qi , qk} =0 and {Qi , Pk} = {pi ,−qk} = {qk, pi } = δik , so the transformation is canonical. Wecan easily see for the transformation

Qi = −pi , Pi = qi ,

that the same result follows.

Problem 2.3 Find generating functions for the transformations (2.7) and (2.8).

Solution: Again we start with the transformations

Qi = pi , Pi = −qi .

For generating functions of the first type,

pi = ∂F1

∂qi, Pi = − ∂F1

∂Qi,

Page 38: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

2.4 Transformations with a Time Dependence … 29

so plugging in for our transformation we find a generating function F1 = ∑i Qiqi .

We could also have used

qi = ∂F4

∂ pi, Qi = −∂F4

∂Pi,

to find a generating function of the fourth type, F4 = −∑i pi Pi . For the transfor-

mation

Qi = −pi , Pi = qi ,

we find generating functions F1 = −∑i Qiqi or F4 = ∑

i Pi pi .

Problem 2.4 Find the generating functions of the second and third type for theidentity transformation

Qi = qi , Pi = pi .

This problem illustrates the fact that the choice of the type of the generatingfunction is not unique.

Solution: The generating function of the second type for this transformation is

F2 =∑

i

qi Pi .

The generating function of the third type has to satisfy

qi = −∂F3

∂ pi, Pi = − ∂F3

∂Qi,

which is solved by F3 = −∑i pi Qi .

Problem 2.5 A change of coordinates which does not involve any dependence onthe conjugate momenta is called a point transformation:

Qi = fi (q1, q2, . . . , qn) i = 1 . . . n , (2.32)

(in general, there can be an explicit time dependence in the coordinate transformation,but for this problemwewill not consider that). Find a generating function for a generaltime-independent point transformation (2.32).

Solution: The problem is solved by a generating function of the second type,F2 = ∑

i fi (q1, q2, . . . , qn)Pi :

Page 39: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

30 2 Canonical Transformations

Qi = ∂F2(q, P)

∂Pi= fi (q1, q2, . . . , qn)

pi = ∂F2(q, P)

∂qi=

k

∂ fk∂qi

Pk =∑

k

∂Qk

∂qiPk .

As an example, take the transformation from Cartesian to polar coordinates:q1 = x, q2 = y → Q1 = r = √

x2 + y2, Q2 = θ = tan−1(y/x). We then have thegenerating function F2 = P1

√x2 + y2 + P2 tan−1(y/x).

Problem 2.6 For a simple harmonic oscillator with two degrees of freedom,

H = p2x/2m + p2y/2m + mω2x x

2/2 + mω2y y

2/2 ,

use a generating function of the second kind to change to variables X = y andY = −x . Find PX , PY , and the new Hamiltonian H ′.

Solution: The generating function of the second type is F2 = yPx − x Py . Thenthe new coordinates are given by

X = ∂F2

∂Px= y , Y = ∂F2

∂Py= −x ,

and we can trivially invert the expressions for the old momenta,

px = ∂F2

∂x= −Py , py = ∂F2

∂y= Px ,

to obtain Px = py and Py = −px .Because F2 has no explicit time dependence, the original Hamiltonian only has

to be rewritten in terms of the new variables,

H ′ = P2x /2m + P2

y /2m + mω2y X

2/2 + mω2xY

2/2 .

Page 40: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

Chapter 3Action-Angle Variables and Liouville’sTheorem

Oneof themost powerful uses of canonical transformations is to express the dynamicsin terms of action-angle variables. These are phase space coordinates which providea simple description of the Hamiltonian motion, and are widely used in particledynamics. A geometrical view of the Hamiltonian flow in phase space leads us to theformulation of Liouville’s theorem that is crucial for understanding the fundamentalproperties of large ensembles of beam particles in accelerators.

3.1 Canonical Transformation for a Linear Oscillator

Wewill now apply the general formalism of canonical transformations from the pre-vious chapter to the particular example of the harmonic oscillator. The Hamiltonianfor an oscillator with a unit mass is

H(x, p) = p2

2+ ω2

0x2

2, (3.1)

where x is the coordinate, p is the conjugate momentum, and ω0 is the oscillatorfrequency. Writing the equations of motion:

p = −∂H

∂x= ω2

0x , x = ∂H

∂ p= p , (3.2)

we easily find their solution,

x = a cos(ω0t + φ0) , p = −aω0 sin(ω0t + φ0) , (3.3)

where a is the amplitude and φ0 is the initial phase of the oscillation. Both a and φ0

are constant.

© Springer International Publishing AG, part of Springer Nature 2018G. Stupakov and G. Penn, Classical Mechanics and Electromagnetismin Accelerator Physics, Graduate Texts in Physics,https://doi.org/10.1007/978-3-319-90188-6_3

31

Page 41: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

32 3 Action-Angle Variables and Liouville’s Theorem

We would like to find a canonical transformation from the old variables, x, p, tothe new ones, φ, I , (where φ is the new coordinate and I is the new momentum),such that the transformation φ, I → x, p takes the form

x = A(I ) cosφ , p = −A(I )ω0 sin φ , (3.4)

where A(I ) is an as-yet unknown function of I . The advantage of the new variablesis clear: the new momentum I is a constant of motion (because the amplitude of theoscillations is constant), and the new coordinate evolves as a linear function of time,φ = ω0t + φ0.

To construct the canonical transformation (3.4)wewill use the generating functionF1(x,φ) of the first type. First, we express p in terms of the old (x) and new (φ)coordinates by eliminating A(I ) from (3.4),

p = −ω0x tan φ . (3.5)

We then integrate the equation (∂F1/∂x)φ = p = −ω0x tan φ to find

F1(x,φ) =∫

p dx = −ω0x2

2tan φ . (3.6)

The new momentum is obtained by differentiating F1 with respect to the new coor-dinate φ,

I = −∂F1

∂φ= ω0x2

2(1 + tan2 φ) = 1

2ω0

(ω20x

2 + p2), (3.7)

where we have used Eq. (3.5) to express tan φ in terms of x and p. The function A(I )is then found when we substitute Eq. (3.4) into (3.7):

A(I ) =√2I

ω0. (3.8)

Equation (3.7) defines the new momentum in terms of the old variables. The newcoordinate can be found from Eq. (3.5):

φ = − arctanp

ω0x. (3.9)

Because the canonical transformation does not depend on time, the newHamiltonianis equal to the old one expressed in new variables. Comparing Eq. (3.1) with (3.7)we find that

H ′ = ω0 I . (3.10)

Page 42: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

3.1 Canonical Transformation for a Linear Oscillator 33

We see that the new Hamiltonian does not depend on the new coordinate. The equa-tions of motion in new variables are:

I = −∂H ′

∂φ= 0 , φ = ∂H ′

∂ I= ω0 . (3.11)

They are easily integrated:

I = const , φ = ω0t + φ0 . (3.12)

Of course, this is the same dynamics as described by the original Eq. (3.3), but it issimpler because one of the coordinates, I , turns out to be an integral of motion andthe other one, φ, is a simple linear function of time.

The (I,φ) pair is called the action-angle coordinates for this particular case.They are especially useful for building perturbation theory for more complicatedsystems that in the lowest approximation reduce to a linear oscillator. In the nextsection, we will derive the action-angle variables for a more general one-dimensionalHamiltonian system.

3.2 Action-Angle Variables in 1D

With a little more effort, we can generalize the action-angle variables introduced inthe previous section for the harmonic oscillator to 1D periodic motion in an arbitrarybut constant potential well U (x). Assuming a unit mass, the Hamiltonian for thissystem is

H(x, p) = p2

2+U (x) . (3.13)

The shape of the potential function is sketched in Fig. 3.1a with several trajectories inthe phase space x, p shown in Fig. 3.1b. Each trajectory is defined by a constant valueof the Hamiltonian, H(x, p) = E , where E is the energy. Both x and p for a giventrajectory are periodic functions of time oscillating with the revolution frequency ωthat depends on the energy, ω(E). The energy dependence of the frequency can beeasily established using the relations dx/dt = p = √

2(E −U (x)) and observingthat half a period of the revolution is given by the following integral:

1

2T = πω−1 =

∫ x2

x1

dx ′

p(x ′)=

∫ x2

x1

dx ′√2(E −U (x ′))

, (3.14)

where T = 2π/ω is the period, and x1 and x2 are the turning points on the orbit (seeFig. 3.1b).

Page 43: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

34 3 Action-Angle Variables and Liouville’s Theorem

Fig. 3.1 Illustration of 1DHamiltonian: a the potentialenergy U (x) and b threeorbits with differentenergies E

(a) (b)

Let us try to make a canonical transformation to new variables choosing thenew momentum to be equal to the energy E ; the corresponding canonical conjugatecoordinate is denoted by Q. We will use the generating function of the second typeF2(x, E). Integrating the equation ∂F2/∂x = p = √

2(E −U (x)), we obtain

F2(x, E) =∫ x

dx ′√2(E −U (x ′)) . (3.15)

Of course, this integral cannot be taken analytically in general, but at least it can becomputed numerically for any given functionU (x). Since this is a time-independenttransformation, the new Hamiltonian H ′ is equal to the old one expressed in termsof the new variables:

H ′(Q, E) = H = E , (3.16)

with the equations of motion for the new variables

Q = ∂H ′

∂E= 1 , E = −∂H ′

∂Q= 0 . (3.17)

We see that the evolution of the variable Q is very simple,

Q = t + t0 . (3.18)

We can say that the conjugate variable to the energy is time.The variable Q is not themost convenient one because in one revolution it increas-

es by one period, T = 2π/ω(E), and this period changes with energy E . A betteroption would be to select a new coordinate, φ, in such a way that in one revolutionit changes exactly by 2π — the same quantity for each trajectory. This coordinate iscalled the angle, and the corresponding generalized momentum, J , is the action.

Aswewill see, the action is a function of energy, J (E), or, conversely, E = E(J );this function will be found below. To find J and φ for the system (3.13) we need to goback and replace our canonical transformation (3.15) by a slightly modified one. Thegenerating function F2(x, J ) that accomplishes the transformation (x, p) → (φ, J )is the same function F2, in which we now replace E by E(J ),

Page 44: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

3.2 Action-Angle Variables in 1D 35

F2(x, J ) = F2(x, E(J )) . (3.19)

With this arrangement, the new Hamiltonian H is

H(φ, J ) = E(J ) , (3.20)

and the equation for φ reads

φ = ∂ H

∂ J= dE

d J. (3.21)

We require that φ be equal to ω(E), so that

φ = ω(E)t + φ0 . (3.22)

With this time dependence, one orbital period corresponds to the change of variableφ by 2π, as desired. Combining (3.21) and (3.22) we obtain the differential equationfor E(J ),

dE

d J= ω(E) . (3.23)

Integrating this equation gives

J (E) =∫ E

Emin

dE ′

ω(E ′), (3.24)

where Emin is the energy corresponding to the bottom of the potential well U . Theinverse function of J (E) gives E(J ), which upon substitution into (3.19) in combi-nation with (3.15) fully defines the generating function F2 and finalizes the canonicaltransformation to φ, J .

The key features of action-angle coordinates are that the action is a constant of themotion, and the angle grows linearly in time, with periodic motion corresponding toa change in phase of 2π. The rate of change in the phase is generally different fordifferent trajectories. The simple harmonic oscillator is a notable exception whereall trajectories have the same period.

Strictly speaking, different Hamiltonians have different corresponding action-angle coordinates, but the resulting transformations will always be canonical. Thetransformation from a simple dynamical system can be useful as a first approximationto the action-angle coordinates in a more complicated system even when as a resultthe action is not exactly constant and the phase does not grow exactly linearly intime.

Page 45: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

36 3 Action-Angle Variables and Liouville’s Theorem

3.3 Hamiltonian Flow in Phase Space and Symplectic Maps

Wewill now take another look at theHamiltonianmotion, focusing on its geometricalaspect. Let us assume that for a givenHamiltonian H(qi , pi , t), for every set of initialconditions q0

i , p0i from some domain, we can solve the equations of motion starting

from initial time t0 and find the values of qi and pi at time t . This gives us a mapof the initial domain in the 2n dimensional phase space to a manifold in the samephase space at time t :

qi = qi (q0i , p

0i , t0, t) , pi = pi (q

0i , p

0i , t0, t) . (3.25)

Varying t in these equations moves each point (qi , pi ) along its trajectory and theset of all trajectories starting from the initial domain constitutes a Hamiltonian flow,as illustrated by Fig. 3.2.

In an accelerator context one can associate, for example, each trajectory (3.25)with a different particle in a beam.Assume that there is a beamdiagnostic at a locationin the ring that measures coordinates of particles when the beam passes by at timet0. On the next turn, at time t = t0 + T , where T is the revolution period in the ring,it measures coordinates again. The relation between the new and old coordinates isgiven by Eq. (3.25) that connects the initial and final coordinates in a Hamiltonianflow.

A remarkable feature of the functional relations (3.25) is that, for a given t0 and t ,they constitute a canonical transformation from q0

i , p0i to qi , pi , which is also called

a symplectic transfer map. While we are not going to prove the canonical propertiesof this map in the general case of n degrees of freedom, we will demonstrate it belowfor n = 1. In what follows, we drop the index i , using the notations q and p.

The proof is based on the calculation of the time derivatives of the Poisson brack-ets {q, p}q0,p0 , {p, p}q0,p0 and {q, q}q0,p0 and showing that they are equal to zero.Since at the initial time t = t0 the transformation from q0, p0 to q, p is the identitytransformation (q = q0, p = p0), it is clearly canonical. After we prove that the

Fig. 3.2 Hamiltonian flowin phase space. Orbitsstarting in the domain M0 attime t0 end up in the domainM1 at time t

Page 46: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

3.3 Hamiltonian Flow in Phase Space and Symplectic Maps 37

Poisson brackets do not change with time we will have verified that the map remainscanonical for all values of t .

Let us calculate the time derivative of {q, p}q0,p0 :

d

dt{q, p}q0,p0 = d

dt

(∂q

∂q0

∂ p

∂ p0− ∂q

∂ p0∂ p

∂q0

)(3.26)

= ∂ p

∂ p0∂

∂q0

dq

dt+ ∂q

∂q0

∂ p0dp

dt− ∂ p

∂q0

∂ p0dq

dt− ∂q

∂ p0∂

∂q0

dp

dt

= ∂ p

∂ p0∂

∂q0

∂H

∂ p− ∂q

∂q0

∂ p0∂H

∂q− ∂ p

∂q0

∂ p0∂H

∂ p+ ∂q

∂ p0∂

∂q0

∂H

∂q.

In the last transformation we have used the Hamiltonian equations of motion. Ap-plying the chain rules,

∂ p0= ∂ p

∂ p0∂

∂ p+ ∂q

∂ p0∂

∂q,

∂q0= ∂ p

∂q0

∂ p+ ∂q

∂q0

∂q, (3.27)

to the derivatives of the Hamiltonian, it is easy to show that all the terms on the right-hand side of (3.26) cancel and we obtain d{q, p}q0,p0/dt = 0. For one degree offreedom, the other two Poisson brackets {q, q} and {p, p} are automatically alwaysequal to 0 by the anticommutative property.

A somewhat different language of symplectic maps is often used in the literaturein connection with canonical transformations (2.2) or (3.25). A symplectic map isdefined with the help of the matrices J2n ,

J2n =

⎛⎜⎜⎜⎝

J2 0 0 00 J2 0 0

0 0. . . 0

0 0 0 J2

⎞⎟⎟⎟⎠ , (3.28)

where the right-hand side of this equation is a block n × n matrix with each elementtreated as 2 × 2 matrix. The diagonal elements are:

J2 =(0 −11 0

), (3.29)

and each zero in (3.28) is a 2 × 2 zero matrix. With the 2n × 2n matrix J2n definedin this way, one utilizes a uniform notation in which the conjugate coordinates andmomenta qi and pi are replaced by 2n variables wi : w2i−1 = qi , w2i = pi , i =1, 2, . . . , n. Similarly, instead of the new coordinates and momenta Qi and Pi oneuses Wi : W2i−1 = Qi , W2i = Pi , i = 1, 2, . . . , n. For example, for n = 2 we havew1 = q1, w2 = p1, w3 = q2, w4 = p2, and W1 = Q1, W2 = P1, W3 = Q2, W4 =P2. The transformation from the old to new variables (2.2) is then replaced by 2nfunctions

Page 47: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

38 3 Action-Angle Variables and Liouville’s Theorem

Wi = Wi (wk), i, k = 1, 2, . . . , 2n . (3.30)

It is a matter of a straightforward calculation to show that the requirement that allpossible Poisson brackets satisfy Eqs. (2.15)–(2.17) (which, aswe know, is equivalentto the requirement for the transformation to be canonical) can be concisely written as

MJ2nMT = J2n , (3.31)

where M is the Jacobian matrix of the transformation with the matrix elements

Mi j = ∂Wi

∂w j, (3.32)

and the superscript T denotes the transposition of a matrix.

3.4 Liouville’s Theorem

Let us consider a general canonical transformation (2.2) that maps a 2n-dimensionaldomainM0 in the phase space (qi , pi ) to a manifoldM1 in the phase space (Qi , Pi ).The volume V1 of M0 is given by the integral

V1 =∫M0

dq1dq2 . . . dqndp1dp2 . . . dpn , (3.33)

where the integration goes over themanifoldM0. ThemanifoldM1 has a volume V2,

V2 =∫M1

dQ1dQ2 . . . dQndP1dP2 . . . dPn . (3.34)

It turns out that the new volume is equal to the old one, V2 = V1. The proof followsfrom the following statement from calculus: the ratio of infinitesimal volumes in atransformation of variables is equal to the absolute value of the determinant of theJacobian matrix of the transformation M ,

dQ1dQ2 . . . dQndP1dP2 . . . dPndq1dq2 . . . dqndp1dp2 . . . dpn

= |det M | . (3.35)

At this point one can see the advantage of using the uniform notations of old, wi ,and new, Wi , variables introduced at the end of the previous section: the Jacobianmatrix is simply given by Eq. (3.32). For a canonical transformation, this matrixsatisfies the Eq. (3.31). Taking the determinant of the right-hand and left-hand sidesof (3.31) and noting that the determinant of a product of matrices is equal to theproduct of the determinants, with the observation that det J2n = 1, we conclude

Page 48: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

3.4 Liouville’s Theorem 39

that |det M | = 1. From the conservation of the infinitesimal volumes then followsthe conservation of integral volumes for arbitrary manifolds that are mapped by acanonical transformation.

Applied to the case of the Hamiltonian flow discussed in the previous section,Liouville’s theorem guarantees that the phase space volume occupied initially by abeam remains the same through its Hamiltonian evolution with time. In particular, itforbids increasing the phase space density of a beam by applying to its componentparticles electromagnetic fields that do not destroy the Hamiltonian nature of theirmotion. This fact has a fundamental importance for beam properties in accelerators.

3.5 Non-conservative Forces in Hamiltonian Dynamics

The dynamics of some physical systems do not fit into the framework of the Hamilto-nian and Lagrangian formalisms that we have been focusing on. In particular, damp-ing and arbitrary externally-applied forces lead to equations of motion that do notquite match those which we have looked at previously. We can consider such termsto be corrections to the equations of motion. This will be most useful when the non-conservative forces are small perturbations on time scales of a single oscillation.

The simplest place to start is with Eq. (1.9) from Chap.1, and generalize it to:

d

dt

∂L

∂qi− ∂L

∂qi= Fi , i = 1, . . . , n , (3.36)

where Fi = Fi (qk, qk) is a generalized force (for example, friction). Sometimes thisforce can be defined in terms of a potential-like term R(qk, qk), called the Rayleighdissipation function, as Fi = −∂R/∂qi . Although R does not represent a true po-tential or relate to any conserved quantity, it is convenient because, in contrast toFi , it does not change under coordinate transformations. Therefore, when changingcoordinates from qi to Qi , it is sufficient to replace R(qk, qk) with

R(Qi , Qi ) = R

⎛⎝qk(Qi ),

n∑j=1

∂qk∂Q j

Q j

⎞⎠ ,

where the functions qk(Qi ) express the old coordinates in terms of the new ones. Inother words, R preserves the covariance of the Lagrangian equations, even though itbreaks some of the conservation rules [1].

The switch to the Hamiltonian formalism uses the exact same definitions as forthe case where F (and R) vanish, with conjugate momentum pi = ∂L/∂qi and

H =n∑

i=1

pi qi − L(qk, qk, t) . (3.37)

Page 49: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

40 3 Action-Angle Variables and Liouville’s Theorem

Repeating the derivation from Sect. 1.3 yields

dpidt

= −∂H

∂qi+ Fi ,

dqidt

= ∂H

∂ pi. (3.38)

In these equations Fi is now understood as a function of the Hamiltonian variablesqk and pk which is obtained by expressing qk in the arguments of Fi through thesevariables, seeEq. (1.27). The formof the Fi maychangedramatically after a canonicaltransformation. Finally, when we calculate the total time derivative of a generalfunction f (qi , pi , t) in Hamiltonian coordinates, which in this context is called theconvective derivative, we find that

d f

dt conv= ∂ f

∂t+ { f, H} +

∑i

Fi∂ f

∂ pi. (3.39)

This equation generalizes Eq. (1.39) for the case of non-conservative forces Fi . Inparticular, the Hamiltonian evolves as

dH

dt= ∂H

∂t+

∑i

Fi∂H

∂ pi= ∂H

∂t+

∑i

Fi qi = ∂H

∂t−

∑i

qi∂R

∂qi, (3.40)

where the last expression is for a frictional force corresponding to a potential R, butthe derivatives of this potential and the qi terms should be viewed as functions of theqi and pi coordinates. Here, it is more obvious that the numerical value of R doesnot have to change for a simple change in coordinates.

We can also use this formalism for one degree of freedom to calculate the impactof frictional forces on the dynamic flow maps defined by (q0, p0) → (q(t), p(t)). Ifwe simplify Eq. (3.26) as far as possible without using the Hamiltonian equations ofmotion, instead only using the definition of the Poisson bracket, we find that for onedegree of freedom

d

dt{q, p}q0,p0 = {q, p}q0,p0

(∂q

∂q+ ∂ p

∂ p

). (3.41)

Using the expressions from Eq. (3.38), we see that the partial derivatives of H cancelin the last bracket, leaving only the correction ∂F/∂ p. We can integrate this equationto find

{q, p}q0,p0 = exp

[∫ t

t0

(∂F

∂ p

)dt ′

], (3.42)

Page 50: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

3.5 Non-conservative Forces in Hamiltonian Dynamics 41

where the integration goes along the trajectory that starts at time t0 with the initialconditions q0, p0, andwe used the fact that at the initial coordinates {q0, p0}q0,p0 = 1.This directly gives us the extent to which phase space volumes are changing, insteadof remaining constant as in Liouville’s theorem.

For multiple degrees of freedom, the generalization of the above is more complexand no simple expression can be derived for the evolution of the Poisson brackets{qi , pk} in the system governed by Eq. (3.36). However, if we focus on Liouville’stheorem, we only need to consider a single scalar, det M . We will show that a simpleexpression can be obtained for the time derivative of det M . In our derivation we willuse Jacobi’s formula [2] for the time derivative of the determinant of an invertiblematrix M(t):

d

dtdet M = det (M) tr

(M−1 dM

dt

), (3.43)

where the trace operator, tr, denotes the sum of the diagonal elements.To simplify the calculation we first look at the time derivative at t = t0. In this

case, the starting matrix M(t = t0) is the Jacobian of the identity map and hence isa unit 2n × 2n matrix, M(t = t0) = I2n . Substituting this matrix to the right-handside of Eq. (3.43) we find

d

dtdet M

∣∣∣∣t=t0

= tr

(dM

dt

)∣∣∣∣t=t0

=∑i

(∂qi∂qi

+ ∂ pi∂ pi

)=

∑i

(∂Fi∂ pi

)q0,p0

,

(3.44)

where we used the definition (3.32) to calculate the diagonal elements of M , andEq. (3.38) for non-conservative forces. To get the derivative at any time t we separateM into two matrices, one for a fixed map from t = t0 to t1 < t , M(t0 → t1), andanother mapping from the t1 coordinates to t , which we denote by M(t1 → t). Incalculation of the time derivative of det M we will use the fact that the determinantof a product of two matrices is equal to the product of the determinants,

d

dtdet M = det M(t0 → t1)

d

dtdet M(t1 → t) . (3.45)

In the limit when t1 approaches t , for the time derivative on the right-hand side ofthis expression we can use the result Eq. (3.44) in which t0 is replaced by t1. Theresult is:

d

dtdet M = det M(t)

∑i

(∂Fi∂ pi

)q(t),p(t)

. (3.46)

Although it is not necessary to know all the details of the map corresponding to thedynamic flow, the right-hand side in this equation does have to be evaluated alongparticle trajectories.

Page 51: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

42 3 Action-Angle Variables and Liouville’s Theorem

Because we are only extracting a scalar, it is straightforward to integrate thisequation:

det M(t) = exp

[∫ t

t0

∑i

(∂Fi∂ pi

)q(t ′),p(t ′)

dt ′]

=∏i

exp

[∫ t

t0

(∂Fi∂ pi

)q(t ′),p(t ′)

dt ′]. (3.47)

In some very simple examples for one degree of freedom the damping forcereduces to F = −γ p for constant γ, and we find that det M = e−γ(t−t0).

Worked Examples

Problem 3.1 Find the action-angle variables for the system with the following po-tential

U (x) ={ ∞, x < 0Fx, x > 0

.

Solution: The Hamiltonian H = (1/2)p2 +U (x) defines the energy E , and inthis case excludes particles from occupying x < 0 so particles “bounce” at x = 0.The period is given by

πω−1 =∫ x2

x1

dx ′√2(E −U (x ′))

=∫ E/F

0

dx ′√2(E − Fx ′)

= 1

F

√2(E − Fx ′)

∣∣∣E/Fx ′=0

= 1

F

√2E ,

with limits 0 and E/F chosen at the turning points. So we find

ω = Fπ√2E

,

and then

I (E) =∫ E

Emin

dE ′

ω(E ′)= 1

πF

∫ E

0

√2E ′dE ′

= 2√2

3πFE ′3/2

∣∣∣EE ′=0

= 2√2

3πFE3/2

= 2√2

3πF

(p2

2+ Fx

)3/2

.

Page 52: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

3.5 Non-conservative Forces in Hamiltonian Dynamics 43

We can then find the generating function F2(x, E) corresponding to the newmomentum equal to the energy E ,

F2(x, E) =∫ x

0dx ′√2(E − Fx ′) = − 1

3F

([2(E − Fx)]3/2 − (2E)3/2

].

The generating function for the action I is F2(x, I ) = F2(x, E(I )). Inverting ourequation for I (E) we find

E = 1

2(3πF I )2/3 ,

so

F2(x, I ) = −1

3

(1

F

[(3πF I )2/3 − 2Fx

]3/2 − 3π I

).

We then find the angle variable from

φ = ∂F2(x, E)

∂ I= −1

3

(1

F

[(3πF I )2/3 − 2Fx

]1/2 3

2(3πF)2/3

2

3I−1/3 − 3π

)

= − 1

3F

([(3πF I )2/3 − 2Fx

]1/2(3πF) (3πF I )−1/3

)+ π

= π − π p√p2 + 2Fx

,

using the identity (3πF I )2/3 = 2E = p2 + 2Fx . The π phase offset comes from thestarting point of the integral used to calculate F2(x, E). Starting at x = xmax insteadof x = 0 would have removed this phase term.

Problem 3.2 Derive Eq. (3.31) for n = 2.

Solution: For n = 2, we have the matrix J4 defined in terms of 2 × 2 submatrixcomponents as

J2n =(J2 00 J2

),

where J2 is given by Eq. (3.29). It is convenient to break down the matrix M in thesame way, as

M =(A11 A12

A21 A22

),

Page 53: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

44 3 Action-Angle Variables and Liouville’s Theorem

where

A11 =(

∂Q1/∂q1 ∂Q1/∂ p1∂P1/∂q1 ∂P1/∂ p1

), A12 =

(∂Q1/∂q2 ∂Q1/∂ p2∂P1/∂q2 ∂P1/∂ p2

),

A21 =(

∂Q2/∂q1 ∂Q2/∂ p1∂P2/∂q1 ∂P2/∂ p1

), A22 =

(∂Q2/∂q2 ∂Q2/∂ p2∂P2/∂q2 ∂P2/∂ p2

).

This readily generalizes to more degrees of freedom. While we do not knowmuch about any single term Mi j , we have constraints related to combinations thatcorrespond to Poisson brackets, becausewe are looking at canonical transformations.

We evaluate

MJ4MT =

(A11 J2AT

11 + A12 J2AT12 A11 J2AT

21 + A12 J2AT22

A21 J2AT11 + A22 J2AT

12 A21 J2AT21 + A22 J2AT

22

).

Now we can consider these 2 × 2 terms individually. We will only calculate theleft column, the quantities in the right column can be found in the same way just byswitching 1 ←→ 2 in the partial derivative terms.

We begin by looking at

A11 J2AT11 =

(M11 M12

M21 M22

)(0 −11 0

)(M11 M21

M12 M22

)

=(−M11M12 + M12M11 −M11M22 + M12M21

−M21M12 + M11M22 −M21M22 + M22M21

)

= (M11M22 − M12M21)

(0 −11 0

)= (det A11)J2 ,

where det A11 is the determinant of that 2 × 2 matrix. Repeating this calculationfor A12 J2AT

12 gives A12 J2AT12 = (det A12)J2 and the top left corner of the matrix

MJ4MT becomes

A11 J2AT11 + A12 J2A

T12 = (det A11 + det A12)J2 .

Writing the elements of A11 and A12 as partial derivatives yields

det A11 + det A12 = ∂Q1

∂q1

∂P1∂ p1

− ∂Q1

∂ p1

∂P1∂q1

+ ∂Q1

∂q2

∂P1∂ p2

− ∂Q1

∂ p2

∂P1∂q2

=∑i

(∂Q1

∂qi

∂P1∂ pi

− ∂Q1

∂ pi

∂P1∂qi

)≡ {Q1, P1}q,p = 1 .

This verifies that the top left corner is J2.

Page 54: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

3.5 Non-conservative Forces in Hamiltonian Dynamics 45

Now it only remains to show that the bottom left corner of MJ4MT is equal to 0.The individual terms in each off-diagonal submatrix are:

A21 J2AT11 =

(−M31M12 + M32M11 −M31M22 + M32M21

−M41M12 + M42M11 −M41M22 + M42M21

)

A22 J2AT12 =

(−M33M14 + M34M13 −M33M24 + M34M23

−M43M14 + M44M13 −M43M24 + M44M23

).

Expressing these quantities as partial derivatives, we see that all of the terms inA21 J2AT

11 give derivatives with respect to q1 and p1, while all the terms in A22 J2AT12

give derivatives with respect to q2 and p2. The 2 × 2 submatrix can be written as

A21 J2AT11 + A22 J2A

T12 =

( {Q1, Q2}q,p {P1, Q2}q,p{Q1, P2}q,p {P1, P2}q,p

).

All of these Poisson brackets must equal 0 for a canonical transformation, and theidentity

A21 J2AT11 + A22 J2A

T12 = 0

is established.

References

1. E. Minguzzi, Rayleigh’s dissipation function at work. Eur. J. Phys. 36, 035014 (2016)2. J.H.Hubbard, B.B.Hubbard,VectorCalculus, Linear Algebra, andDifferential Forms: AUnified

Approach, 2nd edn. (Prentice Hall, Upper Saddle River, 2001)

Page 55: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

Chapter 4Linear and Nonlinear Oscillators

The linear oscillator is a simple model that lies at the foundation of many physicalphenomena and plays a crucial role in accelerator dynamics. Many systems can beviewed as an approximation to a set of independent linear oscillators. In this chapter,we will review the main properties of the linear oscillator including its response toresonant excitations, slowly varying forces, random kicks, and parametric variationof the frequency. We will discuss the impact of damping terms as well as how small,nonlinear terms in the oscillator equation modify the oscillator frequency and leadto nonlinear resonance.

4.1 Harmonic Oscillator Without and with Damping

Wehave already encounteredEq. (1.1) for an ideal harmonic oscillatorwithout damp-ing. It has the general solution,

x(t) = a cos(ω0t + φ0) , (4.1)

where a is the amplitude and φ0 is the initial phase of oscillations.Damping due to a friction force that is proportional to the velocity x adds a term

with the first derivative into the differential equation,

x + γ x + ω20x = 0 , (4.2)

where γ is the damping constant and has the dimension of frequency. When thedamping is not too strong, γ < 2ω0, the general solution to this equation is

x(t) = ae−γt/2 cos(ω1t + φ0) , (4.3)

© Springer International Publishing AG, part of Springer Nature 2018G. Stupakov and G. Penn, Classical Mechanics and Electromagnetismin Accelerator Physics, Graduate Texts in Physics,https://doi.org/10.1007/978-3-319-90188-6_4

47

Page 56: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

48 4 Linear and Nonlinear Oscillators

with

ω1 = ω0

√1 − γ2

4ω20

. (4.4)

If γ � ω0, the frequency ω1 is close to ω0, ω1 ≈ ω0. The damping effect is oftenquantified by the so-called qualtity factor Q defined as Q = ω0/2γ; the regime ofweak damping is characterized by Q � 1.

If the oscillator is driven by an external force we have the following equationfor x(t):

x + γ x + ω20x = f (t) , (4.5)

where f (t) is the force divided by the oscillator mass. The general solution to thisequation can be found by the method of variation of parameters. Below we will needthe special case of this solution for γ = 0,

x(t) = x0 cosω0t + x0ω0

sinω0t + 1

ω0

∫ t

0sin[ω0(t − t ′)] f (t ′)dt ′ , (4.6)

where x0 and x0 are the initial values of the coordinate and velocity of the oscillatorat t = 0 (see a derivation in Problem4.1).

We can apply the formalism of the Rayleigh dissipation function from Sect. 3.5to the harmonic oscillator with friction. The Hamiltonian of an ideal oscillator (3.1)does not have an explicit dependence on time. For a linear oscillator the conjugatemomentum p = x , and when the friction force F = −γ x we find that R = γ p2/2and dH/dt = −γ p2. For a quadratic potential, we know that the average of p2 isH in the absence of damping. Thus at least for the weak damping case, γ � ω0, wecan expect to be able to take an average over a single oscillation (denoted below bythe angular brackets) to find the approximate relation

d

dt〈H〉 � −γ 〈H〉 , (4.7)

so the Hamiltonian decays as H � H0e−γt . Because the Hamiltonian scales as thesquare of the amplitude of motion, this is consistent with Eq. (4.3).

More general forms of the frictional force can also be considered.While the aboveexpression for H is not a replacement for the exact solution when the damping termis strong, it is useful in the case of weak damping, as will be seen in later chapters.

Page 57: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

4.2 Resonance in a Damped Oscillator 49

4.2 Resonance in a Damped Oscillator

Consider now an oscillator driven by a sinusoidal force with the frequency ω, f (t) =f0 cosωt . For this particular case, it is convenient to represent x(t) as the real partof a complex function ξ(t), x(t) = Re ξ(t). The differential equation for ξ(t),

ξ + γξ + ω20ξ = f0e

−iωt , (4.8)

is written in such a way that taking its real part gives Eq. (4.5) with f (t) = f0 cosωt .Due to the linearity of the problem, the real part of a solution to (4.8) is also a solutionto (4.5).

Let us seek a particular solution to (4.8) in the form ξ(t) = ξ0e−iωt where ξ0 is acomplex number, ξ0 = |ξ0|eiφ0 . For such ξ(t) we have x(t) = Re (|ξ0|e−iωt+iφ0) =|ξ0| cos(ωt − φ0), hence |ξ0| is the amplitude of the driven oscillation and φ0 is itsphase. Substituting ξ(t) into (4.8) and solving it for ξ0 we find

ξ0 = f0ω20 − ω2 − iωγ

. (4.9)

Taking the absolute value squared of ξ0 we obtain

|ξ0|2 = f 20(ω2

0 − ω2)2 + ω2γ2. (4.10)

When the damping factor γ is small, γ � ω0, the dependence of |ξ0|2 versus thefrequency ω exhibits resonant behavior: the amplitude of the oscillations increas-es when the driving frequency approaches the resonant frequency ω0, reaching itsmaximum atω ≈ ω0 with the resonant amplitude |ξ0| = f0/ω0γ. The resonant width�ωres is defined as a characteristic width of the resonant curve; a crude estimate for�ωres is �ωres ∼ γ.

A plot of the amplitude of the oscillations versus the frequency ω for severalvalues of the parameter γ is shown in Fig. 4.1.

4.3 Random Kicks

What happens to an oscillator if the external force is a sequence of random kicks?Let us assume that the external force is given by the following expression,

f (t) =∑i

fiδ(t − ti ) , (4.11)

Page 58: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

50 4 Linear and Nonlinear Oscillators

Fig. 4.1 Resonant curvesfor various values of thedamping parameter γ. Theratio γ/ω0 takes the valuesof 0, 0.1, 0.2 and 0.5 withlarger values correspondingto flatter curves. The curvefor γ = 0 tends to infinity atω = ω0

where ti are randommoments of time, and the kick amplitudes fi take random valueswith zero average value, 〈 fi 〉 = 0. The formal solution to this problem (for γ = 0)is given by Eq. (4.6):

x(t) = 1

ω0

∫ t

0sin[ω0(t − t ′)] f (t ′)dt ′ =

∑i

fiω0

sin[ω0(t − ti )] , (4.12)

where we have assumed that at time t = 0 the oscillator was at rest, x0 = x0 = 0.The result is a random process whose particular values are determined by the specificsequence of fi and ti . For any given realization of the random numbers fi and ti , thesolution (4.12)would exhibitwild fluctuations in amplitude and phase. Itmakes sensethen to consider the amplitude squared averaged over many random realizations ofthe random force with the same statistical properties. Note that for free oscillations,as it follows from (4.1), the square of the amplitude is given by x2 + ω−2

0 x2, so weassociate with the averaged amplitude squared the following quantity:

〈x(t)2 + ω−20 x2(t)〉 =

= ω−20

∑i, j

〈 fi f j{sin[ω0(t − ti )] sin[ω0(t − t j )] + cos[ω0(t − ti )] cos[ω0(t − t j )]

}〉= ω−2

0

∑i, j

〈 fi f j cos[ω0(ti − t j )]〉 , (4.13)

where the angular brackets denote the averaging. Let us assume that ti and t j arestatistically independent random numbers, and they are not correlated with the kickamplitudes fi . Then the averaging of fi f j on the last line of Eq. (4.13) can be splitfrom the averaging of the cosine functions, and using 〈 fi f j 〉 = f 2δi j we arrive atthe estimate

〈x(t)2 + ω−20 x2(t)〉 = f 2

ω20

N (t) , (4.14)

Page 59: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

4.3 Random Kicks 51

where N (t) is the average number of kicks in the interval [0, t]. The latter can beestimated as N (t) ≈ t/�t , where �t is the averaged time between the kicks. Wesee that the square of the oscillation amplitude grows linearly with time which is acharacteristic feature of the diffusion process. Hence the random uncorrelated kickslead to diffusion-like behavior of the oscillation amplitude with time.

Note that in the limitω0 → 0, which corresponds to a free particle, we can neglectthe term x(t)2 on the left-hand side to obtain

〈x2(t)〉 = f 2N (t) . (4.15)

This is a well known result for the velocity diffusion of a free particle caused byuncorrelated random kicks.

4.4 Parametric Resonance and Slow Variationof the Oscillator Parameters

Let us now consider what happens if the frequencyω0(t) of the linear oscillator varieswith time,

x + ω20(t)x = 0 , (4.16)

and moreover, ω0(t) is a periodic function of time. Unfortunately, there is no generalsolution to this equation for an arbitrary function ω0(t), so we will narrow our scopeand limit our analysis to the case where ω2

0(t) is given by the following formula,

ω20(t) = �2[1 − h cos(νt)]. (4.17)

Equation (4.16) with this time-dependent frequency is called the Mathieu equation;its solutions are given by the Mathieu functions.

Naively, one might think that if h is small, solutions will be close to those of theharmonic oscillator with the frequency approximately equal to �. This is true, how-ever, only on a limited time interval while for some combinations of the parameters h,� and ν, Eq. (4.16) may exhibit unstable solutions that grow exponentially with time.For small values of h, oscillations become unstable if the ratio of the frequencies�/νis close to n/2where n is an integer or, in other words, for ν ≈ 2�, �, 2

3�, 12�, . . ..

The exact pattern of stable and unstable regions in the plane of �, h parameters israther complicated; it is shown in Fig. 4.2. The unstable gaps between the stableregions become exponentially narrow when h � 1 and �/ν increases. This meansthat for a slow modulation, ν � �, the region h � 1 can be considered as a practi-cally stable area. This is the region of adiabatically slow variation of the oscillatorparameters.

Page 60: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

52 4 Linear and Nonlinear Oscillators

Fig. 4.2 Stability regions forthe Mathieu equation (4.16),(4.17) as functions ofamplitude modulation h. Thestable regions are shadowed

Remarkably, the adiabatic regime allows for a simple analytical treatment evenfor an arbitrary dependence ω0(t). An adiabatically slow variation means that

ω−20

∣∣∣∣dω0

dt

∣∣∣∣ � 1 , (4.18)

which also means that the relative change of the frequency ω0 over time ω−10 is small.

We seek a solution to Eq. (4.16) as a real part of the complex function ξ(t), with ξ(t)given by the following expression:

ξ(t) = A(t) exp

(−i

∫ t

0ω0(t

′)dt ′ + iφ0

), (4.19)

where A(t) is the slowly varying amplitude of the oscillations and φ0 is the initialphase. The integral over time in this expression properly accounts for the accumu-lation of the phase when ω0 is a function of time. Substituting this into Eq. (4.16)yields

A − 2iω0 A − i ω0A = 0 . (4.20)

Sincewe expect that the amplitude A is a slow function of time, we neglect the secondderivative A in this equation in comparison with the second term, which gives

2ω0 A + ω0A = 0 . (4.21)

This equation can also be written as

d

dtln(A2ω0) = 0 , (4.22)

Page 61: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

4.4 Parametric Resonance and Slow Variation of the Oscillator Parameters 53

Fig. 4.3 Illustration of the adiabatic invariance of A2ω0: the left panel shows the function ω0(t); onthe right panel the red curve shows the quantity x(t)2 + x2(t)/ω2

0(t) (which is close to the amplitudesquared, A2) while the blue curve shows the product of this quantity with ω0(t)

and then easily integrated: A(t)2ω0(t) = const. We found an adiabatic invariant forthe harmonic oscillator, from which it follows that, indeed, A(t) varies on the sametime scale as ω0(t), and hence is a slow-varying function as was assumed in thederivation. The value of the constant is defined by the initial values of A and ω0; ata later time the amplitude of the oscillation varies as A ∝ 1/

√ω0.

In Fig. 4.3 we show an example of numerical integration of Eq. (4.16) in anadiabatic regime that demonstrates an approximate conservation of the product A2ω0.It is a common feature of adiabatic invariants that once ω0(t) stops changing, theinvariant settles back to a value very close to its original value.

4.5 Nonlinear Oscillator and Nonlinear Resonance

The linear oscillator is only the lowest order approximation for a system in whichthe potential energy depends on the coordinate x in a fashion more general than thequadratic dependence ω2

0x2/2. Accounting for higher order terms in the potential

energy, Eq. (1.1) for x(t) is replaced by

x = −ω20x + αx2 + βx3 + · · · , (4.23)

where the terms on the right side of the equation are obtained through the Taylorexpansion of the potential energy close to the equilibrium position. The oscillator isweakly nonlinear if the nonlinear terms with α and β are small in comparison withthe linear ones. One of the most important properties of the nonlinear oscillator —in contrast to the linear one — is that its frequency depends on the amplitude.

Instead of trying to solve for the nonlinear oscillator (4.23) we will first analyzea particularly important example for particle accelerators, which is the pendulumequation (1.2),

θ + ω20 sin θ = 0 . (4.24)

Page 62: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

54 4 Linear and Nonlinear Oscillators

Fig. 4.4 Phase space orbitsfor the pendulum with thered curve showing theseparatrix

Note that for small amplitudes, |θ| � 1, we have

sin θ ≈ θ − 1

6θ3 , (4.25)

and we recover Eq. (4.23) with α = 0 and β = ω20/6. The linear approximation for

the pendulum equation is obtained if we neglect the cubic term in this expansion.As was outlined in Sect. 1.1, using the integral of motion that characterizes the

energy, E = 12ω

−20 θ2 − cos θ, we obtain the following equation for the time deriva-

tive θ:

θ = ±ω0

√2(E + cos θ) . (4.26)

This equation allows us to graph the phase portrait of the system where we plottrajectories in the plane (θ, θ/ω0) using contours of constant E , see Fig. 4.4. Thisplot shows the stable points at θ = 2πn, θ = 0 and unstable points at θ = 2πn + π,θ = 0, where n is an integer number. The trajectories that pass through the unstablepoints are called separatrices. Pendulum oscillations correspond to the values ofE such that −1 < E < 1; these trajectories exhibit bounded motion and occupya limited extension in θ. Orbits with E > 1 describe unbounded motion with thependulum rotating about the pivot point — the angle θ on these trajectories varieswithout limits. The separatrices are the orbits with energy E = 1.

We now calculate the period of the pendulum T (and hence the frequency ω =2π/T ) as a function of the oscillation amplitude. Inside the separatrix, for a givenenergy E , the pendulum swings between −θ0 and θ0, where θ0 is defined by theequation cos θ0 = −E . To find a half period of the oscillations we need to integratedt = dθ/θ from −θ0 to θ0:

1

2Tω0 = ω0

∫ θ0

−θ0

θ= 1√

2

∫ θ0

−θ0

dθ√cos θ − cos θ0

. (4.27)

This integral can be expressed in terms of the complete elliptic integral of the firstkind K ,1

1Here we use the definition of the complete elliptic integral following the convention of the software

package Mathematica [1], K (m) = ∫ π/20

(1 − m sin2 θ

)−1/2dθ.

Page 63: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

4.5 Nonlinear Oscillator and Nonlinear Resonance 55

Fig. 4.5 Period T for thependulum as a function ofthe amplitude angle θ0 in therange 0 < θ0 < π

T

T0= 2

πK

[sin2

(θ0

2

)]= 2

πK

(1 + E

2

), (4.28)

where T0 = 2π/ω0 is the period in the linear approximation. A plot of this functionis shown in Fig. 4.5.

For small values of the argument, the Taylor expansion of the elliptic functionis (2/π)K (x) ≈ 1 + x/4. This means that for small amplitudes the oscillation fre-quency is given by

ω ≈ ω0

(1 − θ20

16

). (4.29)

This frequency decreases with the amplitude θ0.Returning to the general case of the weakly nonlinear oscillator (4.23) and us-

ing our knowledge from the analysis of the pendulum in the limit of small ampli-tude (4.29), we can expect a correction to the frequency ω0,

ω(a) ≈ ω0 + νa2 , (4.30)

where a is the amplitude and ν is a constant. Indeed, as detailed calculations show [2],for given nonlinearity parameters α and β, the constant ν is given by the followingexpression:

ν = − 3β

8ω0− 5α2

12ω30

. (4.31)

The approximation (4.30) is valid if the nonlinear correction to the frequency is muchsmaller than ω0, ω0 � |ν|a2. Another interesting property of a weakly nonlinear os-cillation is anharmonicity— its Fourier spectrum contains not only the fundamental

Page 64: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

56 4 Linear and Nonlinear Oscillators

frequency ω(a), but also small contributions from higher harmonics nω(a), where nis an integer.

Nonlinearity also changes the resonance effect. We saw in Sect. 4.2 that for alinear oscillator an external force at the resonant frequency can drive the oscillatoramplitude to very large values, if the damping is small. The situation is different for anonlinear oscillator for a reason that is easy to understand:when the amplitude grows,the frequency of the oscillator drifts from its initial value, detuning the oscillator fromthe resonance. The unlimited growth of the amplitude ceases when the amplitudereaches some value a∗ which depends on the strength of the external force and thenonlinearity.

It is easy to get a rough estimate of the maximum value of a∗ using Eq. (4.10) forthe square of the amplitude in which we set γ = 0 assuming no damping. We replaceω0 by ω0 + νa2∗ and then set ω = ω0 (choosing the frequency of the driving force tobe equal to that of the linear oscillator). Using the smallness of νa2∗ we find

a2∗ ≈ f 20(2|ν|ω0a2∗)2

. (4.32)

This relation should be considered as an equation for a∗, from which we find

a∗ ≈(

f02|ν|ω0

)1/3

. (4.33)

We see that due to the nonlinearity, even at exact resonance the amplitude of theoscillations is finite; it is proportional to |ν|−1/3. For drive frequencies different fromω0, this dependence will change and the peak amplitude may even increase for aforcing term having the same magnitude.

Worked Examples

Problem 4.1 Prove that Eq. (4.6) gives a solution to Eq. (4.5) with γ = 0,

x + ω20x = f (t) . (4.34)

Verify that the initial conditions are satisfied. Generalize the solution Eq. (4.6) forthe case when γ �= 0.

Solution: Calculating the first and the second time derivatives of Eq. (4.6) we find

dx

dt= −x0ω0 sinω0t + x0 cosω0t +

∫ t

0cos [ω0(t − t ′)] f (t ′)dt ′ ,

Page 65: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

4.5 Nonlinear Oscillator and Nonlinear Resonance 57

and

d2x

dt2= f (t) − x0ω

20 cosω0t − x0ω0 sinω0t − ω0

∫ t

0sin [ω0(t − t ′)] f (t ′)dt ′ .

Substitution into Eq. (4.34) verifies the equation. To verify initial conditions, theevaluation of Eq. (4.6) at t = 0 yields

x(0) = x0 , x(0) = x0 . (4.35)

It is instructive to understand the origin of different terms in the solution (4.6).The first two terms, x0 cosω0t + x0ω

−10 sinω0t , is a solution of the homogeneous

equation (with f = 0) with the initial conditions (4.35). The third, integral termis a convolution of the force with the function ω−1

0 sin(ω0t). The latter is also asolution of Eq. (4.34) with f (t) = δ(t) and the initial conditions x(t < 0) = 0 andx(t < 0) = 0. This solution is called the Green function for the harmonic oscillator.Alternatively, ω−1

0 sin(ω0t) can be understood as a solution of the homogeneousequation (4.34) with the initial conditions x(0) = 0 and x(0) = 1.

The Green function approach can also be used for Eq. (4.5). Using the generalsolutions (4.3) of the homogeneous equation we first find a combination that satisfiesthe initial conditions (4.35):

e−γt/2

[x0 cosω1t +

(x0γ

2ω1+ x0

ω1

)sinω1t

]. (4.36)

We then find the Green function as a solution of the homogeneous Eq. (4.5) with theinitial conditions x(0) = 0 and x(0) = 1:

1

ω1e−γt/2 sin (ω1t) . (4.37)

Summing Eq. (4.36) with the convolution of (4.37) with f (t) gives

x(t) = e−γt/2

[x0 cosω1t +

(x0γ

2ω1+ x0

ω1

)sinω1t

]

+ 1

ω1

∫ t

0e−γ(t−t ′)/2 sin [ω1(t − t ′)] f (t ′)dt ′ .

Similar to what we did above for the solution (4.6) it is now straightforward to verifythat this solution satisfies Eq. (4.5).

Problem 4.2 The function f (t) is shown in Fig. 4.6: it is equal to zero for t < −�t ,and is constant for t > �t with a smooth transition in between. Describe the behaviorof the linear oscillator driven by this force in the limits �t � ω−1

0 and �t � ω−10 .

Page 66: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

58 4 Linear and Nonlinear Oscillators

Fig. 4.6 A sketch of thefunction f (t)

Solution: Assume γ = 0, and use the solution (4.5) with initial conditions x0 =x0 = 0. Over time �t , the function f (t) goes from 0 to f0. We have motion givenby

x = 1

ω0

∫ t

−�tsinω0(t − t ′) f (t ′)dt ′

= 1

ω0

[cosω0(t − t ′)

ω0f (t ′)

∣∣∣tt ′=−�t

−∫ t

−�t

cosω0(t − t ′)ω0

f (t ′)dt ′]

= 1

ω20

[f (t) −

∫ t

−�tcosω0(t − t ′) f (t ′)dt ′

].

For the case of �t � ω−10 , f (t) has a steep jump at t = 0, and can be approximated

by the Heaviside function, f (t) ≈ f0h(t), where h = 0 for t < 0 and h = 1 fort > 1. Then f ≈ f0δ(t) and for t > �t we find

x ≈ f0ω20

(1 − cosω0t) .

So a sudden kick from f leads to oscillations centered around f/ω20 (with f the force

per unit mass to give correct units). In the opposite limit, �t � ω−10 , we ignore f as

small, giving

x ≈ f (t)

ω20

.

In this adiabatic limit, the position changes smoothly from x = 0 to x = f0/ω20.

Problem 4.3 Assume γ = 0 in Eq. (4.8). Show that if ω � ω0 then one can neglectthe term ω2

0ξ in the equation. In other words, the oscillator responds to the drivingforce as a free particle. This fact explains why the dielectric response of many mediato x-rays can be computed neglecting the binding of electrons to nuclei.

Solution: For γ = 0 and ω � ω0 from Eq. (4.9) we find

Page 67: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

4.5 Nonlinear Oscillator and Nonlinear Resonance 59

ξ0 = f0ω20 − ω2

≈ − f0ω2

.

But this approximate solution, ξ(t) = −( f0/ω2)e−iωt , solves the original equa-tion (4.8) without the ω0 term (and γ = 0):

ξ = f0e−iωt .

This is an equation of motion for a free particle under the influence of the sinusoidalforce with the frequency ω.

Problem 4.4 Draw the phase portrait of a linear oscillatorwith andwithout damping.

Solution: The phase portraits are shown in Fig. 4.7. The derivative of the positionis scaled to ω0. The left image is for the special case of no damping, γ = 0, the centerimage is a typical example of weak damping, 0 < γ < 2ω0, and the final image isan example of strong damping, γ > 2ω0.

When there is no damping, the trajectories form periodic orbits, each correspond-ing to a fixed energy. With small damping (γ < 2ω0), the trajectories both oscillateand lose energy, and so spiral in to the center. The two trajectories shown both slowlyconverge to the center. When γ > 2ω0, ω1 is imaginary and we do not have oscilla-tory behavior, only rapid decay towards the center. Multiple trajectories are shown,all converging to this point.

Problem 4.5 Derive Eq. (4.29) directly from Eq. (4.27).

Solution: The period is given by

1

2Tω0 = 1√

2

∫ θ

−θ0

dθ√cos θ − cos θ0

.

For small angles we can use the Taylor expansion cos θ ≈ 1 − θ2/2! + θ4/4! andapproximate this as

Fig. 4.7 Linear oscillator phase portraits for γ = 0 (left image), γ = 0.3ω0 (center image), andγ = 2.1ω0 (right image)

Page 68: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

60 4 Linear and Nonlinear Oscillators

Fig. 4.8 Dependence of θversus time for a pendulumtrajectory

1

2Tω0 ≈ 1√

2

∫ θ0

−θ0

[(θ202! − θ40

4!)

−(

θ2

2! − θ4

4!)]−1/2

= 1√2

∫ θ0

−θ0

{(θ202

− θ2

2

) [1 − 1

6

(θ202

+ θ2

2

)]}−1/2

≈∫ θ0

−θ0

dθ[1 + (θ20 + θ2)/24

]√

θ20 − θ2

=∫ 1

−1

dx√1 − x2

[1 + θ20

24

(1 + x2

)],

with x ≡ θ/θ0. This integral can be found as

1

2Tω0 = sin−1 x

∣∣∣1−1+ θ20

24

(3

2sin−1 x − 1

2x√1 − x2

) ∣∣∣1−1

= π

(1 + θ20

16

).

So with T ≡ 2π/ω, we find

ω ≈ ω0

(1 − θ20

16

).

Problem 4.6 Figure. 4.8 shows a numerically computed function θ(t) for a pendu-lumwithω0 = 1. Try to figure out what is the energy E for this trajectory and explainqualitatively the shape of the curve.

Solution: Near the separatrix, that corresponds to θ0 = π and E = 1, the periodof oscillations becomes very large. The argument of the elliptic integral K from

Page 69: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

4.5 Nonlinear Oscillator and Nonlinear Resonance 61

Eq. (4.28) will also be close to unity. Denoting E = 1 − 2ε we can use an approxi-mation for the function K :

K (1 − ε) ≈ 1

2ln

(16

ε

),

valid in the limit ε � 1. This gives us an estimate for the period T near the separatrix:

T

T0≈ 1

πln

32

1 − E.

As we see, the period diverges logarithmically as E approaches its value at theseparatrix. The particle motion comes almost to a stop as it moves close to the ‘x’points of the separatrix,which leads to the long periods of nearly zero θ correspondingto almost reaching the highest possible position. Eventually, the particle falls backdown the way it came, undergoing a nearly full cycle until it approaches the highestposition from the other side.

Inverting the above expression yields

E = 1 − 32 exp

(−πT

T0

).

In the figure, the period of motion is about 60, while the low-amplitude period isT0 = 2π/ω0 = 2π because ω0 = 1. Therefore, the energy must be E � 1 − 32e−30;it takes 12 significant digits to see a difference between this value and 1.

Problem 4.7 Verify that Eq. (4.31) gives the result (4.29) for the pendulum.

Solution: For the pendulum, α = 0 and β = ω20/6, so

a = − 3β

8ω0− 5α2

12ω30

= −ω0

16,

and

ω(A) ≈ ω0 + aA2 = ω0

(1 − A2

16

),

where the amplitude A = θ0 is the maximum amplitude of the motion.

References

1. Wolfram Research, Inc. Mathematica, Version 11.2, 20172. L.D. Landau, E.M. Lifshitz, Course of Theoretical Physics, 3rd edn., Mechanics (Elsevier

Butterworth-Heinemann, Burlington MA, 1976). (translated from Russian)

Page 70: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

Chapter 5Coordinate System and Hamiltonianfor a Circular Accelerator

In this chapter, we derive the Hamiltonian for a particle moving in a circular accel-erator. Our derivation uses several simplifying assumptions. First, we assume thatthere is no electrostatic field, φ = 0, and the magnetic field does not vary with time.Second, the magnetic field is arranged in such a way that there is a closed reference(or nominal) orbit for a particle with a nominal momentum p0—this is achieved bya proper design of the magnetic lattice of the ring. We will also assume that thisreference orbit is a plane curve lying in the horizontal plane. Our goal is to describethe motion in the vicinity of this reference orbit of particles having energies (or,equivalently, momenta) that can slightly deviate from the nominal one.

There are very practical reasons for wanting a circular accelerator to at leastapproximate these properties. Accelerators can be very large and generally need tolie flat on the ground. Horizontal bending is obviously necessary to form the ring,while significant vertical bending will usually lead to coupling of the horizontal andvertical degrees of freedom unless extraordinary measures are taken in designing thebeamline.

More general consideration of the issues related to the derivation of the Hamil-tonian of a charged particle in an accelerator can be found in Refs. [1, 2]. TheHandbook of Accelerator Physics and Engineering [3] is an excellent reference textfor accelerators in general, with extensive references.

5.1 Coordinate System

A segment of the reference orbit is shown in Fig. 5.1. It is specified by the vectorfunction r0(s), where s is the arclength measured along the orbit in the directionof motion. In connection with this orbit, we will define three unit vectors. The firstvector s is tangential to the orbit, s = d r0/ds. The second vector, x, is perpendicularto s and lies in the plane of the orbit. The third vector, y, is perpendicular to the planeof the orbit, y = s × x. The three vectors x, y, and s constitute a right-hand oriented

© Springer International Publishing AG, part of Springer Nature 2018G. Stupakov and G. Penn, Classical Mechanics and Electromagnetismin Accelerator Physics, Graduate Texts in Physics,https://doi.org/10.1007/978-3-319-90188-6_5

63

Page 71: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

64 5 Coordinate System and Hamiltonian for a Circular Accelerator

Fig. 5.1 A segment of aplane reference orbit (thickblue curve), with the globalcoordinate system X , Y , Z ,and a local Cartesiancoordinate system x , y and s.For this orbit, the bendingradius ρ is positive. Aparticle is located at point P

base for the local coordinate system. The coordinate x is measured along x, and thecoordinate y is along y.

Note that the direction of the vector x in our definition is not uniquely specified:it can either be directed along, or against, the curvature vector of the orbit. It iscustomary to direct x away from the interior of the closed orbit, as shown in Fig. 5.1.Vector y, however, is unique after both s and x are established.

The three vectors s, x, and y, being defined locally, vary with s. Frenet-Serretformulas from differential geometry [4] establish the following relations betweentheir derivatives:

d sds

= − xρ(s)

, (5.1a)

d xds

= sρ(s)

, (5.1b)

d yds

= 0 , (5.1c)

where ρ(s) is the radius of curvature of the reference orbit. Since we have made thesimplifying assumption that there is no vertical bending, d x/ds is always parallelto the longitudinal coordinate s and the reference orbit lies in a plane. Therefore,the curvature sufficiently defines the local properties of the orbit. Furthermore, themagnetic field can only have y (vertical field) and/or s (solenoidal field) components.The absolute value of the radius ρ is given by the equation (see Eq. (1.25))

|ρ(s)| = p0|eBy(s)| , (5.2)

where p0 is the kinetic momentum of the reference particle and By is the verticalcomponent of the magnetic field. In the more general case of orbits that move out ofa single plane, the expressions for the derivatives above would have additional termsrelated to torsion, and the magnetic field could also have an x component, but thosewill not be considered here. Under the above constraints, the torsion is always zero.

Beam particles deviate from the reference orbit, but move close to it. A particleposition is defined in the local coordinate system as shown in Fig. 5.1, where a radiusvector r is represented by coordinates s, x , and y such that

Page 72: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

5.1 Coordinate System 65

r = r0(s) + x x(s) + y y . (5.3)

Below we will need to carry out various differential operations in curvilinear coordi-nates. Here are some useful formulae for the gradient of a scalar function φ(x, y, s),and for the curl and divergence of a vector function A = (Ax (x, y, s), Ay(x, y, s),As(x, y, s)):

∇φ = x∂φ

∂x+ y

∂φ

∂y+ s

1

1 + x/ρ

∂φ

∂s, (5.4a)

(∇ × A)x = − 1

1 + x/ρ

∂Ay

∂s+ ∂As

∂y, (5.4b)

(∇ × A)y = − 1

1 + x/ρ

∂As(1 + x/ρ)

∂x+ 1

1 + x/ρ

∂Ax

∂s, (5.4c)

(∇ × A)s = −∂Ax

∂y+ ∂Ay

∂x, (5.4d)

∇ · A = 1

1 + x/ρ

∂Ax (1 + x/ρ)

∂x+ ∂Ay

∂y+ 1

1 + x/ρ

∂As

∂s. (5.4e)

5.2 Hamiltonian in Curvilinear Coordinate System

The general Hamiltonian for a charged particle is given by Eq. (1.37), which forφ = 0 takes the form

H =√

(mc2)2 + c2(π − eA)2 . (5.5)

This Hamiltonian was derived for a Cartesian coordinate system. We now want touse it in the coordinate system related to the reference orbit described in the previoussection. This requires a canonical transformation from the old to new coordinatesthat we will carry out with the help of generating functions introduced in Sect. 2.4.

As a first step, we choose the local coordinates s, x , and y as coordinate variablesfor our new Hamiltonian. For the transformation from the original Cartesian coordi-nates X , Y , and Z (see Fig. 5.1) to the new ones, we use the generating function ofthe third type:

F3(π, x, y, s) = −π · (r0(s) + x x(s) + y y) . (5.6)

In this equation π is the old momentum and x , y and s are the new coordinates. Thenew canonical momentum is denoted by �; it is given by Eq. (2.30),

Page 73: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

66 5 Coordinate System and Hamiltonian for a Circular Accelerator

�x = −∂F3

∂x= π · x = πx ,

�y = −∂F3

∂y= π · y = πy ,

�s = −∂F3

∂s= π ·

(d r0ds

+ xd xds

)= π ·

(s + x

ρs)

= πs

(1 + x

ρ

), (5.7)

where we have used Eq. (5.1b). Note that

(π − eA)2 = (πx − eAx )2 + (πy − eAy)

2 + (πs − eAs)2

= (�x − eAx )2 + (�y − eAy)

2 +(

�s

1 + x/ρ− eAs

)2

, (5.8)

and our Hamiltonian can be written as

H = c

[

m2c2 + (�x − eAx )2 + (�y − eAy)

2 +(

�s

1 + x/ρ− eAs

)2]1/2

. (5.9)

Here we have introduced the notation Ax = A · x, Ay = A · y, and As = A · s1.Equation (5.9) is our new Hamiltonian as a function of the new coordinates x , y,

s and the new conjugate momenta �x , �y and �s .

5.3 Using s as Time Variable

It was assumed at the beginning of this chapter that our Hamiltonian does not dependon time t and hence is a constant of motion. Because a particle has three degreesof freedom, the Hamiltonian depends on three pairs of conjugate variables. It turnsout that, using conservation of H , one can transform the equations of motion insuch a way that they are described by a Hamiltonian with two pairs of conjugatevariables, effectively lowering the number of degrees of freedom from three to two.This is achieved by changing the independent variable from time t to the longitudinalcoordinate s.

Let us assume that we have solved the equations of motion and found all theHamiltonian variables as functions of time, x(t), y(t), s(t), etc. Then the form of,say, x(s) is obtained in the following way. From the equation s = s(t) we find theinverse function t (s) and substitute it into the argument of x : x(t) → x(t (s)). Thelatter becomes a function of s: x(s) = x(t (s)). Repeating the same procedure forthe coordinate y and the components of the momentum vector � we obtain y(s)and �(s).

1In some textbooks the reader can find a different definition of As , As = (1 + x/ρ)A · s.

Page 74: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

5.3 Using s as Time Variable 67

It turns out that the dependence of x , y,�x , and�y versus s can be found directlyby solving Hamiltonian equations with two degrees of freedom, without invoking theintermediate step of reversing the function s(t). We will first show how to constructthe new Hamiltonian, and then will prove that the new Hamiltonian equations areequivalent to the original ones.

We start from equating the Hamiltonian (5.9) to a constant value of h,

h = H(x,�x , y,�y, s,�s) , (5.10)

and then treat this relation as an equation for �s . Solving it, we find �s as a functionof the two pairs of conjugate variables plus h and s,

�s = �s(x,�x , y,�y, h, s) , (5.11)

where h is understood as a constant; recallingEq. (1.35) and remembering thatφ = 0,we see that the value of this constant is equal to γmc2. The crucial next step is tointroduce a new Hamiltonian, which we denote by K , equal to the negative function�s ,

K (x,�x , y,�y, h, s) = −�s(x,�x , y,�y, h, s) . (5.12)

Here x ,�x , y,�y are considered as canonical conjugate variables, s is an independent“time-like” variable, and h is a (constant) parameter. As we see, the Hamiltonian Khas two pairs of conjugate variables, and hence describes motion of a system whichhas two degrees of freedom. However, this Hamiltonian depends on s and hence, incontrast to the original H , is not a conserved quantity. The loss of the Hamiltonianconstancy is the price paid for lowering the number of degrees of freedom from threeto two.

Let us now prove that the s-evolution of the functions x(s), �x (s), y(s), and�y(s) are governed by the Hamiltonian (5.12). Because x(s) is obtained from x(t)and s(t) by eliminating the variable t , we have for dx/ds,

dx

ds= dx/dt

ds/dt= ∂H/∂�x

∂H/∂�s, (5.13)

where we have used the Hamiltonian equations of motion for dx/dt and ds/dt . Onthe other hand, the derivative ∂K/∂�x can be calculated as a derivative of an implicitfunction by differentiating Eq. (5.10) with respect to �x , yielding:

∂K

∂�x= −

(∂�s

∂�x

)

h

= ∂H/∂�x

∂H/∂�s. (5.14)

Page 75: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

68 5 Coordinate System and Hamiltonian for a Circular Accelerator

We see that

dx

ds= ∂K

∂�x, (5.15)

which is a Hamiltonian equation for dx/ds with the Hamiltonian K . The sameapproach works for �x ,

d�x

ds= d�x/dt

ds/dt= −∂H/∂x

∂H/∂�s=

(∂�s

∂x

)

h

= −∂K

∂x. (5.16)

Similarly one can calculate dy/ds and d�y/ds and find out that they are equal to∂K/∂�y and −∂K/∂y, respectively.

Although time is now eliminated from the equations, the time dependence ofs, and hence all of the variables, can be easily recovered. Indeed, in the originalHamiltonian equations we had ds/dt = ∂H/∂�s . Taking the reciprocal gives theequation for t (s):

dt

ds= 1

∂H/∂�s= ∂�s

∂h= −∂K

∂h. (5.17)

The second equality is found by substituting Eq. (5.11) into Eq. (5.10) and thendifferentiating with respect to h. Integrating this equation over s we find t = t (s),with the inverse function defining s(t).

5.4 Small Amplitude Approximation

We will now accept s as an independent variable and switch from the HamiltonianH (5.9) to the Hamiltonian K given by Eq. (5.12). Solving Eq. (5.9) for �s , weobtain the following expression for K :

K = −(1 + x

ρ

) [1

c2h2 − (�x − eAx )

2 − (�y − eAy)2 − m2c2

]1/2

− eAs

(1 + x

ρ

). (5.18)

As we will see in the next chapter, in most cases of interest a single component As

is sufficient to describe themagnetic field in an accelerator, sowe set Ax = Ay = 0 inEq. (5.18).With this choice of the vector potential,�x and�y are equal to the kineticmomenta, �x = px = mγvx and �y = py = mγvy (see Eqs. (5.7) and (1.33)), andwe can use px and py instead of �x and �y :

K = −(1 + x

ρ

) (1

c2h2 − p2x − p2y − m2c2

)1/2

− eAs

(1 + x

ρ

). (5.19)

Page 76: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

5.4 Small Amplitude Approximation 69

Particles in an accelerator beam typically move at small angles relative to the ref-erence orbit. This means that px and py are small in comparison with the totalmomentum p, and the square root in (5.19) can be expanded as a Taylor series. Thisis the main advantage of the local coordinate system, in which x and y are measuredrelative to the reference orbit. Expanding K in px and py and keeping only the lowestorder terms, we obtain:

K ≈ −p

(1 + x

ρ

) (

1 − p2x2p2

− p2y2p2

)

− eAs

(1 + x

ρ

), (5.20)

where p(h) = √h2/c2 − m2c2 is the total kinetic momentum of the particle (which

together with the energy is a conserved quantity in a constant magnetic field).Instead of using the variables px and py it is convenient to introduce dimensionless

quantities Px = px/p0 and Py = py/p0, where p0 is the nominal momentum in thering. The transformation from x , px , y, py to x , Px , y, Py is not canonical, but it doesnot change the Hamiltonian structure of the equations of motion. A simple analysisshows that the new Hamiltonian, which we denote by H, is obtained by dividing Kby p0 (see Problem 2.1 on p.33):

H(x, Px , y, Py) = K

p0

= − p

p0

(1 + x

ρ

) [

1 − 1

2P2x

(p0p

)2

− 1

2P2y

(p0p

)2]

− e

p0As

(1 + x

ρ

).

(5.21)

As mentioned at the beginning of this chapter, we are interested here in the casewhen the energy and the total momentum of the particles deviate only slightly fromthe nominal one, that is

p

p0= 1 + η, (5.22)

with η � 1. This somewhat simplifies H:

H(x, Px , y, Py)

≈ −(1 + η)

(1 + x

ρ

) (1 − 1

2P2x − 1

2P2y

)− e

p0As

(1 + x

ρ

), (5.23)

where we have replaced (p0/p)2 by unity in small quadratic terms proportional toP2x and P2

y .Finally, we note that our dimensionless momenta Px , Py are approximately equal

to the slopes x ′ ≡ dx/ds and y′ ≡ dy/ds, respectively. Indeed, we have

Page 77: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

70 5 Coordinate System and Hamiltonian for a Circular Accelerator

x ′ ≡ dx

ds= vx

vs= px

ps≈ Px , (5.24)

with a similar expression for y′. Some authors use the notations x ′ and y′ instead ofPx , Py for the canonical momenta conjugate to x and y —when using that notation,one has to be careful to avoid confusion between a canonical variable, e.g., Px , withthe rate of change of its conjugate, dx/ds.

5.5 Time-Dependent Hamiltonian

While we emphasized above that for a time-independent Hamiltonian the transitionfrom t to s as an independent variable eliminates one degree of freedom, the re-quirement of being independent of time is actually not needed. Without changes,our derivation in Sect. 5.3 can be easily generalized to the case when H is also afunction of t . Writing down Eq. (5.10) we now have H(x,�x , y,�y, s,�s, t) onthe right-hand side and, correspondingly, the newHamiltonian K becomes a functionof time,

K (x,�x , y,�y, t, h, s) = −�s(x,�x , y,�y, t, h, s). (5.25)

Time t in thisHamiltonian should nowbeunderstood as a third coordinate (in additionto x and y). Indeed, Eq. (5.17) is a Hamiltonian equation for t (s), if−h is associatedwith the momentum conjugate to t . For such an association to be valid, we also needthe complementary Hamiltonian equation,

dh

ds= ∂K

∂t. (5.26)

This equation is easily proven if one takes into account that dh/ds = (1/s)∂H/∂t ,expresses the time derivative of H in terms of the partial derivatives of K , ∂H/∂t =−(∂K/∂t)/(∂K/∂h), and uses Eq. (5.17). So we conclude that for a time-dependentHamiltonian H , in the transformation to the Hamiltonian K , one has to treat timeand the negative energy as a canonically conjugate pair of variables.

One important physical effect that requires a time-dependent Hamiltonian for itsdescription is acceleration of charged particles by radio-frequency (RF) electromag-netic fields. In a simple model which assumes that the field is localized in a short RFcavity which is powered to voltage V , an additional term that needs to be added tothe Hamiltonian (5.18) is [1]

eV

ωRFδ(s − s0) sin(ωRFt + φ) , (5.27)

Page 78: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

5.5 Time-Dependent Hamiltonian 71

whereωRF is theRF frequency,φ is the RF phase, and s0 is the coordinate of the cavitylocation in the ring. The cavity length in this equation is assumed to be infinitesimallysmall, so that its action is localized at one point in the ring. Using Eq. (5.26), it iseasy to calculate that a passage through the point s0 at time t changes the kineticenergy h of the particle by eV cos(ωRFt + φ), which reaches the maximum valueof �h = eV when the argument of the cos function is equal to a multiple of 2π.Because the Hamiltonian (5.23) is obtained from K through division by p0, whenadding the RF term to H we also need to divide it by the same factor p0.

Worked Examples

Problem 5.1 Check that Eq. (5.1) holds for a circular orbit.

Solution: A circular reference orbit of radius ρ0 is parametrized in Cartesiancoordinates by

r0 = (ρ0 cos(s/ρ0), ρ0 sin(s/ρ0), 0) ,

where we assumed that the orbit lies in the plane Z = 0. Using x = r0/ρ0 wefind that

d r0ds

= (− sin(s/ρ0), cos(s/ρ0), 0) = s ,

d sds

= d

ds(− sin(s/ρ0), cos(s/ρ0), 0)

=(

− 1

ρ0cos(s/ρ0),− 1

ρ0sin(s/ρ0), 0

)= − r0

ρ20= − x

ρ0,

d xds

= 1

ρ0

d r0ds

= sρ0

,

d yds

= d

ds(0, 0, 1) = (0, 0, 0) .

Problem 5.2 Verify that if we extend Eq. (5.2) to define a specific choice for thesign of ρ,

ρ(s) = p0qBy(s)

, (5.28)

this choice of sign is correct within Eq. (5.1) for arbitrary sign of the charge q andof the direction of the motion in the reference orbit.

Solution: Because s is directed along v we can write v = v s. Substituting thisvelocity into the magnetic force on the charged particle F = qv × B we obtain

Page 79: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

72 5 Coordinate System and Hamiltonian for a Circular Accelerator

Fig. 5.2 Electron trajectoryin a chicane with the colorrectangles showing thepositions of the four dipolemagnets. Assume that the yaxis is directed out of thepage

F = qvBy s × y = −qvBy x, where we have used the relation s × y = −x for aright-handed coordinate system x , y, s. Substituting Eq. (5.28) into Eq. (5.1a) wefind that d s/ds = −qBy x/p0, which means that d s/ds is co-linear with the forceF, and the trajectory turns in the direction of the force. Thus, this choice of signis consistent with the sign in Eqs. (5.1a, 5.1b, 5.1c) for either sign of q or By . Weemphasize that for this consistency to hold true it is necessary to align s, the directionof increasing s, with the velocity of the particle.

Problem 5.3 Figure5.2 shows the electron trajectory in a four-dipole chicane (typ-ically used for bunch compressions). Indicate the direction of axis x assuming thatthe y axis is directed out of the page toward you. Determine the sign of the orbitradius ρ and the magnetic field direction of each of four dipoles along the orbits.What happens with this sign if the particle is moving in the direction opposite to theone shown in the figure?

Solution: For the chicane, taking the y axis out of the plane and s to the right,the x axis is pointing up. Tracking the rotation of x through the chicane, it is easy tosee that in the first and the last magnets x rotates in the +s direction. In the secondand third magnets it rotates in the −s direction. From Eq. (5.1b) we then concludethat ρ is positive in the first and the last magnets and negative in the second and thirdones. The direction of the magnetic field is then determined from Eq. (5.28) takinginto account that q = −e negative for electrons: By is negative in the first and thelast magnets and positive in the second and the third.

If the electron follows the same path in the opposite direction, then all of theB fields change direction to preserve the bending direction. If we keep y in thesame direction as before, the signs of ρ are flipped. Because the direction of theelectron trajectory is reversed, the geometry requires x to be reversed as well, whichis another way of seeing that ρ changes sign even though the vector form of the radiusof curvature is a function of the curve alone.

Problem 5.4 Verify that Eqs. (5.4a)–(5.4e) hold for a circular orbit.

Solution: For a circular orbit, the curvilinear coordinate system (x, y, s) reducesto the standard cylindrical coordinates (r, θ, z) if we establish the following corre-spondence between the unit vectors:

r = x , θ = s , z = − y ,

Page 80: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

5.5 Time-Dependent Hamiltonian 73

while for the coordinates

r = ρ + x , θ = s

ρ, z = −y .

Theminus sign in the y → z correspondence is needed to have the (r, θ, z) coordinatesystem be right-handed. Changing the notation Ax → Ar , Ay → −Az and As → Aθ

it is then a straightforward calculation to check that Eqs. (5.4a)–(5.4e) are the sameas the corresponding differential operators in the cylindrical coordinates.

Problem 5.5 Find the Hamiltonian K for the following model Hamiltonian H :

H(x,�x , s,�s) = �2x

2+ ω2(s)

x2

2+ v�s ,

where v is a constant. Prove that both Hamiltonians describe the same dynamics.

Solution: Solving equation H(x,�x , s,�s) = h for �s we find

K (x,�x , s, h) = −�s = 1

v

(−h + �2

x

2+ ω2(s)

x2

2

).

The equations of motion that follow from the old Hamiltonian are:

x = ∂H

∂�x= �x

�x = −∂H

∂x= −xω2

s = ∂H

∂�s= v

�s = −∂H

∂s= −x2ω

ds.

The new Hamiltonian K has two conjugate variables, x and �x , and the twoequations of motion are

dx

ds= ∂K

∂�x= �x

v

d�x

ds= −∂K

∂x= − xω2

v.

These equations are complemented by Eq. (5.17),

dt

ds= −∂K

∂h= 1

v.

Page 81: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

74 5 Coordinate System and Hamiltonian for a Circular Accelerator

The last equation that we need relates d�s/ds to the rate of change of theHamiltonian K ,

d�s

ds= −dK

ds= −∂K

∂s= − x2ω

v

ds.

We can then compare the two sets of equations, and see that ds/dt = v anddt/ds = 1/v are consistent with each other. The remaining equations differ only bya factor of 1/v, as we expect because we are taking derivatives with respect to sinstead of t . Multiplying by ds/dt = v makes these equations identical.

References

1. J.M. Jowett, Introductory statistical mechanics for electron storage rings. AIP Conf Proc 153,864–970 (1987)

2. K.R. Symon, Derivation of Hamiltonians for accelerators. Report ANL/APS/TB-28, ArgonneNational Laboratory, 1997

3. A.W. Chao, M. Tigner, Handbook of Accelerator Physics and Engineering, 3rd edn. (WorldScientific Publishing, Singapore, 2006)

4. E. Kreyszig, Differential Geometry (Dover Publications, 1991)

Page 82: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

Chapter 6Equations of Motion in Accelerators

A typical accelerator uses a sequence of various types of magnets separated bysections of free space (so-called drifts) to control the motion of the particle beam.To specify the Hamiltonian (5.23) we need to know the vector potential, As , forthese magnets. We evaluate the fields and Hamiltonian for the major magnet typesassuming that the field profiles are uniform over their length. Often in analysis andsimulations, one has to take into account that at the end points of themagnets differentfield geometries appear, called fringe fields. The impact of these fields are usuallytreated as highly localized corrections which are calculated separately from the bulkof the magnet, and involve higher order terms that we will simply neglect in thischapter. When fringe fields are weak they can be treated as field errors, which arecovered in Chap.8.

6.1 Vector Potential for Different Types of Magnets

There are several types of magnets that are used in accelerators and each of them ischaracterized by a specific dependence of the longitudinal component of the vectorpotential, As , versus x and y. In this section, we will list several magnet types andwrite down approximate expressions for As(x, y). In our approximations, wewill usethe fact that we are only interested in fields near the reference orbit, |x |, |y| � |ρ|,so we can neglect higher powers of the ratios x/ρ and y/ρ.

We first consider the dipolemagnets that are used to bend the orbit and, in circularaccelerators, to eventually make it close on itself. Assuming that the field is directedalong y and, in the lowest approximation, neglecting its variation in the transverseplane (that is, neglecting its dependence on x and y), we have

B = yB(s) . (6.1)

© Springer International Publishing AG, part of Springer Nature 2018G. Stupakov and G. Penn, Classical Mechanics and Electromagnetismin Accelerator Physics, Graduate Texts in Physics,https://doi.org/10.1007/978-3-319-90188-6_6

75

Page 83: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

76 6 Equations of Motion in Accelerators

The function B(s) characterizes the longitudinal variation of the field, and vanishesoutside of themagnets.Within the dipole the field can be represented by the followingvector potential:

As = −B(s)x

(1 − x

). (6.2)

Indeed, using Eq. (5.4c) we obtain

By = − 1

1 + x/ρ

∂As(1 + x/ρ)

∂x

≈ B(s)

(1 − x

ρ

)∂

∂x

[x

(1 − x

) (1 + x

ρ

)]

≈ B(s) + O

(x2

ρ2

). (6.3)

This is an approximation in which we only keep terms to the first order in |x/ρ|.The second type is a quadrupole magnet. It is used to focus off-orbit particles so

that they remain close to the reference orbit. It has two components of the field, Bx

and By , that linearly increase with the distance from the axis:

B = G(s)( yx + xy) , (6.4)

where the functionG(s) again isolates the longitudinal variation of the field.Apictureof quadrupole field lines is shown in Fig. 6.1a.

Note that the field on the axis is zero, which means that the reference orbit isa straight line, ρ = ∞. It is straightforward to verify that the corresponding vectorpotential is

As = 1

2G(s)

(y2 − x2

). (6.5)

A skew quadrupole is a normal quadrupole rotated by 45◦:

B = Gsq(s)(− yy + xx) , (6.6)

with

As = Gsq(s)xy . (6.7)

Finally, we also consider a sextupole magnet. Sextupoles are used to correct someproperties of the transverse oscillations of the beam particles around the reference or-bit. This magnet has a nonlinear dependence of the magnetic field with the transversecoordinates:

Page 84: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

6.1 Vector Potential for Different Types of Magnets 77

(a) (b)

Fig. 6.1 Quadrupole (a) and sextupole (b) field lines

B = S(s)

[1

2y(x2 − y2) + xxy

]. (6.8)

The vector potential corresponding to this field is

As = S(s)

(1

2xy2 − 1

6x3

). (6.9)

The field lines of a sextupole are shown in Fig. 6.1b. There is also a skew version ofthe sextupole, and for higher order magnets as well.

6.2 Taylor Expansion of the Hamiltonian

We are now in position to specify the Hamiltonian (5.23) for a circular acceleratorthat has dipole and quadrupole magnets in the ring. Replacing As with the sum ofthe vector potentials (6.2) and (6.5) we obtain

H = −(1 + η)

(1 + x

ρ

) (1 − 1

2P2x − 1

2P2y

)

− e

p0

[−B(s)x

(1 − x

)+ 1

2G(s)

(y2 − x2

)] (1 + x

ρ

)

≈ −1 − η − ηx

ρ+ 1

2P2x + 1

2P2y + x2

2ρ2− e

p0

1

2G(s)

(y2 − x2

), (6.10)

wherewe havemade use of ρ = p0/eB(s) and neglected terms of the third and higherorders (assuming that η, as well as all four canonical variables x , Px , y and Py are ofthe first order). In what follows, we drop the constant first term −1 on the last lineof Eq. (6.10).

Page 85: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

78 6 Equations of Motion in Accelerators

Our main interest in the following chapters will be the case of on-momentumparticles, η = 0. In this case, the Hamiltonian is the sum of two terms correspondingto the vertical and horizontal degrees of freedom,

H = Hx + Hy , (6.11)

with

Hx = 1

2P2x + x2

2ρ2+ 1

2

e

p0G(s)x2 (6.12)

and

Hy = 1

2P2y − 1

2

e

p0G(s)y2 . (6.13)

The fact that the Hamiltonian is split into two pieces each of which involves onlyvariables corresponding to one degree of freedom means that the horizontal andvertical motion are decoupled. The skew quadrupole and sextupole magnets whichhave been left out of this example can in practice be used to correct unintendedcoupling as needed.

The quadrupole magnetic field acts in opposite ways in x and y: positiveG meansthat the effective potential energy in Hx has a minimum on axis x = 0 and leads tostable oscillations around this equilibrium point. At the same time, the effectivepotential energy in Hy has a maximum at y = 0, which is unstable. Negative Gchanges the sign of the effect in x and y. In the following sections, we will show thatnotwithstanding the defocusing effect in one of the transverse directions, a sequenceof quadrupoles with alternating polarities can make the transverse motion stable inboth directions and confine it near the reference orbit. As a result, a particle near theequilibrium orbit executes stable betatron oscillations.

Note that in the absence of quadrupoles, there is a focusing force in the horizontaldirection inside dipole magnets due to the second term in Eq. (6.12) for Hx . Beinginversely proportional to ρ2, this term is typically small and does not play a big rolein the beam dynamics (it is referred to as the weak focusing effect).

To study general properties of the motion in both transverse planes, in the nextsection we will use a generic Hamiltonian

H0(x, Px , s) = 1

2P2x + 1

2K (s)x2 , (6.14)

where K = ρ−2 + eG/p0 for the horizontal plane, and K = −eG/p0 for the verticalplane.

Page 86: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

6.3 Hill’s Equation, Betatron Function and Betatron Phase 79

6.3 Hill’s Equation, Betatron Function and Betatron Phase

From the Hamiltonian (6.14) we find the following equation of motion in the trans-verse plane:

x ′′(s) + K (s)x(s) = 0 , (6.15)

where the prime denotes the derivative with respect to s. In a circular accelerator,K (s) is a periodic function of s with a period that we denote by L (which may beequal to the ring circumference or a fraction of it). Equation (6.15) with a periodicK is called Hill’s equation; it describes the so-called betatron oscillations around thereference orbit. Note that we have encountered the same equation, Eq. (4.16), in thediscussion of the parametric resonance, with the only difference that we now haves as an independent variable instead of t . We know that this equation can have bothstable and unstable solutions. Of course, to successfully contain particles inside thevacuum chamber of an accelerator, one has to design a magnetic system that avoidsunstable solutions.

While we cannot explicitly solve Eq. (6.15) in the general case, some of its funda-mental properties can be studied without specifying the function K (s). Let us seekits solution in the following form,

x(s) = Aw(s) cosψ(s) , (6.16)

where A is an arbitrary constant, and the functionsw(s) and ψ(s)will be determinedby the requirement that (6.16) satisfies (6.15). Note that the function w(s) is notuniquely defined: we can always multiply it by an arbitrary factor w0 and redefinethe amplitude A → A/w0, so that x(s) is not changed. It turns out that if the particlemotion is stable, we can require that w(s) and dψ/ds be periodic functions of s withthe period L . The requirement on the function ψ, which is called the betatron phase,can equivalently be formulated as ψ(s + L) = ψ(s) + ψ0, where ψ0 is a constant.These two properties of the functions w and ψ are guaranteed by the Floquet theoryof differential equations with periodic coefficients [1]. Introducing the two unknownfunctions w(s) and ψ(s) instead of one x(s) gives us the freedom to impose a con-straint of our choice on the functions w and ψ to obtain an optimal parametrizationof the solution.

Substituting (6.16) into (6.15) we obtain

[w′′ − wψ′2 + K (s)w

]cosψ − (

2w′ψ′ + wψ′′) sinψ = 0 . (6.17)

We now use the freedom mentioned above and set to zero each of the terms thatmultiply cosψ and sinψ. Themotivation here is to have a solutionwhich is insensitiveto the constant part of the phase ψ, and instead only depends on its derivatives. Thisgives us two equations:

Page 87: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

80 6 Equations of Motion in Accelerators

w′′ − wψ′2 + K (s)w = 0 ,

−2w′ψ′ − wψ′′ = 0 . (6.18)

The last equation can be written as

1

w(ψ′w2)′ = 0 . (6.19)

At this point we integrate (6.19) and introduce the beta function, β(s) = w2(s),

ψ′ = a

β(s), (6.20)

where a is an arbitrary constant of integration. Without loss of generality, we canassume that a > 0; if this is not the case, we can always change its sign by redefiningthe angleψ → −ψ, which does not change x(s) in (6.16). As was pointed out above,the function w can be multiplied by an arbitrary constant factor. Choosing this factorequal to

√a and replacing β → aβ eliminates a from (6.20) and converts it into

ψ′ = 1

β(s). (6.21)

The first equation in (6.18) now becomes

w′′ − 1

w3+ K (s)w = 0 . (6.22)

Finally, substituting w(s) = √β(s) into (6.22) yields

1

2ββ′′ − 1

4β′2 + K (s)β2 = 1 . (6.23)

We have derived a nonlinear differential equation of the second order for β(s).For a given periodic function K (s), its solution uniquely defines β(s) such thatβ(s) = β(s + L), although in most practical cases the equation can only be solvednumerically, usuallywith dedicated computer codes developed for this purpose.Afterβ(s) is found, the betatron phase ψ is obtained by a straightforward integration ofEq. (6.21). Note that, as follows from (6.21), for a periodic β(s) the derivative ψ′is also periodic with the same period L , thus satisfying the requirement formulatedafter Eq. (6.16). In Fig. 6.2we show the beta functions for theAdvanced Light Sourceat Lawrence Berkeley National Laboratory (LBNL).

An important characteristic of the magnetic lattice of a ring is the betatron phaseadvance over its circumferenceC ,�ψ = ∫ C

0 ds/β(s). The quantity�ψ/2π is calledthe tune ν (also denoted by Q in the European literature),

Page 88: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

6.3 Hill’s Equation, Betatron Function and Betatron Phase 81

Fig. 6.2 The beta functionsβx and βy for the AdvancedLight Source at LBNL

ν = 1

∫ C

0

ds

β(s). (6.24)

As we will see in the following chapters, the tune plays an important role in beamdynamics.

The main advantage of introducing the two functions β(s) and ψ(s) is that, fora given magnetic lattice, they need to be calculated only once. Having found them,the general solution to the equation of motion (6.15) can be written as

x(s) = A√

β(s) cos[ψ(s) − ψ0] , (6.25)

where A and ψ0 are two arbitrary constants that depend on the initial conditions.Note that the phase term only needs to be adjusted by a constant offset for differentinitial conditions. Even without detailed knowledge of β and ψ, this equation givesimportant information about the structure of particle trajectories in the ring.

To conclude this section, we mention that although our analysis was motivated bycircular accelerators, the same representation (6.16) of particle orbits is often usedin linear accelerators. In the absence of the periodicity condition in such machines,to solve Eq. (6.23) one needs either to specify the initial β and its derivative β′ at theentrance to the system or to impose equivalent boundary conditions.

Worked Examples

Problem 6.1 The magnetic field Bs(s) of a solenoid cannot be described with asingle longitudinal component As of the vector potential. Show that this magneticfield can be represented with a vector potential that has two transverse components:

Ax = −1

2Bs(s)y , Ay = 1

2Bs(s)x , As = 0 .

Solution: For the above vector potential, we can use Eqs. (5.4b)–(5.4d) settingρ → ∞ (there is no bending magnetic field inside the solenoid). We then find thatthe fields of a solenoid are

Page 89: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

82 6 Equations of Motion in Accelerators

Bs = −∂Ax

∂y+ ∂Ay

∂x= Bs(s) ,

Bx = −∂Ay

∂s+ ∂As

∂y= − x

2

∂Bs

∂s,

By = −∂As

∂x+ ∂Ax

∂x= − y

2

∂Bs

∂s.

The transverse fields are related to longitudinal derivatives of Bs and vanish insidethe solenoid where Bs is constant. They play a role at the edges of the solenoid andare required to conserve magnetic flux (∇ · B = 0).

Problem 6.2 Using Eq. (5.17) and the Hamiltonian (6.10) show that

dt

ds= 1

v

(1 + x

ρ

).

Explain the meaning of this relation. It follows from this equation that, for a relativis-tic particle, ds/dt can be larger than c. Does this constitute violation of the specialtheory of relativity which forbids motion of bodies faster than the speed of light?

Solution: We need to evaluate dt/ds = −∂K/∂h. Using K = p0H, η = (p −p0)/p0, and p = √

h2/c2 − m2c2 for a relativistic particle, one obtains:

−∂K

∂h= −p0

∂H∂η

∂η

∂ p

∂ p

∂h= −∂H

∂η

h

c2 p= 1

v

(1 + x

ρ

),

where we used the relation v = c2 p/h. It does not violate the special theory ofrelativity, since the change in the s coordinate differs from the actual space travelledby the particle. In particular, if for negative values of x the particle is travelling on acircle of shorter radius than ρ the rate of change of s can be larger than c.

Problem 6.3 Find terms in the HamiltonianH responsible for the skew quadrupole(the magnetic field given by Eq. (6.6)).

Solution: A skew quadrupole has vector potential As = Gsqxy, adding an addi-tional term to the Hamiltonian of

Hsq = − e

p0

(1 + x

ρ

)Gsqxy

≈ − e

p0Gsqxy ,

which makes the motion in x and y coupled.

Problem 6.4 Using the vector potential for the solenoid

Page 90: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

6.3 Hill’s Equation, Betatron Function and Betatron Phase 83

Ax = −1

2Bs y , Ay = 1

2Bsx ,

and starting from the Hamiltonian Eq. (5.18), find the contribution to K of themagnetic field of the solenoid. Assume that x and y, and hence Ax and Ay , are smalland use the Taylor expansion in the Hamiltonian, keeping only linear and secondorder terms.

Solution: We set As = 0 and ρ = ∞ in the Hamiltonian (5.18),

K = −[1

c2h2 − (�x − eAx )

2 − (�y − eAy)2 − m2c2

]1/2

.

Defining p2 = h2/c2 − m2c2 and expanding the square root for small Ax and Ay

yields

K = −p

[1 − 1

p2(�2

x − 2e�x Ax + e2A2x + �2

y − 2e�y Ay + e2A2y)

]1/2

≈ −p + 1

2p0(�2

x − 2e�x Ax + e2A2x + �2

y − 2e�y Ay + e2A2y) .

Substituting the specified values of the vector potential into this expression yields

K ≈ −p + 1

2p0

[�2

x + �2y + 1

4e2B2

s (x2 + y2) + eBs(�x y − �yx)

].

Note that the e2B2s (x

2 + y2)/8p0 term in the Hamiltonian provides focusing in bothtransverse planes, and the last term in K introduces a coupling between x and y.

There is a strong similarity between this expression and the result of Problem 1.2for a coordinate system rotating in time, especially if we were to continue that earlierexercise to calculate the Hamiltonian. Although in the current exercise we considera relativistic particle moving with small transverse angles and we are using s as theindependent variable, this can be taken into account by replacing the mass with thenominal momentum and the rotation frequency by a rate of change with respect tothe coordinate s. The rotating frame result will match the current example when thewavenumber of the rotation in s is set to eBs/2p0. This is half the rate at which aparticle performs transverse oscillations inside the solenoid field.

Problem 6.5 Find the solutions of Eq. (6.23) in free space where K = 0.

Solution: Taking the case of K = 0 (drift space) we rewrite Eq. (6.23) as

1

2β′′ = 1

β+ β′2

4β. (6.26)

Page 91: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

84 6 Equations of Motion in Accelerators

Then differentiating both sides of the equation yields

1

2β′′′ = − β′

β2− β′3

4β2+ β′′β′

= β′

β2

(−1 − 1

4β′2 + 1

2ββ′′

)= 0 .

The vanishing third derivative of β means that it is a quadratic polynomial of s,

β(s) = As2 + Bs + C ,

where A, B and C are constants. They however are not independent which can beobserved by substituting this solution back into Eq. (6.26),

AC = 1 + 1

4B2 .

A more convenient form for the beta function in free space that is often used inpractice is obtained by introducing two parameters: s0 = −B/2A and β0 = 1/A,which then requires C = β0 + s20/β0. This allows us to rewrite the beta function as

β(s) = β0 + 1

β0(s − s0)

2 . (6.27)

The parameter s0 is the location of the minimum of the beta function and β0 is thevalue of the beta function at the minimum.

Problem 6.6 Calculate a jump of the derivative of the beta function through a thinquadrupole. Such a quadrupole is defined by K (s) = f −1δ(s − s0), where f is calledthe focal length of the quadrupole.

Solution: We can integrate Eq. (6.23) over a narrow window around s0:

∫ s0+

s0−ds

1

2ββ′′ −

∫ s0+

s0−ds

1

4β′2 +

∫ s0+

s0−ds

β2

fδ(s − s0) =

∫ s0+

s0−ds .

For an infinitesimally narrow window, the right-hand side goes to 0. We also setthe second term on the left to zero; because β is continuous, β′ must be finite, andthe integral in the limit of a narrow window must go to zero. The first term can beintegrated by parts to find

β0

2

(β′0+ − β′

0−) + f −1β2

0 = 0 .

So the jump in the first derivative of the β function is given by �β′ = −2β/ f .

Page 92: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

6.3 Hill’s Equation, Betatron Function and Betatron Phase 85

Fig. 6.3 The beta functionversus longitudinal positionin a FODO lattice, bothscaled to the periodicitylength l. This example is forK0 = 2/ l

Problem 6.7 A FODO lattice is a sequence of thin quadrupoles with alternatingpolarities:

KFODO(s) =∞∑

n=−∞K0δ(s − nl) − K0δ

(s −

[n + 1

2

]l

),

where l is the period of the lattice. Solve Eq. (6.23) for the FODO lattice and findβ(s). For a given value of l find the maximum value of K0 for which the motion isstable.

Solution: The FODO lattice is periodic, with focusing quadrupole, drift, defo-cusing quadrupole, drift. Its beta function is also periodic with the period l and issymmetric about each quadrupole, see Fig. 6.3.

Due to the symmetry it is sufficient to find β(s) in the first half of the cell,0 < s < l/2, where we can use Eq. (6.27) in which β0 and s0 are two unknowns.

From the previous problemwe can identifyK0 = 1/ f , where f is the focal length,and calculate the jump of the beta function �β′ = −2β(0)/ f at s = 0, and the jump�β′ = −2β(l/2)/ f at s = l/2. From the symmetry of the beta function we concludethat the values of the derivatives of the beta function at the location of the quadrupolesare equal to the halves of the corresponding jumps:

β′(0+) = − 1

fβ(0), β′

(l

2− 0

)= − 1

(l

2

).

Calculating β′(s) from Eq. (6.27) and substituting it into these equations we obtaintwo equations for the two unknowns β0 and s0,

2s0β0

= 1

f

(β0 + s20

β0

), 2

s0 − l/2

β0= 1

f

[β0 + (s0 − l/2)2

β0

].

Page 93: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

86 6 Equations of Motion in Accelerators

Solving these equations we find

s0 = 1

4(4 f + l) , β0 = 1

4

√16 f 2 − l2 .

This yields a constraint that l < 4 f so that β is real. This is as expected — if thedrift is much longer than the focusing length, we would not expect the focusing tobe sufficient. In terms of the magnetic strength this condition becomes |K0| < 4/ l.

For the maximum and minimum values of the beta function we find

βmax = β(0) = 2 f

√4 f + l

4 f − l, βmin = β

(l

2

)= 2 f

√4 f − l

4 f + l.

Problem 6.8 Consider two rings with circumferencesC1 andC2. Assume thatC1 =λC2 and K2(s) = λ2K1(λs), and prove that β2(s) = λ−1β1(λs).

Solution: The second ring is related to the first by having its circumference andall lengths scaled by 1/λ, and all focusing strengths around the ring scaled by λ2.We start from the equation for β2,

1

2β2β

′′2 − 1

4β′22 + K2β

22 = 1 .

If we take β1(λs) = λβ2(s), we find β′2(s) = (1/λ)β′

1(λs)λ = β′1(λs) and β′′

2 (s) =λβ′′

1 (λs). For each derivative, we extract an extra factor of λ. Substituting in for β1

above we find

1

2β1β

′′1 − 1

4β′21 + K1β

21 = 1 ,

confirming that β2 is a valid solution so long as β1 is a solution for its ring.

Reference

1. W. Magnus, S. Winkler. Hill’s Equation (Dover Publications, Mineola, 2004)

Page 94: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

Chapter 7Action-Angle Variables for BetatronOscillations

In Chap. 3 we showed that choosing the action-angle canonical variables in a one-dimensional Hamiltonian system dramatically simplifies the dynamics: the actionremains constant and the angle increases linearly with time. With minor modifica-tions, the same transformation can be applied to theHamiltonian (6.14) that describesbetatron oscillations in an accelerator. This yields an invariant of the motion and isalso a useful starting point for analyzing more complicated dynamics.

7.1 Action-Angle Variables

In Sect. 6.3, we proved that the general solution of the equations of motion for theHamiltonian (6.14) can be written in the form (6.25),

x(s) = A√

β(s) cosψ(s) , (7.1)

(where ψ0 is now included into ψ) which looks similar to the solution (4.1) of thelinear harmonic oscillator, although the amplitude A

√β(s) varies with position, and

the phase ψ(s) is not necessarily a linear function of s. The canonical momentum Pxcorresponding to x(s) is equal to x ′(s) and is obtained by differentiating the aboveequation,

Px (s) = x ′(s) = A√βcosψ(s)

(β′

2− tanψ(s)

)

= x

β

(β′

2− tanψ(s)

). (7.2)

© Springer International Publishing AG, part of Springer Nature 2018G. Stupakov and G. Penn, Classical Mechanics and Electromagnetismin Accelerator Physics, Graduate Texts in Physics,https://doi.org/10.1007/978-3-319-90188-6_7

87

Page 95: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

88 7 Action-Angle Variables for Betatron Oscillations

Fig. 7.1 Plots of x and x ′ versus s for a particular solution to Eq. (6.15) with initial conditionsx(0) = 0 and x ′(0) = 1

Fig. 7.2 Function K (s) (left panel) and the corresponding β function (right panel)

These two dependences are illustrated in Fig. 7.1 for a particular choice of functionK (s) shown in the left panel of Fig. 7.2 — a sequence of rectangular pulses ofopposite polarity with the period equal to 3. One can see that both functions showa complicated pattern consisting of an approximately sinusoidal oscillation with asuperimposed jitter caused by the jumps in the function K (s).Wewill now show howto “rectify” this complicated dynamics making a transformation to the action-anglevariables φ and J .

We will use the same approach as in the case of the linear harmonic oscillator,i.e., assuming that A = A(J ) and replacing the phase in the representation of thesolution (7.1) by the angular variable φ:

x(s) = A(J )√

β(s) cosφ (7.3a)

Px (s) = − x

β(s)(α + tan φ) , (7.3b)

where we have introduced the notation

α(s) = −1

2β′(s) . (7.4)

Page 96: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

7.1 Action-Angle Variables 89

The next step is to calculate the generating function (3.6) of the first kind,

F1(x,φ, s) =∫

Pxdx = − x2

2β(α + tan φ) . (7.5)

With this generating function, we find the action

J = −∂F1

∂φ= x2

2β(1 + tan2 φ) , (7.6)

and expressing tan φ from Eq. (7.3b),

tan φ = −βPxx

− α , (7.7)

we obtain J in terms of x and Px :

J = 1

[x2 + (βPx + αx)2

]. (7.8)

Equations (7.7) and (7.8) give us the transformation from the old conjugate variablesx and Px to the new ones φ and J . The inverse transformation (φ, J ) → (x, Px ) canalso be found. From Eq. (7.6) we have

x = √2β J cosφ . (7.9)

Substituting this relation to Eq. (7.3b) we find Px in terms of J and φ,

Px = −√2J

β(sin φ + α cosφ) . (7.10)

To find the new Hamiltonian which we denote by H(φ, J ), we need to take intoaccount that the generating function depends on the time-like variable s and useEq. (2.28b):

H = H + ∂F1

∂s

= 1

2P2x + 1

2K (s)x2 + x2

4

β′′β − β′2

β2+ x2β′

2β2tan φ . (7.11)

With the help of Eq. (6.23) we eliminate β′′ from this equation and then substituteEq. (7.7) for tan φ:

Page 97: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

90 7 Action-Angle Variables for Betatron Oscillations

H = 1

2P2x + 1

2β2x2 + α2

2β2x2 + α

βPx x

= J

β. (7.12)

In the last step we have used the definition (7.8) of J in terms of the old variables.Since the new Hamiltonian is independent of φ the equation of motion for J is

J ′ = −∂H∂φ

= 0 , (7.13)

which means that J is an integral of motion. The quantity 2J is called the Courant-Snyder invariant [1]. The Hamiltonian equation for φ reads

φ′ = ∂H∂ J

= 1

β(s). (7.14)

Comparing this equation with Eq. (6.21) we see that the new coordinate φ is actuallyequal to the old betatron phase, φ = ψ + φ0. This is not surprising, because wereplaced the betatron phase in Eqs. (7.1) and (7.2) by φ to arrive at Eqs. (7.3a, 7.3b).

Note that Eq. (7.12) is not a universal expression relating the Hamiltonian to theaction in action-angle coordinates. Equation (7.12) is a special, particularly simpleform that is valid when the focusing is linear and the action is directly obtained fromthe particle’s physical displacement.

7.2 Eliminating Phase Oscillations

As follows from Eq. (7.14), the phase coordinate φmonotonically grows with s, witha rate of change that oscillates around some average value due to the oscillations of thebeta function as shown in Fig. 7.2b. We can do one more canonical transformation ofthe variables and “straighten out” these oscillations. This transformation replaces φ,J with new canonical variablesφ1 and J1. It is carried outwith the help of a generatingfunction of the second type, F2(φ, J1, s), which is given by the following equation:

F2(φ, J1, s) = J1

(2πνs

C−

∫ s

0

ds ′

β(s ′)

)+ φJ1 , (7.15)

whereC is the circumference of the accelerator and the tune ν is given by Eq. (6.24).The new angle is given by the derivative of F2 with respect to J1,

φ1 = ∂F2

∂ J1= φ + 2πνs

C−

∫ s

0

ds ′

β(s ′)= φ + 2πνs

C− ψ(s) , (7.16)

Page 98: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

7.2 Eliminating Phase Oscillations 91

and the action is unchanged,

J = ∂F2

∂φ= J1 . (7.17)

We denote the new Hamiltonian by H1,

H1 = H + ∂F2

∂s= 2πν

CJ1 . (7.18)

With this Hamiltonian, the equation of motion for φ1 reads

φ′1 = ∂H1

∂ J1= 2πν

C, (7.19)

which means that φ1 is a linear function of s with the slope given by 2πν/C . We seethat, indeed, we got rid of the oscillations exhibited by the phaseφ and obtained a newphase that follows a straight line in s. The tune is identical in both sets of coordinates.

7.3 Phase Space Motion at a Given Location

As a particle travels in a circular accelerator, every revolution period it arrives at thesame longitudinal position s. Let us consider the phase plane x , Px at this locationand plot the particle coordinates every time it passes through s. Because there isan integral of motion J , all these points are located on the curve J = const. Fromexpression (7.8) for J it follows that this curve is an ellipsewhose size and orientationdepend on the values of J , β, and α. A set of consecutive points xi , Px,i , i = 1, 2 . . .,are shown in Fig. 7.3, where for convenience x is normalized by the beta functionat this location, β(s). Particles with different values of J have geometrically similarellipses enclosed inside each other.

From expression (7.8) it is easy to see that the ellipse turns into a circle if α = 0and again x is normalized by β(s). In this case, the trajectory is very simple: on eachrevolution the representative point rotates by the betatron phase advance in the ring�ψ = 2πν in the clockwise direction.

A set of ellipses at another location in the ring will have a different shape which isdefined by the local values of β and α. When one travels along the circumference ofthe accelerator, one sees a continuous transformation of these sets with the coordinates. For a collection of particles in a bunch, this effect includes changes not only in thesize of the beam but also in statistical correlations between x and Px .

Worked Examples

Problem 7.1 UsingEqs. (7.7) and (7.8), showbydirect calculationofPoissonbrack-ets that the transformation x, Px → φ, J is canonical.

Page 99: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

92 7 Action-Angle Variables for Betatron Oscillations

Fig. 7.3 The phase spaceellipse (solid curve) and aparticle’s positions atconsecutive turns. Dashedlines show ellipses forparticles with smaller andlarger values of action J .The vertical axis is markedby x ′ which is equal to Px

Solution: The only Poisson bracket that needs to be evaluated is {φ, J }x,Px , because itis trivially true that {φ,φ} = {J, J } = 0. The coordinate transformation is defined by

tan φ = −βPxx

− α , J = 1

[x2 + (βPx + αx)2

].

We use the fact that d(tan φ) = (1/ cos2 φ)dφ to express

∂φ

∂x= cos2 φ

∂ tan φ

∂x,

and similarly for ∂φ/∂Px . This yields

{φ, J }x,Px = ∂φ

∂x

∂ J

∂Px− ∂φ

∂Px

∂ J

∂x

= (cos2 φ)βPxx2

1

2β2(βPx + αx)β

− (cos2 φ)

(−β

x

)1

2β[2x + 2(βPx + αx)α]

= cos2 φ

[

1 + (βPx + αx)2

x2

]

= cos2 φ(1 + tan2 φ

) = 1 .

Thus, the value of the Poisson bracket is 1 as required for a canonical transformation.

Problem 7.2 Find the major and minor half axes, and the tilt of the ellipse shownin Fig. 7.3 which is based on Eq. (7.8).

Page 100: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

7.3 Phase Space Motion at a Given Location 93

Solution: In Eq. (7.8) we change variables to ζ = x/β and multiply both sides by2/β to obtain:

2J

β= ζ2 + (Px + αζ)2 = (1 + α2)ζ2 + P2

x + 2αPxζ .

This is the formula for an ellipse, which, because of the Pxζ cross term, is tilted. Welook for a rotation to new variables (r, u) defined by

ζ = r cos θ − u sin θ ,

Px = r sin θ + u cos θ ,

such that we no longer have a cross term. Substituting in above, we find that the crossterm is

2αru(cos2 θ − sin2 θ − α sin θ cos θ

).

For α �= 0 (otherwise the ellipse is already a circle with radius√2J/β), setting this

to zero requires α = 2 cos 2θ/ sin 2θ so

tan 2θ = 2

α.

This last expression can also be obtained by looking for an extremum in ζ2 + (Px )2

by calculating the variation d J under the constraint that ζdζ + PxdPx = 0, and thensolving for d J = 0. This yields the equation

(Pxζ

)2

+ αPxζ

− 1 = 0 ,

which gives the same condition after setting Px/ζ = tan θ and multiplying by cos2 θ.The major and minor half axes are just the maximum and minimum value of

the radius under fixed 2J/β and fixed α. As we already defined a pair of rotatedcoordinates for which the ellipse is upright, we can rewrite the equation for theellipse in terms of r and u:

2J

β= r2

(1 + α2 cos2 θ + 2α sin θ cos θ

) + u2(1 + α2 sin2 θ − 2α sin θ cos θ

)

= r2 cot2 θ + u2 tan2 θ ,

where the second equality comes from eliminating α using α = 2/ tan 2θ, whichjustifies dropping the ru cross term, and from trigonometric identities. Themaximumand minimum values are found by setting either u = 0 or r = 0, giving the twovalues for the radius or half-axis in these scaled coordinates as

√2J/β | tan θ| and

Page 101: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

94 7 Action-Angle Variables for Betatron Oscillations

√2J/β | cot θ|; forα > 0 the tan θ termwill be the shorter, minor axis, and forα < 0

it will be the longer, major axis.

Problem 7.3 Prove that the transformation x, Px → x, Px with

x = 1√βx , Px = 1√

β(βPx + αx) ,

is canonical. Prove that phase space orbits plotted in variables x , Px are circles.

Solution: For a linear transformation in one degree of freedom, in order to show it issymplectic and thus canonical it is sufficient to show that the determinant is 1. Thetransformation is equivalent to

(xPx

)=

⎜⎜⎝

1√β

0

α√β

√β

⎟⎟⎠

(xPx

),

and the determinant is clearly equal to 1.This transformation is canonical regardless of what the quantities x and Px

represent. When we use the original parametrization for the particle position,x = √

2β J cosφ, with the corresponding derivative being Px , we find that

x = √2J cosφ ,

and

Px = √2J sin φ .

Note that Px is β times the derivative of x . Because the action J is constant, theseexpressions give the equation for a circle with radius

√2J .

Reference

1. E.D. Courant, H.S. Snyder, Theory of the alternating-gradient synchrotron. Ann. Phys. 3, 1–48(1958)

Page 102: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

Chapter 8Magnetic Field and Energy Errors

Due to small magnetic field errors the magnetic field in any real machine differs fromthe ideal linear lattice studied in the previous chapters. Also, the particle energy in thebeam deviates from the reference energy for which the ideal closed orbit is designed.It is important to understandwhat new effects the non-ideal magnetic fields introduceto the particle motion. In this chapter, we study how the dipole and quadrupole fielderrors as well as small energy deviations affect the orbits.

8.1 Closed Orbit Distortions

We first consider what happens with the reference orbit if the dipole magnetic fieldis not exactly equal to the design one.

Let B0(s) be the design vertical magnetic field and consider the magnetic fielderror to be also directed along y with the total field B(s) = y[B0(s) + �B(s)]. Thedeviation from the ideal field �B is small, |�B| � B0. Our curvilinear coordinatesystem x, y, s is associated with the ideal reference orbit of the magnetic field B0,which is the line x = y = 0. The vector potential �As that describes the error fieldis given by Eq. (6.2) in which, because of the smallness of �B, we keep only thefirst order term, �As = −�B(s)x . This vector potential should be added to theHamiltonian (6.12):

Hx = 1

2P2x + 1

2K (s)x2 + e�B(s)

p0x , (8.1)

where K = ρ−2 + eG/p0. From this Hamiltonian we find the following differentialequation for x ,

x ′′ + K (s)x = −e�B(s)

p0. (8.2)

© Springer International Publishing AG, part of Springer Nature 2018G. Stupakov and G. Penn, Classical Mechanics and Electromagnetismin Accelerator Physics, Graduate Texts in Physics,https://doi.org/10.1007/978-3-319-90188-6_8

95

Page 103: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

96 8 Magnetic Field and Energy Errors

This equation describes general betatron oscillations in the accelerator, but can alsobe used to find a new closed orbit in the perturbed magnetic field B(s). Because theperturbation is small, this new orbit should be close to the old one; it is obtained as aperiodic solution to Eq. (8.2). We denote this solution by x0(s); it is called the closedorbit distortion. By definition, closed orbits must satisfy the periodicity conditionx0(s + C) = x0(s), where C is the circumference of the ring.

To calculate x0(s)we first consider the case of a field perturbation localized at onepoint: �B(s) = �B0(s ′)δ(s − s ′). Since the right-hand side of Eq. (8.2) is equal tozero everywhere except for the point s = s ′, we seek a solution in the form of (6.25)with an unknown amplitude A and an initial phase ψ0,

x0(s) = A√

β(s) cos[ψ(s) − ψ0] . (8.3)

To be a periodic solution to (8.2), this function should satisfy two requirements. Thefirst one is that x0(s) should be continuous at s = s ′. This is achieved if we chooseψ0 = ψ(s ′) + πν and agree thatψ(s) varies fromψ(s ′) toψ(s ′) + 2πν when s variesfrom s ′ to s ′ + C . Indeed, when ψ(s) = ψ(s ′), the argument of the cos function isequal to −πν, and x0(s ′) = A

√β(s ′) cos(−πν). After one turn around the ring, the

argument of the cos function becomes equal to πν, and since cos is an even function,x0(s ′ + C) = x0(s ′). Hence,

x0(s) = A√

β(s) cos[ψ(s) − ψ(s ′) − πν] . (8.4)

The second requirement is obtained when we integrate Eq. (8.2) over s from s ′ − εto s ′ + ε, where ε is an infinitesimally small number. The delta function in �B(s)gives a nonzero value on the right-hand side, and the integration of the left-hand sidegives a jump in the first derivative of x0 at s ′:

x ′0(s

′) − x ′0(s

′ + C) = −e�B0(s ′)p0

. (8.5)

Note that according to our convention of measuring s from s ′ to s ′ + C (whichmeansthat s cannot be smaller than s ′), for the derivative of x0 before the delta function weuse the notation x ′

0(s′ + C) instead of x ′

0(s′ − ε). Substituting (8.4) into (8.5) we find

A = −√

β(s ′)2 sin(πν)

e�B0(s ′)p0

. (8.6)

Having solved the problem for the delta-function perturbation, we can now extendthis solution to the case of an arbitrary �B(s). For this, we represent B(s) as asuperposition of delta-function contributions using the following identity:

�B(s) =∫ s+C

sds ′�B(s ′)δ(s − s ′) . (8.7)

Page 104: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

8.1 Closed Orbit Distortions 97

Fig. 8.1 An ideal (dashedblue line) and a distorted(solid orange line) orbit in aring. The thick green lineshows a betatron orbit withthe offset ξ(s) measuredrelative to the distorted orbit.Note that both the ideal anddistorted orbits are closedlines, but the betatronoscillation is not

Because x0 is linear in �B, we need to add contributions from all locations s ′,integrating the right-hand side of Eq. (8.4) over s ′:

x0(s) = −e

2p0 sin(πν)

∫ s+C

sds′�B(s′)

√β(s)β(s′) cos[ψ(s) − ψ(s′) − πν] . (8.8)

An important immediate consequence of this formula is that integer values for thetune ν are not allowed in a realistic magnetic lattice: such a lattice would be unstablewith respect to small errors of the ideal magnetic field.

We found x0 as a particular solution of the inhomogeneous differential equa-tion (8.2). To find its general solution x(s) we seek it as a sum

x(s) = x0(s) + ξ(s) . (8.9)

Substituting this into (8.2) we see that ξ(s) satisfies the homogeneous equation

ξ′′ + K (s)ξ = 0 , (8.10)

which is the equation for the horizontal betatron oscillations. Thismakes themeaningof ξ(s) clear — it is the betatron oscillation around the perturbed closed orbit x0(s).This is illustrated by Fig. 8.1.

The action-angle variables (7.8) derived in Sect. 7.1 for theHamiltonian (6.14) canbe naturally generalized for the Hamiltonian (8.1). The result of this generalizationcan be understood if we recall that in the expression for the action (7.8) the coordinatex and the momentum Px = x ′ are measured relative to the ideal closed orbit x = 0.With the perturbedmagnetic field, both the distance and the angle has to be measuredrelative to the perturbed closed orbit x0(s). It is not surprising then that the new actionis obtained from (7.8) by simply replacing x → x − x0(s) and Px → Px − x ′

0(s):

J (x, Px , s) = 1

{(x − x0)

2 + [β(Px − x ′

0) + α(x − x0)]2}

. (8.11)

The derivation of this result is the subject of Problem 8.1.

Page 105: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

98 8 Magnetic Field and Energy Errors

As the perturbation of the fields becomes larger, higher-order terms that wereneglected in this derivation become more important. At some point, it may evenbecome impossible to find a closed orbit at all. Our derivation was carried out for thecase when the perturbation of the magnetic field is vertical, i.e., in the same directionas themainmagnetic field. The case of the horizontal perturbation, B(s) = yB0(s) +x�B(s), can be considered in a similar fashion: to lowest order the perturbationpropagates into the Hamiltonian Hy , Eq. (6.13), and the perturbed closed orbit getsa distortion y0(s) in the vertical direction.

8.2 Effect of Energy Deviation

Another effect that causes a distortion of the closed orbit is the deviation of the particleenergy from the nominal one. We can directly adapt the results of the previous sectionto find the distortion for a particle with a relative energy that differs from the nominalone by η. From the Hamiltonian (6.10), we see that it has a term −ηx/ρ that couplesη to the horizontal motion. Hence, instead of (6.12) we have

Hx = 1

2P2x + 1

2K (s)x2 − η

ρx . (8.12)

This Hamiltonian can be formally obtained from (8.1) by the replacement

�B → −η p0eρ

. (8.13)

Making this replacement in Eq. (8.8), we find the orbit distortion caused by theenergy deviation η,

x0(s) = D(s)η , (8.14)

where the function D is

D(s) = 1

2 sin(πν)

∫ s+C

sds ′

√β(s)β(s ′)ρ(s ′)

cos[ψ(s) − ψ(s ′) − πν] . (8.15)

This function is called the dispersion function of the ring, and it too is a periodicfunction of s.

Using the expressions (8.11) and (8.14) we can also find the action variable for aparticle with an energy deviation η,

J (x, Px , η, s) = 1

([x − ηD(s)]2 + {

β[Px − ηD′(s)] + α[x − ηD(s)]}2)

.

(8.16)

Page 106: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

8.3 Quadrupole Errors 99

8.3 Quadrupole Errors

Wewill now address the issue of errors in the quadrupole component of the magneticfield in the ring. Let us assume that due to a perturbation of the quadrupole field theideal focusing strength K (s) is changed to K (s) + �K (s), where |�K | � |K |. TheperturbedHamiltonian is obtained through the replacement K (s) → K (s) + �K (s),

H = 1

2P2x + 1

2K (s)x2 + 1

2ε�K (s)x2 , (8.17)

wherewe have introduced a formal smallness parameter εwhichwill be set to unity atthe end of the calculation. With the proper choice of function K (s), this Hamiltoniancan be applied to both horizontal (x) and vertical (y) coordinates. Since we knowthat the focusing function K (s) determines the betatron oscillations in the system,it is clear that changing the focusing strength would result in the perturbation of thebeta function and, hence, the tune of the ring.

To calculate these changes we first transform to the action-angle variables J and φdefined in Sect. 7.1. This transformation casts the first two terms of the Hamiltonianinto J/β:

1

2P2x + 1

2K (s)x2 → J

β(s). (8.18)

In the last term of Eq. (8.17) we express x in terms of J and φ using Eq. (7.9):

H = J

β(s)+ ε�K (s)Jβ(s) cos2 φ

= J

(1

β(s)+ 1

2ε�K (s)β(s)

)+ 1

2ε�K (s)Jβ(s) cos 2φ , (8.19)

where we have split the perturbation term into an averaged part and one oscillatingas cos 2φ. We will denote the last term in this equation by εV (φ, J, s):

V (φ, J, s) = 1

2�K (s)Jβ(s) cos 2φ . (8.20)

To solve for the Hamiltonian (8.19) we will use perturbation theory based on canon-ical transformations with the goal of eliminating the perturbation term (8.20) fromthe transformed Hamiltonian, thereby removing any dependence on the angular vari-able. If this goal is achieved and the transformed Hamiltonian becomes dependentonly on the action, this action is a constant of motion and the equation of motionfor the angular variable that is conjugate to the action can be easily integrated, asdiscussed in Sects. 7.1 and 7.2. As it turns out, there is no general method that com-pletely eliminates the angular dependence on φ; however, this dependence can be

Page 107: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

100 8 Magnetic Field and Energy Errors

made of higher order in ε. For our purposes, it is enough to obtain the transformedHamiltonian with an accuracy of the order of ε2.

Our first step is to make a canonical transformation to new variables, (φ, J ) →(ξ, I ), where ξ and I are the new angle and action.Wewill use the generating functionF2 of the second type,

F2(φ, I, s) = φI + εG(φ, I, s) , (8.21)

where the appropriate function G will be determined below. Using Eq. (2.29a) weobtain the following relations between the old and new variables:

ξ = φ + εGI (φ, I, s) , J = I + εGφ(φ, I, s) , (8.22)

where we use the subscripts I and φ to denote differentiation with respect to thecorresponding variables. Because G is multiplied by the small parameter ε, the dif-ference between the old and the new variables is small, on the order of ε. We cansolve these equations to the first order in ε:

ξ ≈ φ + εGI (φ, J, s) , I ≈ J − εGφ(φ, J, s) , (8.23)

where terms of the order of ε2 are neglected.To find the new Hamiltonian,H1(ξ, I, s), we use Eq. (2.29b) and then express φ

and J in terms of I and ξ using (8.23) and neglecting terms of the second and higherorder in ε,

H1 = I

(1

β+ 1

2ε�Kβ

)+ εV + 1

βεGφ + εGs + O(ε2) . (8.24)

We can cancel out the angle-dependent part of the perturbation V in the new Hamil-tonian by choosing G in such a way that the φ-dependent terms in (8.24), which areall linear in ε, cancel:

V + 1

βGφ + Gs = 0 . (8.25)

Note that because the old and new variables differ by small terms of the order of ε,we can write V as if it were a function of the variables φ, I, s to match the form ofthe function G from the generating function. It does not matter that H1 has to beexpressed as a function of I, ξ, s because we are ensuring that to first order in ε thedependence on φ will vanish.

One needs to find a solution to (8.25) that is periodic in s with the period equalto the ring circumference C . This solution is given by the expression

G(φ, I, s) = − I

4 sin(2πν)

∫ s+C

sds′�K (s′)β(s′) sin 2[φ − ψ(s) + ψ(s′) − πν] . (8.26)

Page 108: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

8.3 Quadrupole Errors 101

Indeed, this function is periodic in s, because the integrand does not change whenboth s ′ and s change to s ′ + C and s + C , respectively. It also satisfies Eq. (8.25) asis directly established by differentiation of the right-hand side of (8.26). We see thatto avoid a singularity in the formula we need to require that ν is not equal to integeror half-integer values.

With the function G defined above, the new Hamiltonian becomes

H1 = I

(1

β+ 1

2ε�Kβ

), (8.27)

where we have neglected higher order terms. Because H1 does not depend on theangle ξ, the action I is an integral of motion. Equation (8.23) now gives the trans-formation from the old to new variables to the lowest order in ε. For the new actionwe have

I = J + εJ

2 sin(2πν)

∫ s+C

sds ′�K (s ′)β(s ′) cos 2[φ − ψ(s) + ψ(s ′) − πν] .

(8.28)

Changing the strength of the focusing in the lattice by ε�K also modifies the betafunction in the ring. Let us denote the new beta function and corresponding betatronphase by β1 and φ1, respectively. It is tempting to use the variable ξ given by thefirst equation in Eq. (8.22) for the new betatron phase φ1, but this is not justified. Wehave already seen from Eq. (7.17) that the same action can be paired with differentconjugate angle coordinates, and by comparing the expressions (7.14) and (7.19)we note that the derivative of the phase terms are different. In the same example,we found that the form of the Hamiltonian also changed, so we also cannot simplyassume H1 = I/β1. The proper definition of the beta function and betatron phasecomes from Eqs. (7.8)–(7.10). These expressions should hold equally in the newaction-angle coordinates by taking J → I , β → β1 and φ → φ1:

x = √2β1 I cosφ1 ,

Px = −√2I

β1(sin φ1 + α1 cosφ1) ,

I = 1

2β1

[x2 + (β1Px + α1x)

2] , (8.29)

where we again define α1 = −β′1/2.

Even though both β1 and φ1 are unknown, we can overcome this by using the factthat they are only functions of s, and that the above expressions have to hold true forall combinations of x and Px . Thus we may limit our attention to the case φ = π/2.This requires x = 0, so φ1 is constrained to be π/2 as well and

J = 1

2βP2

x , I = 1

2β1P

2x , (8.30)

Page 109: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

102 8 Magnetic Field and Energy Errors

which implies

β1 = β

JI |φ=π/2 . (8.31)

We now use Eq. (8.28), our solution for the new action to first order in the perturbedfields. For brevity, we write this expression as I = J + ε�J , and similarly we willwrite the new beta function to first order in ε as β1 = β + ε�β, where ε�β is thedifference between the new and the old beta functions. Then

�β(s) = β(s)

J�J |φ=π/2

= β(s)

2 sin(2πν)

∫ s+C

sds ′�K (s ′)β(s ′) cos 2[π/2 − ψ(s) + ψ(s ′) − πν]

= − β(s)

2 sin(2πν)

∫ s+C

sds ′�K (s ′)β(s ′) cos 2[−ψ(s) + ψ(s ′) − πν] ,

(8.32)

using cos(π + θ) = − cos θ.An important conclusion that follows from the above equation is that one should

avoid integer or half-integer values of the tune — they are unstable with respect toerrors in the focusing strength of the lattice. Having found the correction to the betafunction, we can find the correction to the tune by using Eq. (6.24).

Worked Examples

Problem 8.1 The action-angle variables defined by Eqs. (7.8)–(7.10) have to bemodified in the case of dipole field errors. Starting from the Hamiltonian (8.1) trans-form to the action-angle variables using the following generating function:

F1(x,φ, s) = [x − x0(s)]22β(s)

(β′(s)2

− tan φ

)+ xx ′

0(s) .

Show that in this case

J (x, Px , s) = 1

2β(s)

([x − x0(s)]2 + {

β(s)[Px − x ′0(s)] + α(s)[x − x0(s)]

}2),

and obtain the Hamiltonian (7.12).

Solution: For the given generating function of the first type the new, action-anglevariables satisfy

J = −∂F1

∂φ= [x − x0(s)]2

2β(s)sec2 φ ,

Page 110: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

8.3 Quadrupole Errors 103

and

Px = −∂F1

∂x= x − x0(s)

β(s)

(β′(s)2

− tan φ

)+ x ′

0(s) . (8.33)

We want to express J in terms of x and Px . Solving for tan φ from our expressionfor Px we find (in the following we leave out most of the s-dependences):

tan φ = β′

2− β(Px − x ′

0)

x − x0= −β(Px − x ′

0) + α(x − x0)

x − x0,

and using sec2 = 1 + tan2 the result is

J = 1

{(x − x0)

2 + [β(Px − x ′

0) + α(x − x0)]2}

.

Finally, our new Hamiltonian is related to the old one by

H = H + ∂F1

∂s= 1

2P2x + 1

2K (s)x2 + e�B(s)

px + (x − x0)

2

2β2

1

2ββ′′

− (Px − x ′0)

[x ′0 − α

β(x − x0)

]+ xx ′′

0 ,

where we used α = −β′/2 and the fact that

β′

2− tan φ = β(Px − x ′

0)

x − x ′0

,

that follows from Eq. (8.33). Finally, we use Eq. (6.23) for β′′ and the fact that x0 isa solution of Eq. (8.2) to obtain:

H = 1

2(Px − x ′

0)2 + 1 + α2

2β2(x − x0)

2 + α

β(x − x0)(Px − x ′

0) + 1

2(x ′

0)2 − 1

2Kx20

= J

β+ 1

2(x ′

0)2 − 1

2Kx20 .

The last two terms are functions of s only and do not affect the dynamics. If we hadchosen the last term in F1 to be (x − x0)x ′

0, then the residual term would be simplyx0e�B/2p.

Problem 8.2 For a ring, how does the sensitivity of the beam orbit in the horizontalplane to the size of the error magnetic field �By(s) scale with the parameters of thering?

Page 111: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

104 8 Magnetic Field and Energy Errors

Solution: A crude estimate of the magnitude of the closed orbit distortion fromEq. (8.8) is

|x0| ∼ e

p0 sin(πν)�BβC ,

where β is the typical magnitude of the beta function in the ring. From the definitionof the bending radius,

|x0| ∼ 1

sin(πν)

�B

BβC

ρ,

and usually the circumference is roughly 2πρ. We see from this equation that theeffect of magnetic field errors is less in machines with small beta functions, that iswith stronger focusing. The effect of errors can be greatly enhanced when the tuneis close to an integer.

Problem 8.3 Verify by direct calculation that G given by Eq. (8.26) satisfiesEq. (8.25).

Solution: We start with G(φ, I, s), and keep in mind that all partial derivativesinvolve holding two of the arguments fixed while varying the third one. To calculateGφ = ∂G/∂φ, we note thatφonly appears in the sin termat the end, and the derivativechanges this to 2 cos 2[φ − ψ(s) + ψ(s ′) − πν]. To evaluate Gs = ∂G/∂s, we notethat in addition to ψ(s) both limits of integration also depend on s. We can calculatethese terms separately, then add them. The ψ(s) derivative changes the sin termat the end to −(2dψ/ds) cos 2[φ − ψ(s) + ψ(s ′) − πν]. Because dψ/ds = 1/β(s),this exactly cancels (1/β)Gφ. The remainder comes from evaluating the integrandat s ′ = s + C and s ′ = s, which combine to form

− I

4 sin(2πν)�K (s)β(s)

{sin 2[φ − ψ(s) + ψ(s + C) − πν]

− sin 2[φ − ψ(s) + ψ(s) − πν]} ,

where we use the fact that �K and β are periodic in s. The phases are not periodic,however; ψ(s + C) = ψ(s) + 2πν by definition of the tune. This gives

1

βGφ + Gs = − I

4 sin(2πν)�K (s)β(s) [sin 2(φ + πν) − sin 2(φ − πν)]

= − I

4 sin(2πν)�K (s)β(s)2 cos(2φ) sin(2πν) = −1

2I�K (s)β(s) cos(2φ) .

This is equal and opposite to V up to order ε, and because these terms are all smallto begin with the difference will be neglected as being O(ε2).

Page 112: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

8.3 Quadrupole Errors 105

Fig. 8.2 Transformation ofthe integration domain in thedouble integral

Problem 8.4 Show that due to the perturbation of the quadrupole strength �K (s)the tune change is given by the following equation:

�ν = 1

∫ C

0ds �K (s)β(s) . (8.34)

Solution: From the definition of the tune we find

�ν = 1

∫ C

0ds �

(1

β(s)

)� − 1

∫ C

0ds

�β(s)

β2(s)

= 1

4π sin(2πν)

∫ C

0

ds

β(s)

∫ s+C

sds ′ �K (s ′)β(s ′) cos 2

[−ψ(s) + ψ(s ′) − πν],

where we used Eq. (8.32). Note that because dψ(s)/ds = 1/β(s), the quantity

1

β(s)cos 2

[−ψ(s) + ψ(s ′) − πν] = −1

2

d

dssin 2

[−ψ(s) + ψ(s ′) − πν]

is a total derivative, which indicates that a change in the order of integration mightbe helpful. The only problem is that the limits of the internal integral have an explicitdependence on s.We first split the double integral into two regions in the s, s ′ plane as

∫ C

0ds

∫ s+C

sds ′ =

∫ C

0ds

∫ C

sds ′ +

∫ C

0ds

∫ s+C

Cds ′ .

The integration area in the second integral on the right-hand side is then downshiftedby the ring circumference C in both variables, as shown Fig. 8.2, to yield

∫ C

0ds

∫ s+C

Cds ′ →

∫ 0

−Cds

∫ s

0ds ′ .

Page 113: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

106 8 Magnetic Field and Energy Errors

This transformation does not change the result because �K (s ′) and β(s ′) areperiodic functions with the period C , and the combination ψ(s) − ψ(s ′) does notchange when we simultaneously replace s ′ with s ′ − C and s with s − C (bothphases have offsets of 2πν that cancel each other out). It is then straightforward tosee that the resulting two integrals are equal to one double integral with the reversedorder of integration,

∫ C

0ds

∫ C

sds ′ +

∫ 0

−Cds

∫ s

0ds ′ =

∫ C

0ds ′

∫ s ′

s ′−Cds .

The internal integral can now be easily integrated,

�ν = 1

4π sin(2πν)

∫ C

0ds′�K (s′)β(s′)

(−1

2

) {sin 2

[−ψ(s) + ψ(s′) − πν] ∣∣∣

s′

s=s′−C

= 1

4π sin(2πν)

∫ C

0ds′�K (s′)β(s′)

(−1

2

)[sin(−2πν) − sin(2πν)] .

We can write this more simply as Eq. (8.34).Because the beta function correction is proportional to 1/ sin(2πν)which in theory

could be sensitive to small changes in ν, the result is often expressed in terms ofthe change in cos(2πν); this is proportional to sin(2πν)�β and may be a betterapproximation. The resulting expression is

� cos(2πν) = −1

2sin(2πν)

∫ C

0ds�K (s)β(s) .

Problem 8.5 Calculate the beta beat and the tune change for a localized perturba-tion of the lattice: �K = �K0δ(s − s0).

Solution: We use the formula (8.34) from Problem 8.4 for calculating the tunechange from an arbitrary quadrupole error. Substituting in �K = �K0δ(s − s0)yields

�ν = 1

4π�K0β(s0) .

We calculate the beta beat for the same example by using Eq. (8.32). The betatronfunction becomes

β1(s) = β(s) − β(s)

2 sin(2πν)�K0(s0)β(s0) cos 2 [−ψ(s) + ψ(s0) − πν] .

The remaining dependence on lattice location s shows that the beta beating termoscillates twice as fast as the betatron motion itself.

Page 114: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

Chapter 9Nonlinear Resonance and ResonanceOverlapping

Higher-order components in the magnetic field in a ring introduce nonlinear termsinto theHamiltonian and generate nonlinear resonances. This can lead to complicatedmotion for particles with large amplitudes of betatron oscillations. We derive theresonant structure in the phase space due to a sextupole magnet when the fractionalpart of the tune is close to± 1

3 . For a Hamiltonian systemwith many resonances, theycan interactwith each other and lead to stochastic orbits in phase space. To understandthis effect, we study a model called the standard map, that illustrates qualitativefeatures of what can occur in a Hamiltonian system with many resonances. Theimpact on dynamics is similarwhether originating fromeffects as diverse as nonlinearmagnetic fields, RF cavities, space-charge forces among the charged particles in abunch, or interactions between bunches.

9.1 The Third-Order Resonance

We will now study the effect of sextupoles on betatron oscillations. The sextupolevector potential is given by Eq. (6.9). We will limit our analysis to one-dimensionalbetatron oscillations in the x direction, set y = 0, and use As = −S(s)x3/6 for thevector potential. Recalling that, in the lowest order, the vector potential enters theHamiltonian in the combination −eAs/p0 (see Eq. (5.23)) we need to add it to thegeneric Hamiltonian (6.14),

H = 1

2P2x + 1

2K (s)x2 + 1

6S(s)x3 , (9.1)

where S = eS/p0. In what follows, we will assume that the last term on the right-hand side is small compared to the first two terms and treat it as a perturbation.

We first make a transformation to the action-angle variables J1 and φ1 definedin Sect. 7.2 (in what follows we will drop the subscript 1 to simplify the notation).

© Springer International Publishing AG, part of Springer Nature 2018G. Stupakov and G. Penn, Classical Mechanics and Electromagnetismin Accelerator Physics, Graduate Texts in Physics,https://doi.org/10.1007/978-3-319-90188-6_9

107

Page 115: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

108 9 Nonlinear Resonance and Resonance Overlapping

This transformation converts the first two terms of the Hamiltonian (9.1) into a linearfunction of J :

1

2P2x + 1

2K (s)x2 → 2πν

CJ . (9.2)

Transforming the last, nonlinear term we obtain:

H = 2πν

CJ +

√2

3J 3/2S(s)β3/2(s) cos3

[φ − 2πνs

C+ ψ(s)

]. (9.3)

It is convenient to change the scale of the independent variable s replacing it bythe angle θ = 2πs/C , so that θ increases by 2π every revolution in the ring. It iseasy to check that the new Hamiltonian which incorporates the rescaling is obtainedfrom (9.3) through multiplication by (C/2π),

H = ν J + V (φ, J, θ) , (9.4)

where the perturbation V is

V (φ, J, θ) =√2

3

C

2πJ3/2S(θ)β3/2(θ) cos3[φ − νθ + ψ(θ)] (9.5)

= C

12π√2J3/2S(θ)β3/2(θ) {cos 3 [φ − νθ + ψ(θ)] + 3 cos [φ − νθ + ψ(θ)]} .

The new Hamiltonian is periodic in θ with period 2π. The equations of motion forthe action-angle variables are:

∂ J

∂θ= −∂H

∂φ= −∂V

∂φ,

∂φ

∂θ= ∂H

∂ J= ν + ∂V

∂ J. (9.6)

Let us now analyze the relative role of the two terms on the right-hand side of (9.5).Note that if we neglect the perturbation V in the Hamiltonian (9.4), the phase evolvesas a linear function of θ, φ = νθ + φ0. After one turn, the combination νθ − ψ(θ)does not change, and hence the argument φ − νθ + ψ(θ) in (9.5) changes by 2πν.If the fractional part of ν is close to one-third or two-thirds, ν ≈ n ± 1/3, wheren is an integer, cos 3[φ − νθ + ψ(θ)] returns to approximately to same value afterθ changes by 2π, and the effect of this part of the perturbation accumulates witheach subsequent period leading to relatively large excursions in J on the orbit. Onthe contrary, cos[φ − νθ + ψ(θ)] has a phase jumping by ≈ ±2π/3 after each turn,and due to continuous change of the sign of the cos function the effect of this termaverages out almost to zero over many revolutions in the ring. This term would be

Page 116: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

9.1 The Third-Order Resonance 109

resonant for the tune close to an integer but, as we know, the integer values of thetune are already unstable.

In the rest of this section we will focus on the most interesting case when ν ≈n ± 1/3 and drop the cos[φ − νθ + ψ(θ)] term in the perturbation,

V (φ, J, θ) = C

12π√2J 3/2S(θ)β3/2(θ) cos 3[φ − νθ + ψ(θ)] . (9.7)

To further simplify our analysis let us consider a ring with one sextupole magnetof length much shorter than the ring circumference C .1 Without loss of generalitywe can assume that the magnet is located at θ = 0. For such a magnet, S(θ) can beapproximated by a periodic delta function,

S(θ) = S0δ(θ) , (9.8)

where δ(θ) = ∑∞k=−∞ δ(θ + 2πk). The requirement of the periodicity of S(θ) fol-

lows from the fact that two values of θ that differ by 2π correspond to the sameposition in the ring. The term V can now be written as

V (φ, J, θ) = C

12π√2J 3/2S0β

3/2(θ)δ(θ) cos 3[φ − νθ + ψ(θ)]

= 1

3RJ 3/2δ(θ) cos 3[φ − νθ + ψ(θ)] , (9.9)

where we have introduced the notation R = CS0β3/20 /(4π

√2), with β0 = β(0). The

equations of motion (9.6) take the form

∂ J

∂θ= RJ 3/2δ(θ) sin 3[φ − νθ + ψ(θ)],

∂φ

∂θ= ν + 1

2RJ 1/2δ(θ) cos 3[φ − νθ + ψ(θ)] . (9.10)

Let us consider how J andφ evolve over one turn in the ring, when θ changes from0 to 2π, starting from θ = −0, that is right before the delta function kick.Without lossof generality, we can assume thatψ(0) = 0.We first need to integrate these equationsthrough the delta-function kick, that is from θ = −0 to θ = +0, for arbitrary initialvalues J1 and φ1. For this one integration, the equations are simplified:

∂J∂θ

= J 3/2δ(θ) sin 3φ ,

∂φ

∂θ= 1

2J 1/2δ(θ) cos 3φ , (9.11)

1To be more precise, the length of the magnet should be much shorter than the betatron wavelength,which typically is a fraction of the circumference C .

Page 117: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

110 9 Nonlinear Resonance and Resonance Overlapping

where we rescaled the action introducing J = R2 J , set θ = 0 everywhere exceptin the argument of the delta function, and discarded the constant ν term from thesecond part of Eq. (9.10) in comparison to the delta function. The initial conditionsfor these equations are J = J1 = R2 J1 and φ = φ1.

Going back to our starting point of Eq. (9.1), we see that when crossing θ = 0 (andthus s = 0) there is no direct contribution from the sextupole field to the evolution ofx , based on dx/ds = ∂H/∂Px . From the equation dPx/ds = −∂H/∂x , it followsthat a short magnet only gives a transverse kick of magnitude �Px = −CS0x2/4π.Unfortunately, when we switch to action-angle coordinates we obtain two sets ofcoupled equations for φ and J which are more challenging to solve, and once wediscard the cosφ term from the perturbation V it is no longer consistent to insist onusing the same value for �Px nor requiring �x = 0. So while our equations (9.11)are physically meaningful when the tune is close to an integer ±1/3, we have morework to do in order to evaluate the effect of the sextupole.

One step we can take analytically is find an integral of motion across the sextupoleusing the following formal transformation which takes advantage of the fact that thedelta function δ(θ) is a derivative of the step function h(θ),

dh

dθ= δ(θ) , (9.12)

where h(θ) is equal to 1 for θ > 0 and zero otherwise. Replacing the delta functionby this derivative and then noting that

∂θ= dh

∂h= δ(θ)

∂h, (9.13)

we replace Eq. (9.11) with

∂J∂h

= J 3/2 sin 3φ ,

∂φ

∂h= 1

2J 1/2 cos 3φ , (9.14)

where the independent variable h now changes from 0 to 1 when θ traverses thedelta-function. It is straightforward to check that Eq. (9.14) represents Hamilton’sequations of motion for the following Hamiltonian:

H(φ,J ) = 1

3J 3/2 cos 3φ , (9.15)

Note that the Hamiltonian would be proportional to x3 if we had not neglected thecosφ term. Since the Hamiltonian does not depend on the independent variable h, itis conserved, and its trajectories can be easily found from the equation H(φ,J ) =const. They are shown in Fig. 9.1. Remarkably, due to the rescaling of variables,

Page 118: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

9.1 The Third-Order Resonance 111

Fig. 9.1 Trajectories in thephase space φ − J for theHamiltonian (9.15). Note thesymmetry along the φ axiswith the period π/3

our newHamiltonian (9.15) is universal— it does not contain any original parametersof the problem, such as the strength of the sextupole or the circumference of the ring.

We have now recovered a method to calculate the action-angle variables at theend of the sextupole, which we will denote as J and φ. By numerically integrat-ing Eq. (9.14) from h = 0 to h = 1, we can generate the map that moves ourparticle through the sextupole magnet according to our approximate Hamiltonian:J = f (φ1,J1) and φ = g(φ1,J1). To get an entire full-turn map we then need toevaluate the equations of motion from the exit from the sextupole, θ = +0, throughthe whole ring, to the next entrance to the magnet, θ = 2π − 0. We will denote theaction-angle variables at θ = 2π − 0 asJ2 andφ2. The action remains unchanged un-til the sextupole is again crossed, and the phase linearly grows until by the sextupoleentrance it has increased by 2πν, so we obtain:

J2 = f (φ1, J1), φ2 = g(φ1,J1) + 2πν . (9.16)

It is now clear that every revolution in the ring repeats the transformation (9.16), andthe values Jn , φn at the start of then-th revolution are expressed through the valuesat the previous one,

Jn = f (φn−1,Jn−1), φn = g(φn−1,Jn−1) + 2πν . (9.17)

To illustrate the dynamics of this map, we iterated it 300 times starting fromdifferent initial conditions and numerically integrating Eq. (9.14) on each step. Eachpairφn ,Jn was converted to the original canonical variables x and Px using Eqs. (7.9)and (7.10) and the relation J = J /R2:

x R/√

β0 = √2J cosφ, Px R

√β0 = −√

2J sin φ , (9.18)

where for simplicity we have assumed α = 0. The result of this simulation whenthe fractional part [ν] of the tune is equal to 1

3 + 0.1 is shown in Fig. 9.2. Thehorizontal and vertical axes are the normalized dimensionless coordinate x R/

√β0

and the normalized angle Px R√

β0. We see that the orbits with large amplitudes of

Page 119: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

112 9 Nonlinear Resonance and Resonance Overlapping

Fig. 9.2 Orbits in phasespace for the case[ν] − 1/3 = 0.1. Particlesstarting from outside of thelargest, triangular-shapedorbit quickly leave thesystem

betatron oscillations get a triangular shape reflecting the three-fold symmetry of thesextupole field (see Fig. 6.1b). The largest orbit shown in Fig. 9.2 is a separatrix —orbits that start outside of it move away from the axis and leave the system in severaliterations.

An example of experimentally measured third-order resonance orbits at the IUCFcooler ring (at the Indiana University Bloomington campus) can be found in Ref. [1].

9.2 Standard Model and Resonance Overlapping

As we saw in the previous section, the effect of a sextupole on betatron oscillationscan be reduced to a map, Eq. (9.17), which demonstrates particle confinement nearthe axis and particle losses outside of the separatrix. Many other nonlinear beamdynamics phenomena can also be formulated in terms of Hamiltonian maps. In thissection, wewill consider one suchmap often called the standard, or Chirikov-Taylor,map. The remarkable feature of this map is that it demonstrates a transition fromregular to chaotic motion in a non-integrable, time-dependent Hamiltonian systemwith only one degree of freedom.

The standard map describes the evolution in time of a system with the followingHamiltonian:

H(θ, I, t) = 1

2I 2 + K δ(t) cos θ , (9.19)

where K is a parameter, δ(t) = ∑∞n=−∞ δ(t + n) is the periodic δ function that

describes kicks repeating with the unit period (note that this definition of δ(t) isdifferent from the one defined in Sect. 9.1 where the period was equal to 2π). Here Icanbe considered as an action, and θ as an angle variable; they are both dimensionless.

Page 120: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

9.2 Standard Model and Resonance Overlapping 113

The equations of motion for I and θ are

I = −∂H

∂θ= K δ(t) sin θ , θ = ∂H

∂ I= I . (9.20)

If In and θn are the values at t = n − 0 (before the delta-function kick), then integrat-ing the first of Eq. (9.20) from t = n − 0 to t = n + 0 (through the delta-functionkick) gives In+1 = In + K sin θn , which is then conserved over the rest of the unitinterval from t = n + 0 to t = (n + 1) − 0 (where there are no kicks). Integratingthe second equation in (9.20) along this part of the interval and remembering that theaction here is already equal to In+1 gives θn+1 = θn + In+1. Combining these twoequations we obtain

In+1 = In + K sin θn ,

θn+1 = θn + In+1 . (9.21)

These equations transform the action-angle variables from their values at time t = nto time t = n + 1. This transformation is called the standard map.2

The periodic delta-function used in Eq. (9.19) can be expanded as a Fourier series,

δ(t) = 1 + 2∞∑n=1

cos (2πnt) . (9.22)

Substituting this representation into the Hamiltonian (9.19) we can rewrite the latterin the following form:

H(θ, I, t) = 1

2I 2 + K cos(θ) + 2K cos(θ)

∞∑n=1

cos(2πnt)

= 1

2I 2 + K cos(θ) + K

∞∑n=−∞n �=0

cos(θ − 2πnt) , (9.23)

where we have used the relation 2 cos(θ) cos(2πnt) = cos(θ − 2πnt) + cos(θ +2πnt). The two first terms on the right-hand side comprise the Hamiltonian of thependulum3 and the sum over terms with n �= 0 is a time-dependent periodic driverwith the frequencies equal to 2πn.

We can get some insight into the structure of the phase space by selecting onlyone term n in the infinite sum over all n (similar to what we did analyzing theHamiltonian (9.4)):

2One can also find in the literature a definition of the standard map which differs from Eq. (9.21)by numerical factors.3This can be seen if one sets ml2 → 1 in the pendulum Hamiltonian (1.32) and associates K with−ω2

0 . In case of positive K , one also needs to shift the origin of the angle, θ → θ + π.

Page 121: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

114 9 Nonlinear Resonance and Resonance Overlapping

Fig. 9.3 Sketch illustratingthe superposition of thephase portraits for variousresonances of the standardmap. The width of eachresonance is 4

√|K |

H(θ, I, t) = 1

2I 2 + K cos(θ − 2πnt) . (9.24)

Making the canonical transformation θ, I → φ, J where

J = I − 2πn , φ = θ − 2πnt , (9.25)

we eliminate the time variable and find the new Hamiltonian,

H′(φ, J ) = 1

2J 2 + K cosφ , (9.26)

which is again the Hamiltonian of a pendulum. The structure of the phase space forthis Hamiltonian is shown in Fig. 4.4; the separatrix that encompasses trajectorieswith a limited angular variation is defined by the equationH′(φ, J ) = |K |. From thisequation we find that the maximum deviation of J on the separatrix is J = ±2

√|K |.For the original action variable I this translates into the deviation relative to the value2πn, I = 2πn ± 2

√|K |.Trying to understand the overall structure of the phase space of the original Hamil-

tonian, we can naively superimpose the phase portraits for the Hamiltonians (9.26)with different values of n. This makes a cartoon shown in Fig. 9.3. Such superposi-tion makes sense only if |K | � 1, when the resonances for different values of n arewell separated and, in the first approximation, do not interact with each other.

Computer simulations of the standard map show that, indeed, as long as the dis-tance between the islands is much larger than the width of the separatrix, then to agood approximation resonanceswith different values of n can be considered separate-ly. However, increasing |K | leads to an overlapping of the resonances and the motionbecomesmuchmore complicated. Because the distance between the resonances is 2πand the width is 4

√|K |, formally the overlapping occurs for |K | > π2/4, but given

Page 122: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

9.3 Dynamic Aperture in Accelerators 115

the qualitative nature of our argument one should not expect a drastic transition at thisexact value of K . Indeed, simulations show that when |K | increases, there is a grad-ual transformation of the flow shown in Fig. 9.3 into a regime in which the laminarorbits are destroyed and themotion becomes stochastic. This is illustrated by Fig. 9.4.Qualitatively, the transition from the laminar to stochastic motion occurs at

K ∼ 1 . (9.27)

When |K | becomesmuch larger than one, regular orbits are destroyed and the particlemotion becomes chaotic. In this limit, after each kick the particle loses memory of itsprevious phase, and the consecutive phases θn can be considered to be uncorrelated.As a result, the subsequent values of action In can be described as a random walk,and over many steps a statistical description of the process as a diffusion along theI axis becomes appropriate.

We can easily estimate the rate of diffusion for the action. From Eq. (9.21) wecalculate the change of the action in one step, �In = K sin θn , and taking the squareof �In and averaging it over the random phase θn we obtain

〈�I 2n 〉 = 1

2K 2 . (9.28)

In a random walk, the average squares 〈�I 2n 〉 add up, and after N steps the averagesquare of the accumulated action IN is

〈I 2N 〉 ≈ 〈�I 2n 〉N ≈ 1

2K 2N . (9.29)

The linear growth of 〈I 2N 〉 with the number of steps is a characteristic feature of thediffusion process.

Figure 9.5 shows the result of a numerical simulation of the map (9.21) for aparticular value K = 8.41. The black line shows one sample orbit that demonstrateswild fluctuations of I 2N on top of a systematic growth with N . The red line is 〈I 2N 〉averaged over an ensemble of 400 orbits starting from different initial conditions —it agrees well with the analytical formula (9.29).4

9.3 Dynamic Aperture in Accelerators

We have seen in this chapter and the previous one that the severity of field errors andnonlinearities depends strongly on the tune. Nonlinear fields (both external and self-

4A more detailed study [2] reveals that Eq. (9.29) is only the leading term in the formula for 〈I 2N 〉in the limit K � 1. The correction to (9.29) is noticeable when K � 10, however it happens tovanish for the particular value of K = 8.41. This explains the very good agreement between thesimulation and theory in Fig. 9.5.

Page 123: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

116 9 Nonlinear Resonance and Resonance Overlapping

Fig. 9.4 The result of computer simulations for the standard map. Shown is a part of the phasespace bounded in the vertical direction by the inequality −π < I < π, for four different values ofthe parameter K , K = 0.1, 0.3, 1, 5 from left to right and from top to bottom. The last two picturesshow an increase and then an almost complete domination of the stochastic component of themotionin a large part of the phase space

Fig. 9.5 The calculatedensemble average of 〈I 2N 〉versus the number ofiteration N for K = 8.41(red line). The straight blueline is the theoreticalprediction of Eq. (9.29). Theblack line shows oneparticular orbit that startsfrom I = π, θ = π/10

Page 124: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

9.3 Dynamic Aperture in Accelerators 117

forces) have the effect of varying the tune corresponding to different particle orbits,leading to a tune spread. This means that even if a particle following the reference or-bit is not near a resonance, other particles at higher amplitudes may be strongly per-turbed by the nonlinear fields. We have not discussed coupling between different de-grees of freedom, but that leads to more opportunities for resonances to appear (forexample, if the vertical and horizontal tunes differ by 1/2). In a typical situation, non-linear fields make the phase space at some distance from the reference orbit prone tostochastic motion, leading to a random walk of the particle until it is lost.

At best there can only be a limited region near the reference orbit where particlesare properly confined. This region in phase space is called the dynamic apertureof the machine. It is computed with the help of accelerator codes by launchingparticles at various locations away from the reference orbit and tracking their motion.Rather than having a sharp boundary, the dynamic aperture is usually surroundedby an intermediate zone where the rate of diffusion which particles experience getsgradually worse. A related concept that focuses on the short-term tune of particleswithin a bunch is the analysis of frequency maps [3].

A modern circular accelerator has many magnets that play various roles in con-fining the beam in the ring. Even as industry and researchers learn to reduce errors inthe manufacture and installation of magnets, more aggressive designs and improve-ments in other areas tend to make nonlinearities a major constraint on the operationof accelerators, limiting the total charge contained in storage rings and the luminosityof colliders.

Worked Examples

Problem 9.1 Prove that the standard map defines a canonical transformation(In, θn) → (In+1, θn+1).

Solution: We can see that {In+1, In+1} = 0 and {θn+1, θn+1} = 0 trivially, since{ f, f } always vanishes for the Poisson bracket. Then with

In+1 = In + K sin θn ,

θn+1 = θn + In + K sin θn ,

we find

{θn+1, In+1} = ∂θn+1

∂θn

∂ In+1

∂ In− ∂θn+1

∂ In

∂ In+1

∂θn

= (1 + K cos θn) − K cos θn = 1 .

Therefore the map is canonical.

Problem 9.2 Prove the following property of the standard map: for two trajectoriesstarting from the same initial value θ0 but with different values I

(1)0 and I (2)

0 , such thatI (2)0 − I (1)

0 = 2πm, where m is an integer, the difference I (2)n − I (1)

n remains equal to2πm for all values of n.

Page 125: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

118 9 Nonlinear Resonance and Resonance Overlapping

Solution: We will do a proof by induction. First we assume it holds true forsome value of n: I (2)

n − I (1)n = 2πm. While the orbits start from the same value of

angle θ0, we assume that on step n their angles differ by an integer number of 2π,θ(2)n = θ(1)

n + 2π p, where p can be any integer. Then

I (2)n+1 − I (1)

n+1 = (I (2)n − I (1)

n ) + K (sin θ(2)n − sin θ(1)

n )

= 2πm ,

θ(1)n+1 − θ(2)

n+1 = (θ(2)n − θ(1)

n ) + (I (2)n − I (1)

n ) + K (sin θ(2)n − sin θ(1)

n )

= 2π(m + p) .

We are given that it holds for n = 0, so it must hold for all n.

Problem 9.3 Prove that Eq. (9.25) defines a canonical transformation, find the cor-responding generating function F2, and obtain the Hamiltonian (9.26).

Solution: First we show the transformation is canonical:

{φ, J } = ∂ J

∂ I

∂φ

∂θ− ∂ J

∂θ

∂φ

∂ I= 1 .

Then we can find the generating function F2(θ, J, t) by integrating the expressionsfor φ and I in terms of the partial derivatives of F2:

φ = ∂F2

∂ J= θ − 2πnt =⇒ F2 = J (θ − 2πnt) + g(θ, t) ,

I = ∂F2

∂θ= J + 2πn =⇒ F2 = θ(J + 2πn) + f (J, t) ,

where g and f are arbitrary functions of their arguments. Combining these twoequations for F2 we conclude that

F2 = θJ + 2πnθ − 2πnJ t + h(t) ,

where h is an arbitrary function of time. The quantity h(t) can be ignored because itdoes not couple to the dynamic variables. Finally, we can find the new Hamiltonianfrom

H ′ = H + ∂F2

∂t

= I 2

2+ K cos(θ − 2πnt) − 2πnJ

= (J + 2πn)2

2+ K cos(φ) − 2πnJ

= J 2

2+ K cos(φ) + 2π2n2 ,

Page 126: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

9.3 Dynamic Aperture in Accelerators 119

as given. Again, the constant term or any purely time-dependent term does not affectthe dynamics and can be ignored.

Problem 9.4 For Eq. (9.21), the main fixed points (ignoring periodicity) when K >

0 are an (unstable) saddle point at θ = 0, I = 0 and, for small K , a stable fixed pointat θ = π, I = 0. Find the value of K > 0 above which the motion around θ = π,I = 0 also becomes unstable.

Solution: Becausewe are expanding around θ = π, it is convenient to define a shiftedvariable δ = θ − π. Then the map becomes In+1 = In − K sin δn , δn+1 = δn + In+1.Expanding for δ, I � 1 and only keeping the linear term, sin δ ≈ δ, we can writethe linear single-turn map in matrix form as

(δn+1

In+1

)=

(1 − K 1−K 1

) (δnIn

).

This matrix, obtained by linearizing the single-turn map, is identical to the Jacobianmatrix for thismap as described inChap. 3, but specifically at the fixed point (note thatthe determinant is 1). We can evaluate the local stability by looking for eigenvaluesof the matrix:

0 =∣∣∣∣1 − K − λ 1

−K 1 − λ

∣∣∣∣ = λ2 + λ(K − 2) + 1 .

Solving this equation we find

λ = 1

2

(2 − K ±

√K 2 − 4K

).

Whenever there is some |λ| > 1, the motion is unstable because there is some com-bination of small offsets which will grow exponentially each turn. A simple analysisshows that |λ| > 1 for K > 4. The loss of the stable point at θ = π and I = 0 canbe clearly seen in Fig. 9.4 comparing the phase space at K = 1 and K = 5.

References

1. D.D.Caussyn,M.Ball, B. Brabson, J. Collins, S.A. Curtis, V.Derenchuck,D.DuPlantis, G. East,M. Ellison, T. Ellison, D. Friesel, B. Hamilton, W.P. Jones, W. Lamble, S.Y. Lee, D. Li, M.G.Minty, T. Sloan, G. Xu, A.W. Chao, K.Y. Ng, S. Tepikian, Experimental studies of nonlinearbeam dynamics. Phys. Rev. A 46(12), 7942–7952 (1992)

2. A.B. Rechester, R.B. White, Calculation of turbulent diffusion for the Chirikov-Taylor model.Phys. Rev. Lett. 44, 1586–1589 (1980)

3. D.Robin,C. Steier, J. Laskar, L.Nadolski,Global dynamics of the advanced light source revealedthrough experimental frequency map analysis. Phys. Rev. Lett. 85, 558 (2000)

Page 127: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

Chapter 10The Kinetic Equation

In the preceding chapters we focused our attention on the motion of a single particle.In this chapter, wewill introduce the concept of the distribution function and describethe formalism of the kinetic equation for treating large ensembles of particles ina beam. While this chapter focuses on deterministic Hamiltonian motion, kineticequations in general can also include stochastic motion and damping.

10.1 The Distribution Function in Phase Spaceand the Kinetic Equation

We begin from a simple case of one degree of freedom when each particle is charac-terized by two canonically conjugate variables q and p. A large ensemble of particlesthat represents a beam shall contain various pairs of values of q and p and constitutesa “cloud” in the phase space as illustrated in Fig. 10.1. With time, each particle willmove along its own orbit in the physical space, and the corresponding phase pointtravels along a trajectory in the phase space. The “cloud” gradually changes shape.The particle motion is governed by external fields, as well as interactions betweenthe particles. In this chapter, however, we neglect the interaction effects, and assumethat each particle moves independently due to external electromagnetic fields only.

Consider an infinitesimally small region dq × dp in the phase plane with thecenter located at p, q, as shown in Fig. 10.2, and let the number of particles at timet in this phase space element be given by dN . This mathematically infinitesimalphase element should be considered to be physically large enough to include manyparticles, so that dN � 1. We define the distribution function of the beam f (q, p, t)so that

dN (t) = f (q, p, t)dp dq . (10.1)

© Springer International Publishing AG, part of Springer Nature 2018G. Stupakov and G. Penn, Classical Mechanics and Electromagnetismin Accelerator Physics, Graduate Texts in Physics,https://doi.org/10.1007/978-3-319-90188-6_10

121

Page 128: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

122 10 The Kinetic Equation

Fig. 10.1 Illustration of thephase space of an ensembleof particles with the positionof each particle indicated bya red point

Fig. 10.2 A characterizationof the distribution functionf . The red dots showparticles inside aninfinitesimally smallrectangle with the sides dqand dp

We can say that the distribution function gives the density of particles in the phasespace.

As was emphasized above, the phase points are moving along trajectories, andthe distribution function evolves with time. Our goal is to derive an equation thatgoverns this evolution.

At time t + dt the number of particles in the region dq × dp will change becauseof the flowof particles through the four boundaries of the rectangle. Due to themotionin the q direction, the number of particles that flow in through the left boundary is

f

(q − 1

2dq, p, t

)× dp × q

(q − 1

2dq, p, t

)dt . (10.2)

In this equation, qdt is the distance from which the flow brings new particles intothe rectangle during time dt , and we take the values of both f and q in the middle ofthe left side of the rectangle. Similarly, the number of particles that flow out throughthe right boundary is

f

(q + 1

2dq, p, t

)× dp × q

(q + 1

2dq, p, t

)dt . (10.3)

Page 129: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

10.1 The Distribution Function in Phase Space and the Kinetic Equation 123

Using the same logic, we calculate the number of particles which flow in throughthe lower horizontal boundary,

f

(q, p − 1

2dp, t

)× dq × p

(q, p − 1

2dp, t

)dt , (10.4)

and the number of particles that flow out through the upper horizontal boundary,1

f

(q, p + 1

2dp, t

)× dq × p

(q, p + 1

2dp, t

)dt . (10.5)

We are now ready to calculate the change of the number of particles in the phasevolume dq × dp. On one hand, this number is due to the change of the distribu-tion function during the time interval dt , dN (t + dt) − dN (t) = [ f (q, p, t + dt) −f (q, p, t)]dp dq. On the other hand, it is equal to the sum of the four contributionscalculated above. Equating these two expressions we obtain,

[ f (q, p, t + dt) − f (q, p, t)]dp dq

= f

(q − 1

2dq, p, t

)q

(q − 1

2dq, p, t

)dp dt − f

(q + 1

2dq, p, t

)q

(q + 1

2dq, p, t

)dp dt

+ f

(q, p − 1

2dp, t

)p

(q, p − 1

2dp, t

)dq dt − f

(q, p + 1

2dp, t

)p

(q, p + 1

2dp, t

)dq dt .

(10.6)

Expanding both sides of this equation in the Taylor series and keeping only linearterms in dp, dq, dt , and then dividing both sides by dp dq dt , we arrive at thefollowing result:

∂ f (q, p, t)

∂t+ ∂

∂q[q(q, p, t) f (q, p, t)] + ∂

∂ p[ p(q, p, t) f (q, p, t)] = 0 .

(10.7)

This is the continuity equation for the function f — it is a mathematical expressionof the fact that the particles in the phase space are not created and do not disappear;they are being transported from one place to another along smooth paths. Integratingthis equation over the whole phase space and taking into account that f vanishesat infinity when |q|, |p| → ∞, gives dN/dt = 0, where N = ∫

f dq dp is the totalnumber of particles in the system. As expected, the total number of particles isconserved.

In the derivation above, we did not use the Hamiltonian nature of the phase flow.We will now show that, for a Hamiltonian system, the medium represented by thedistribution function f is incompressible. This follows immediately from Liouville’s

1Our classification of “flow in” and “flow out” tacitly assumes that the corresponding velocities qand p are positive. For negative values of these velocities, the values of the “flow in” and “flow out”will be negative, corresponding to an actual outflow and inflow respectively.

Page 130: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

124 10 The Kinetic Equation

theorem (see Sect. 3.4) that states that the volume dp dq of a space phase elementdoes not change in Hamiltonian motion. Since f is proportional to the number ofparticles in this volume, and this number is conserved, f too is conserved, but onlywithin a moving phase space volume element. The density at a given point q, p ofthe phase space does, however, change because a fluid element located at this pointat a given time will be replaced by a new element at a later time.

Mathematically, the fact of incompressibility is reflected in the following trans-formation of the continuity equation (10.7). From the Hamiltonian equations for qand p it follows that

∂qq = ∂

∂q

∂H

∂ p,

∂ pp = − ∂

∂ p

∂H

∂q, (10.8)

and hence ∂q/∂q + ∂ p/∂ p = 0. Using this relation, we can rewrite Eq. (10.7) asfollows:

∂ f

∂t+ ∂H

∂ p

∂ f

∂q− ∂H

∂q

∂ f

∂ p= 0 , (10.9)

where, for brevity, we have dropped the arguments of f and H . In accelerator andplasma physics this version of the kinetic equation is often called the Vlasov equa-tion. Specifically, there are no scattering or damping terms. It provides an extreme-ly powerful tool in accelerator physics giving a detailed description of the beamdynamics.

Note that we can also use the formalism of the Poisson bracket to write the Vlasovequation as

∂ f

∂t= {H, f } . (10.10)

In this form, the Vlasov equation is also valid for n degrees of freedom, with canon-ical variables qi and pi , i = 1, 2, . . . , n, and the distribution function f definedas a density in 2n-dimensional phase space and depending on all these variables,f (q1, . . . , qn, p1, . . . , pn, t). Equivalently, the Vlasov equation can be written as

∂ f

∂t=

n∑i=1

(∂H

∂qi

∂ f

∂ pi− ∂H

∂ pi

∂ f

∂qi

). (10.11)

To conclude this section, we note that in some cases it is more convenient tonormalize f by the number of particles N ; in this case, the integral of f over the phasespace is equal to one. With such a normalization, f (q, p, t)dq dp can be understoodas a probability to find a particle in the phase volume dq dp in the vicinity of thephase point q, p.

Page 131: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

10.2 Integration of the Vlasov Equation Along Trajectories 125

Fig. 10.3 A trajectory in theextended phase spaceconnecting two neighboringpoints

10.2 Integration of the Vlasov Equation Along Trajectories

In this sectionwewill prove by direct calculation that any distribution functionwhichsatisfies the Vlasov equation (10.11) remains constant in each “fluid” element of thephase space as it moves along a particle trajectory. This property of the distributionfunction immediately follows from the Liouville theorem; because the phase volumeof a small fluid element is conserved, the value of the distribution function equal tothe ratio of the number of particles in this element to its volume does not change.In addition to confirming this important property, we will derive from it a powerfulmethod for solving theVlasov equation. For simplicitywewill limit our considerationto Hamiltonian systems with one degree of freedom.

Let us consider a trajectory in the extended phase space that in addition to q andp axes includes the time axis t , as shown in Fig. 10.3. We want to calculate thedifference of f at two close points, at time t and t + dt , on this single trajectory. Wehave

d f = f (q + dq, p + dp, t + dt) − f (q, p, t)

= ∂ f

∂tdt + ∂ f

∂qdq + ∂ f

∂ pdp . (10.12)

Because the two points are on the same trajectory, dq = qdt = (∂H/∂ p) dt anddp = pdt = −(∂H/∂q) dt . Substituting these relations into (10.12) we find that

d f = ∂ f

∂tdt − ∂H

∂q

∂ f

∂ pdt + ∂H

∂ p

∂ f

∂qdt = 0 , (10.13)

or

d f

dt= 0 . (10.14)

On the last step we invoked Eq. (10.9). Equation (10.14) is the mathematical expres-sion of the fact that f remains constant along a trajectory. This derivative, which

Page 132: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

126 10 The Kinetic Equation

describes changes along a trajectory, is referred to as the convective derivative, andcan be written as

d

dt conv≡ ∂

∂t+ q

∂q+ p

∂ p. (10.15)

We have encountered this derivative in Sect. 1.5.Knowing that f is constant along trajectories, we can find solutions to the Vlasov

equation if the phase space orbits are known. Let q(q0, p0, t) and p(q0, p0, t) besolutions of the Hamiltonian equations with initial values q0 and p0 at t = 0, andF(q0, p0) be the initial distribution function at t = 0. To find the value of f at q, p attime t we need to trace back the trajectory that passes through q, p at t , and find theinitial values q0, p0 where it starts at t = 0. Hence, we need to invert the relations

q = q(q0, p0, t) , p = p(q0, p0, t) , (10.16)

and find q0, p0 in terms of q, p: q0 = q0(q, p, t) and p0 = p0(q, p, t). The value off at q, p at time t is then equal to the value of F at q0, p0:

f (q, p, t) = F(q0(q, p, t), p0(q, p, t)) . (10.17)

For simple trajectories, the inversion of (10.16) can be done analytically, andEq. (10.17) then defines f for an arbitrary initial function F .

As an illustration of the method, let us consider an ensemble of linear oscillatorswith frequency ω, whose motion is described by the Hamiltonian

H(x, p) = p2

2+ ω2 x

2

2. (10.18)

The distribution function f (x, p, t) for these oscillators satisfies theVlasov equation,

∂ f

∂t+ p

∂ f

∂x− ω2x

∂ f

∂ p= 0 . (10.19)

Solving the Hamiltonian equations, it is easy to find the trajectory which has initialvalue x0 and p0 at t = 0,

x = x0 cos(ωt) + p0ω

sin(ωt),

p = −ωx0 sin(ωt) + p0 cos(ωt) . (10.20)

Inverting these equations, we find

x0 = x cos(ωt) − p

ωsin(ωt),

p0 = ωx sin(ωt) + p cos(ωt) . (10.21)

Page 133: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

10.2 Integration of the Vlasov Equation Along Trajectories 127

Fig. 10.4 Plots of the distribution function f at three distinct moments in time: ωt = 0, ωt = π/2and ωt = π. A bell-shaped distribution function is initially shifted along the x-axis

If F(x, p) is the initial distribution function at t = 0, then, according to Eq. (10.17)we have

f (x, p, t) = F(x cos(ωt) − p

ωsin(ωt),ωx sin(ωt) + p cos(ωt)

). (10.22)

This solution describes a rotation of the distribution function in the phase spacewhich is illustrated by Fig. 10.4. The figure shows three consecutive positions of abell-shaped function f with an initial offset in the x direction.

10.3 Action-Angle Variables in the Vlasov Equation

The Vlasov equation (10.9) has the same form independent of the choice of thecanonical variables q and p. For a particular problem, a judicious choice of thesevariables cangreatly simplifyfinding a solution to theVlasov equation. In this section,we will demonstrate the advantages of using in the Vlasov equation the action-anglevariables φ, J introduced for 1D systems in Sect. 3.2.

Consider a 1D systemwith the action-angle variablesφ, J and a time-independentHamiltonian H(J ). For a distribution function f that depends on φ, J , and t , f =f (φ, J, t), the Vlasov equation reads,

∂ f

∂t+ ∂H

∂ J

∂ f

∂φ− ∂H

∂φ

∂ f

∂ J= ∂ f

∂t+ ∂H

∂ J

∂ f

∂φ= 0 , (10.23)

where we have used the fact that H does not depend on φ. Noting that accordingto Eqs. (3.20) and (3.23) the derivative ∂H/∂ J is equal to the revolution frequencyω(J ) along the orbit with action J , we rewrite (10.23) as

∂ f

∂t+ ω(J )

∂ f

∂φ= 0 . (10.24)

This equation is satisfied by an arbitrary function f of the following form:

f (φ, J, t) = F(φ − ω(J )t, J ) , (10.25)

Page 134: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

128 10 The Kinetic Equation

which is easily verified by substituting f into (10.24). This result has a simplegeometrical meaning: the values of the distribution function on the orbit with a givenaction J rotate along this orbit with the angular frequency ω(J ). In general, thisis a differential rotation — different layers of the phase space rotate with differentfrequencies.

UsingEq. (10.25)we can find a general formof a steady-state distribution functionthat does not depend on time. Because ∂ f/∂t = −ω(J )∂F/∂φ, from ∂ f/∂t = 0 isfollows that F does not depend on φ. We come to the conclusion that any functionf that depends only on J is a steady-state solution to the Vlasov equation.The particular form of the function f (J ) for a beam in an accelerator cannot be

found from Eq. (10.24) alone, and in practice is often determined by either initialconditions (how the beamwas generated or injected into an accelerator) or some slowdiffusion or collision processes in the ring. In many cases, a negative exponentialdependence of f versus J is a good approximation to the equilibrium beam state,

f = const e−J/ε0 = const exp

(− 1

2βε0

[x2 + (βPx + αx)2

]), (10.26)

wherewehaveused the expression (7.8) for J in a linearmagnetic lattice.Thequantityε0 is called the beam emittance. It is an important characteristic of the beam quality.

10.4 Phase Mixing

FromEq. (10.25) we can draw some important conclusions about the evolution of thedistribution function in the limit t → ∞. Because φ is an angular variable, two val-ues of φ that differ by 2π correspond to the same point in phase space. Hence F is aperiodic function of φ with period 2π and can be expanded into the Fourier series

F(φ, J ) =∞∑

n=−∞Fn(J )einφ , (10.27)

where

Fn(J ) = 1

∫ 2π

0F(φ, J )e−inφ dφ . (10.28)

Using this representation of F we can rewrite Eq. (10.25) as

f (φ, J, t) =∞∑

n=−∞Fn(J )ein[φ−ω(J )t] . (10.29)

Page 135: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

10.4 Phase Mixing 129

In the limit t → ∞ all terms in this sum, except for n = 0, become rapidly oscillatingfunctions of the action J due to the factor e−inω(J )t . When calculating any integral off over the phase space, the contribution of these terms averages out to almost zeroand becomes negligible. Hence in this limit we only have to keep the n = 0 term:

f (t,φ, J ) → F0(J ) ≡ 1

∫ 2π

0f (0,φ, J ) dφ . (10.30)

This is simply the average over the angle coordinate of the initial distribution func-tion f . This derivation naturally explains why the steady-state distribution functiondepends only on action — the fact established in the previous section.2

The mechanism that is responsible for the evolution of the distribution function toa steady state through rapid oscillations of the phase factor e−inω(J )t is called phasemixing. We can make a rough estimate of the time needed to approach equilibriumin this scenario. If we use �ω to characterize the frequency spread in the system dueto the function ω(J ) and the distribution of particles found in the beam, the phasevariation nω(J )t at time t can be estimated as n�ωt , and the phases on differentorbits start to diverge at t � π/n�ω. The longest time needed to mix the phasescorresponds to the n = 1 term, giving an estimate t � π/�ω. Hence, the distributionfunction reaches the steady state at times t � π/�ω.

10.5 Damping and Stochastic Motion

In previous chapters we have discussed how the amplitude of motion of a singleparticle can decrease due to damping, or take a random walk as a result of stochasticmotion. Here, we briefly describe how these effects are incorporated in the formalismof the kinetic equation.

The most straightforward way to see the impact of damping is to return to the con-tinuity equation and recalculate the convective derivative, this time with correctionsto the Hamiltonian dynamics from Eq. (3.38) due to non-conservative forces:

0 = ∂ f

∂t+

∑i

[∂

∂qi(qi f ) + ∂

∂ pi( pi f )

]

= ∂ f

∂t+

∑i

[qi

∂ f

∂qi+ pi

∂ f

∂ pi+ f

∂qi∂qi

+ f∂ pi∂ pi

]

= d f

dt conv+ f

∑i

∂Fi∂ pi

, (10.31)

2A linear oscillator in which ω is constant and does not depend on J is an exception: it does notexhibit phase mixing.

Page 136: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

130 10 The Kinetic Equation

thus we find that

d f

dt conv= − f

∑i

∂Fi∂ pi

. (10.32)

Here we have used the expression for the convective derivative of Eq. (10.15). Thisresult is consistent with the more general result Eq. (3.39) given the additional con-straint on the distribution function from the continuity equation. Using Eq. (3.46) wecan connect the evolution of the distribution function with the time derivative of thedeterminant of the Jacobian matrix M of the dynamic flow,

1

f

d f

dt conv= − 1

det M

d det M

dt. (10.33)

Integrating this equation over time, we find that f (t) evaluated along a particletrajectory scales in time as the inverse of the determinant of the matrix M(t):f (qi , pi , t)/ f (qi , pi , 0) = 1/[det M(t)]. This is consistent with the notion that thephase space density increases only when trajectories converge in phase space due tonon-Hamiltonian dynamics.

For one degree of freedom and F = −γ x = −γ p, the right-hand side ofEq. (10.32) becomes simply γ f , and

f (x(t), p(t), t) = f (x(0), p(0), 0) eγt . (10.34)

Random kicks with a small correlation time can also be incorporated into the for-malism of the distribution function in a natural way if the coordinates are chosen sothat only the momenta are directly impacted by the kicks. Because these kicks leadto a random walk of individual particles, this appears in the distribution functionas a diffusion process when a large ensemble is considered. Considering a singledegree of freedom and very short time scales, so that the dynamics have a negligibleimpact, uncorrelated random kicks with typical magnitude �p and a typical time�t between kicks lead to a random walk with

⟨[p(t) − p0]2⟩ = (t − t0)〈(�p)2/�t〉,

where p(t0) = p0 (cf. Eq. (4.15)). This process, convolved with the initial distribu-tion, leads to a spreading out of the distribution function. Statistically, it can also bedescribed as the result of a differential operator

∂ f

∂t= Ds

∂2 f

∂ p2, (10.35)

where Ds = 〈(�p)2/�t〉. Because this expression of the dynamics depends on in-finitesimal time scales, it is only necessary to add back the full dynamics by replacing∂ f/∂t with the convective derivative — for completeness we include the correctionfrom frictional forces:

Page 137: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

10.5 Damping and Stochastic Motion 131

d f

dt conv= − f

∂F

∂ p+ Ds

∂2 f

∂ p2. (10.36)

As seen in Sect. 4.3, the impact of this differential operator will be mixed with thatof the particle dynamics to yield a spread in both momentum and position, especiallywhen the frequency of motion is fast compared to the impact of the scattering. For aone-dimensional system and the simple form of damping F = −γ p, we can expandthis to find the partial time derivative:

∂ f

∂t= −∂H

∂ p

∂ f

∂x+ ∂H

∂q

∂ f

∂ p+ γ p

∂ f

∂ p+ γ f + Ds

∂2 f

∂ p2

= −{ f, H} + ∂

∂ p(γ p f ) + Ds

∂2 f

∂ p2. (10.37)

The second term in the final expression, related to damping, combines the effect ofthe distribution function having a convective derivative, as found above, with the factthat the flow in phase space is itself no longer fully defined by the Poisson bracket, asseen in Eq. (3.39). In some literature, Eq. (10.37) is called the Vlasov-Fokker-Planckequation.

Worked Examples

Problem 10.1 Write the Vlasov equation for a beam distribution f (x, Px , s) interms of variables x and Px . Use the Hamiltonian from Eq. (6.14).

Solution: The Vlasov equation is given by

−∂ f

∂s+ ∂H

∂x

∂ f

∂Px− ∂H

∂Px

∂ f

∂x= 0 .

Using our Hamiltonian (6.14) for an accelerator, H0 = P2x /2 + K (s)x2/2, we find

−∂ f

∂s+ K (s)x

∂ f

∂Px− Px

∂ f

∂x= 0 . (10.38)

Problem 10.2 Give a direct proof that the function (10.26) satisfies the Vlasovequation.

Solution: The partial derivatives of the distribution function (10.26) can bewritten as

Page 138: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

132 10 The Kinetic Equation

1

f

∂ f

∂x= − 1

ε0

(1 + α2

βx + αPx

),

1

f

∂ f

∂Px= − 1

ε0(βPx + αx) ,

1

f

∂ f

∂s= − 1

ε0

[− (1 + α2)β′

2β2x2 + αα′

βx2 + β′

2P2x + α′x Px

].

The partial derivative in s only acts on β andα = −β′/2. From the previous problem,the Vlasov equation (10.38) has 3 terms proportional to x2, P2

x , and x Px . The P2x

term is (−α − β′/2)P2x = 0 by the definition of α. The rest of the equation is then:

0 = x2[αK + (1 + α2)β′

2β2− αα′

β

]+ x Px

(βK − 1 + α2

β− α′

)

= x

β2(βPx + αx)

(Kβ2 − 1 − α2 − βα′) .

The last expression vanishes because that is the equation which defines the betafunction, see Eq. (6.23).

Problem 10.3 Show how the average of the squares of x and Px are expressedthrough ε0 and β for the distribution function of Eq. (10.26).

Solution: To find an average value of some quantity, we need to calculate thefollowing integrals:

〈. . .〉 =∫(. . .) f dx d Px∫

f dx d Px.

Webegin by calculating the normalization,∫

f dx d Px . It is simplest to do the integralover Px first:

∫f dx d Px =

∫dx e−x2/2βε0

∫dPx exp

[− 1

2βε0(βPx + αx)2

]

=∫

dx e−x2/2βε0

√2πε0

β= √

2πε0β

√2πε0

β= 2πε0 ,

where we used the fact that we could change Px to u = Px + αx/β to simplify theexponent and still have an integral over all u. To calculate the mean of x2, the integralover Px is unaffected:

Page 139: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

10.5 Damping and Stochastic Motion 133

⟨x2

⟩ = 1

2πε0

∫dx x2e−x2/2βε0

∫dPx exp

[− 1

2βε0(βPx + αx)2

]

= 1

2πε0

√2π (ε0β)3/2

√2πε0

β= βε0 .

We can again use the trick u = Px + αx/β to calculate the moment

⟨(βPx + αx)2

⟩= 1

2πε0

∫dx e−x2/2βε0

∫dPx (βPx + αx)2 exp

[− 1

2βε0(βPx + αx)2

]

= 1

2πε0

√2πε0β

√2πε0β = ε0β .

Finally, we can see from symmetry that the average of x(βPx + αx) has to vanishbecause again we can switch to the variable u whose integral is 0 by symmetry. Asa consequence,

⟨(βx Px + αx2

)⟩ = 0. Combining these expressions yields

⟨x2

⟩ = βε0 , 〈x Px 〉 = −αε0 ,⟨P2x

⟩ = 1 + α2

βε0 .

Page 140: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

Part IIElectricity and Magnetism

Page 141: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

Chapter 11Self Field of a Relativistic Beam

The electromagnetic field generated by a high-current, relativistic beam of chargedparticles plays an important role in beam dynamics. On the one hand, the forceexerted on the beam by this field can lead to beam instabilities and a deteriorationof its properties in the process of beam generation, acceleration and transport. Onthe other hand, this field induces currents and charges in the beam environment thatcan be used for diagnostic purposes. Hence calculation of this field and the forcesassociated with it constitutes an essential part of beam physics for accelerators.

In this chapter, we analyze the electromagnetic field of a relativistic beammovingwith a constant velocity in free space. In the following chapter we will look at theimpact of material surfaces around the beam path. The authors recommend books byJackson [1], byLandau andLifshitz [2], and byLandau andLifshitz and Pitaevskii [3]for an overview of the electromagnetic field and its interactions with matter.

11.1 Relativistic Field of a Particle Moving with ConstantVelocity

A beam consists of many charged particles, so it makes sense to begin with a reviewof the electromagnetic field of a point charge q moving with a relativistic velocity vin free space. The most straightforward way to derive this field is to use the Lorentztransformation starting from the particle rest frame where there is only the staticCoulomb field,

E′ = 1

4πε0

qr′

r ′3 , (11.1)

with r′ defined as the vector drawn from the position of the particle to the observationpoint and E′ being the electric field at the observation point. The magnetic field B′is zero in this frame. Here and below the prime denotes variables in the rest frame.

© Springer International Publishing AG, part of Springer Nature 2018G. Stupakov and G. Penn, Classical Mechanics and Electromagnetismin Accelerator Physics, Graduate Texts in Physics,https://doi.org/10.1007/978-3-319-90188-6_11

137

Page 142: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

138 11 Self Field of a Relativistic Beam

To find the electric and magnetic fields in the lab frame we will use the Lorentztransformation for the fields, Eq. (B.17). We assume that the particle is moving alongthe z-axis with its trajectory given by x = y = 0 and z = vt , so that the origin in therest frame (x ′ = y′ = z′ = 0) coincides with the particle. For cases like this wherethe magnetic field vanishes in the rest frame we have

Ex = γE ′x , Ey = γE ′

y , Ez = E ′z . (11.2)

We also need to transform the vector r′ in Eq. (11.1) into the lab frame usingEq. (B.2).For the length of this vector we obtain r ′ = √

x2 + y2 + γ2(z − vt)2, which givesfor the electric field E in the lab frame expressed in Cartesian coordinates:

Ex = 1

4πε0

qγx

[x2 + y2 + γ2(z − vt)2]3/2 ,

Ey = 1

4πε0

qγy

[x2 + y2 + γ2(z − vt)2]3/2 ,

Ez = 1

4πε0

qγ(z − vt)

[x2 + y2 + γ2(z − vt)2]3/2 . (11.3)

These three equations can be combined into a vectorial one:

E = 1

4πε0

qrγ2R3

. (11.4)

Here the vector r is drawn from the current position of the particle to the observationpoint, r = (x, y, z − vt), and R denotes the following expression:

R =√

(z − vt)2 + γ−2(x2 + y2) . (11.5)

As follows from Eq. (B.17), the moving charge also carries a magnetic field,

B = 1

c2v × E . (11.6)

The magnetic field points in the azimuthal direction, with the magnetic field linesencircling the z-axis.

The above result can also be obtained through the Lorentz transformation ofthe potentials. Indeed, in the particle rest frame we have the Coulomb electrostaticpotential φ′ and zero vector potential A′,

φ′ = 1

4πε0

q

r ′ , A′ = 0 . (11.7)

Page 143: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

11.1 Relativistic Field of a Particle Moving with Constant Velocity 139

Fig. 11.1 A sketch of the regions of intense electromagnetic fields around a particle moving alonga trajectory z = vt (left), and a plot of the magnitude of the electric field E (right, measured inunits q/4πε0ρ

2) for a given transverse offset ρ = √x2 + y2 and as a function of the polar angle

θ = arctan[ρ/(z − vt)] relative to the particle and its direction of motion, for several values of γ.The width of each curve scales as 1/γ. Both the electric and magnetic fields of a relativistic particleare mainly localized around the plane transverse to the velocity

Using the Lorentz transformation (B.18) we find their values in the lab frame,

φ = γφ′ , A = 1

cβφ . (11.8)

Expressing r ′ in terms of the coordinates in the lab frame, r ′ = γR, gives

φ = 1

4πε0

q

R , A = Z0

4πβq

R , (11.9)

where Z0 = 1/(ε0c) = μ0c � 377 Ohm. This is often called the wave impedance offree space because it is equal to the ratio of the field amplitudes E/H = μ0E/B foran electromagnetic wave traveling in a vacuum. Using the expressions for the fieldsthrough the potentials, it is now straightforward to show that Eq. (11.9) gives thefields (11.4) and (11.6).

The electromagnetic field of a relativistic particle moving with velocity v close tothe speed of light, or equivalently γ � 1, has some remarkable properties that followfrom Eq. (11.4). For a given distance r from the path of the charge, this field reachesa maximum in the mid-plane z = vt , where it scales as E ∝ γ/r2 and is γ timeslarger than the field of a charge at rest. The field remains strong within an angulardistance ∼ 1/γ from this plane, see Fig. 11.1. On the other hand, the field on axis(x = y = 0) is suppressed by a factor of γ2, E ∼ 1/r2γ2. The absolute value of themagnetic field is almost equal to that of the electric field,1 cB ≈ E .

1We actually compare themagnitude of E with cB which have the same dimensions in the SI systemof units that we use in this book. In the Gaussian system of units, the electric and magnetic fieldshave the same dimensions and their magnitude can be compared directly without extra factors.

Page 144: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

140 11 Self Field of a Relativistic Beam

In the case of very large values of γ, towhichwe refer as the ultra-relativistic limit,we can take the limit γ → ∞ and consider the electromagnetic field as localized inan infinitely thin “pancake” region around the transverse plane z = vt . Because inthis plane the field is directed along the vector drawn from the current position of thecharge to the observation point, we can write E = Aρ δ(z − ct)where ρ = xx + yyand A is a factor which is determined by the requirement that the areas under thecurves Ex (z) and Ey(z) agree with the ones given by Eq. (11.4) in the limit γ →∞. The calculation of these integrals is carried out in Problem 11.3; they give thefollowing expressions for the fields:

E = 1

4πε0

2qρ

ρ2δ(z − ct) , B = 1

cz × E . (11.10)

11.2 Interaction of Moving Charges in Free Space

Charged particles in beamsmove in the same general direction and almost on parallelpaths. To get a feeling of the strength of the electromagnetic interaction inside thebunch we consider two such particles and calculate the force of their interaction.

Consider a source particle of charge q that is moving with velocity v along thez-axis, z = vt , and a test particle of the same charge moving on a parallel path. Thetest particle lags behind the leading one at a distance l with an offset a: x = a, y = 0,z = vt − l. We denote by F the force with which the source particle acts on the testone. Using Eq. (11.4), we find for the longitudinal component Fz ,

Fz = qEz = − 1

4πε0

q2l

γ2(l2 + a2/γ2)3/2. (11.11)

The transverse component of the force is directed along x and is equal to

Fx = q(Ex − vBy) = 1

4πε0

q2a

γ4(l2 + a2/γ2)3/2. (11.12)

Consider the limit γ � 1 and assume that the offset x is fixed. In the region l �|a|/γ we can neglect the terms a2/γ2 in the denominators of (11.11) and (11.12) andobtain the scalings Fz ∝ 1/ l2γ2 and Fx ∝ |a|/ l3γ4 which means that the interactionforces rapidly decrease as γ increases. The reasonwhy the transverse force so rapidlydecreaseswith γ is due to the fact that in the relativistic limit the electric andmagneticforces are almost equal to each other and are in opposite directions. The scalings aresomewhat different inside the region l � |a|/γ, however for a given a the length ofthis region decreases as 1/γ, and hence the probability to find particles in that regionis relatively small.

Page 145: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

11.2 Interaction of Moving Charges in Free Space 141

From this simple analysis, we see an indication that in the relativistic limit thedirect interaction between particles in a bunch gets suppressed as the particle energyincreases. In the next section, we will confirm this observation by direct calculationof the fields inside and outside of a relativistic bunch.

11.3 Field of a Relativistic Bunch of Particles

We first calculate the transverse radial component of the electric field outside of thebunch at a distance ρ = √

x2 + y2 from the axis, again assuming that the bunch ismoving along the z-axis. We will be interested in distances ρ that greatly exceed thetransverse size of the beam, ρ � σ⊥. Here, we can completely neglect the transversesize σ⊥, and consider the beam as an infinitely thin line charge. The magnetic fieldsfollow from Eq. (11.6).

To simplify the notation, we assume that the fields are calculated at time t = 0 anddrop the variable t from our equations. The one-dimensional longitudinal distributionfunction of the bunch at this time is denoted by λ(z) and is normalized to unity,∫ ∞−∞ λ(z)dz = 1. The time dependence can be easily recovered in the final resultsby simply replacing λ(z) → λ(z − vt).

We carry out calculations in the lab frame. Each infinitesimally small element ofthe beam generates the electric field given by Eq. (11.4). From this equation we findthat the radial component dEρ created by charge dq ′ located at coordinate z′ is

dEρ(z, z′, ρ) = 1

4πε0

ρ dq ′

γ2[(z − z′)2 + ρ2/γ2]3/2 , (11.13)

where z and ρ are the coordinates of the observation point. To find the field of thewhole bunch we note that the charge dq ′ within dz′ is equal to Qλ(z′)dz′, with Qthe total charge of the bunch. The total field is then obtained by integration of theelementary contributions dEρ:

Eρ(z, ρ) =∫

dEρ(z, z′, ρ)

= Qρ

4πε0γ2

∫ ∞

−∞λ(z′)dz′

[(z − z′)2 + ρ2/γ2]3/2 . (11.14)

This integral can be simplified in two limiting cases. The denominator of theintegrand has a sharp peak of width �z ∼ ρ/γ around z′ = z. At distances ρ � γσz

from the bunch, the width of the peak is smaller than the width σz of the distributionfunction and we can replace the inverse denominator by the delta function:

1

[(z − z′)2 + ρ2/γ2]3/2 → 2γ2

ρ2δ(z − z′) . (11.15)

Page 146: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

142 11 Self Field of a Relativistic Beam

The factor in front of the delta function on the right-hand side follows from therequirement that the area under both functions should be equal, using the mathemat-ical identity

∫ ∞

−∞dz′

[(z − z′)2 + a2]3/2 = 2

a2. (11.16)

The approximation (11.15) is equivalent to using Eqs. (11.10) instead of (11.4).Substituting Eq. (11.15) into (11.14) we obtain

Eρ(z, ρ) = 1

4πε0

2Qλ(z)

ρ. (11.17)

We see that the Lorentz factor γ does not enter this formula — in agreement withwhat one would expect remembering that Eqs. (11.10) are valid in the limit γ → ∞.

Note that in the above calculationwe assumedρ � γσz andσ⊥ � ρ, whichmeansσ⊥ � γσz . In the beam frame of reference, its length is γ times larger than in thelab frame, σ′

z = γσz , while σ′⊥ = σ⊥ (see also Sect. 11.4). We see that in the beam

frame the length of the bunch is much larger than its transverse size, σ′z � σ′

⊥, andhence this regime corresponds to a long-thin approximation for the bunch.When thebunch does not satisfy the long-thin approximation, then the above calculation doesnot apply. It is important to remember that for γ � 1 the long-thin approximationcan hold even when σz ≈ σ⊥ in the lab frame.

In the opposite limit, ρ � γσz , the distribution function λ(z) in Eq. (11.14) canbe considered as a relatively narrow function and replaced by the delta function δ(z),which gives for the bunch field an expression identical to the field of a point charge,

Eρ(z, ρ) = 1

4πε0

Qγρ

(γ2z2 + ρ2)3/2. (11.18)

This result should not be surprising — at a large distance the field cannot resolvethe details of the charge distribution in the bunch and is determined only by the totalcharge Q.

In the intermediate region, ρ ∼ γσz , the field transitions from Eq. (11.17)to (11.18). This transition is illustrated in Fig. 11.2 which shows a numericallycomputed radial electric field (11.14) for a Gaussian distribution function λ(z) =(1/

√2πσz)e−z2/2σ2

z for several intermediate values of the parameter ρ/σzγ, stillassuming σ⊥ � ρ, σzγ.

We now calculate the longitudinal component of the electric field inside the bunch.It is instructive to try again to use the model of an infinitely thin beam as we did in thecalculation of the radial electric field above. In this line-charge model, an elementarycharge dq ′ located at coordinate z′ creates the following field dEz at the observationpoint z on the axis:

dEz(z, z′) = dq ′

4πε0γ2

z − z′

|z − z′|3 . (11.19)

Page 147: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

11.3 Field of a Relativistic Bunch of Particles 143

Fig. 11.2 Transverseelectric field of a relativisticbunch with Gaussiandistribution for variousvalues of the parameterρ/σzγ. This parameter takesthe values of 0.1, 0.5, 1 and 3with larger valuescorresponding to broadercurves. The field isnormalized by(4πε0)

−1Q/γσ2z to yield a

universal set of curves thatonly depend on ρ/γσz

To find the total field, we need to integrate this expression as we did in the derivationof (11.14),

E‖(z) =∫

dEz(z, z′) = Q

4πε0γ2

∫dz′λ(z′)

z − z′

|z − z′|3 . (11.20)

Unfortunately, this integral diverges at z′ → z. The divergence indicates that in thisparticular problem, one cannot assume σ⊥ → 0, and has to take into account thefinite transverse size of the beam.

With the understanding of the importance of the radial charge distribution, let uscalculate the longitudinal electric field in a specific model where the beam radiusis a, and the charge is uniformly distributed over each cross section from ρ = 0 toρ = a. We now slice the beam into infinitesimal disks of thickness dz′ and chargedq ′ = Qλ(z′)dz′. The longitudinal electric field generated by this slice on the axisat point z is

dEz(z, z′) = − 1

4πε0

2dq ′

a2(z − z′)

(1

√a2/γ2 + (z − z′)2

− 1

|z − z′|

)

. (11.21)

We delegate the derivation of this expression to Problem 11.4. The longitudinalelectric field on the axis of the bunch is obtained by integrating over all slices,

Ez(z) =∫ ∞

−∞dEz(z, z

′) (11.22)

= − Q

4πε0

2

a2

∫ ∞

−∞dz′λ(z′)(z − z′)

(1

√a2/γ2 + (z − z′)2

− 1

|z − z′|

)

.

Now the integral converges at z′ = z and can be calculated numerically. For theGaussian distribution function λ(z) the result of the calculation is shown in Fig. 11.3.One can show that in the limit a/γσz � 1 a crude estimate for Ez is given by the

Page 148: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

144 11 Self Field of a Relativistic Beam

Fig. 11.3 Longitudinalelectric field of a relativisticbunch with Gaussiandistribution function forvarious values of theparameter a/γσz . Thisparameter takes the values0.1, 0.01, and 0.001 withsmaller values correspondingto higher fields. The field isnormalized by(2πε0)

−1Q/γ2σ2z

following expression:

Ez ∼ 1

4πε0

Q

γ2σ2z

logγσz

a. (11.23)

This expression diverges in the limit of an infinitely thin beam (a → 0) whichexplains our failure to integrate Eq. (11.20). Note that due to the factor 1/γ2 inthis expression the effect of the longitudinal electric field for relativistic beams isusually small. It is often referred to as the space charge effect.

11.4 Electric Field of a 3D Gaussian Distribution

In accelerator physics, a bunch of charged particles is often represented as having aGaussian distribution function in all three directions so that the charge density ρ of abunch moving with velocity v along the z-axis is given by the following expression:

ρ(x, y, z, t) = Q

(2π)3/2σxσyσze−x2/2σ2

x−y2/2σ2y−(z−vt)2/2σ2

z , (11.24)

where Q is the total charge of the bunch and σx , σy , and σz are the rms bunch lengthsin the corresponding directions. Calculation of the electromagnetic field of such abunch is more complicated than in the simplified models considered in the previoussection. Here we will outline the main steps in this calculation.

First, we will transform into the beam frame where the beam is at rest. The chargedistribution in this frame, ρ′(x ′, y′, z′), does not depend on time and is obtainedfrom (11.24) with the help of the Lorentz transformation for coordinates. Expressingx , y and z − vt through x ′, y′ and z′ we find that the exponential factor in (11.24)becomes

Page 149: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

11.4 Electric Field of a 3D Gaussian Distribution 145

e−x ′2/2σ2x−y′2/2σ2

y−z′2/2σ2z γ

2, (11.25)

from which we conclude that the rms bunch length along z′ is γ times larger thanσz — a relativistic bunch is much longer in the rest frame than in the lab frame. Thetransverse dimensions of the bunch, however, are the same. From the total chargeconservation in the Lorentz transformation, it is now clear that

ρ′(x ′, y′, z′) = Q

(2π)3/2σxσyσ′z

e−x ′2/2σ2x−y′2/2σ2

y−z′2/2σ′2z , (11.26)

where σ′z = γσz .

The electrostatic potential φ′ of the beam at rest is given by Coulomb’s law,

φ′(x ′, y′, z′) = 1

4πε0

∫ρ(ξ,ψ, ζ)dξdψdζ

[(x ′ − ξ)2 + (y′ − ψ)2 + (z′ − ζ)2]1/2 . (11.27)

Unfortunately, with the distribution function given by Eq. (11.26) this integral cannotbe done analytically. There is however a useful trick that considerably simplifies thecalculation. It uses the following mathematical identity,

1

R=

√2

π

∫ ∞

0e−λ2R2/2dλ . (11.28)

Setting R = [(x ′ − ξ)2 + (y′ − ψ)2 + (z′ − ζ)2]1/2 and replacing 1/R in the inte-grand of Eq. (11.27) by (11.28) we first arrive at a four-dimensional integral

φ′ = 1

4πε0

√2

π

∫ ∞

0dλ

∫e−λ2[(x ′−ξ)2+(y′−ψ)2+(z′−ζ)2]/2ρ(ξ,ψ, ζ)dξdψdζ . (11.29)

With the Gaussian distribution (11.26) the integration over ξ, ψ and ζ can now beeasily carried out, e.g.,

∫ ∞

−∞e− 1

2 λ2(x ′−ξ)2−ξ′2/2σ2

x dξ =√2π

√λ2 + σ−2

x

exp

[− x ′2λ2

2(λ2σ2x + 1)

], (11.30)

which gives for the potential

φ′ = 1

4πε0

√2

π

Q

σxσyσ′z

∫ ∞

0

dλ[(λ2 + σ−2

x )(λ2 + σ−2y )(λ2 + σ′−2

z )]1/2

× exp

[

− x ′2λ2

2(λ2σ2x + 1)

− y′2λ2

2(λ2σ2y + 1)

− z′2λ2

2(λ2σ′2z + 1)

]

. (11.31)

Page 150: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

146 11 Self Field of a Relativistic Beam

This integral involves only one integration and is much easier to evaluate numericallythen the original expression (11.27).

Having found the potential in the beam frame, it is now easy to transform it to thelab frame using the Lorentz transformation. First we need to express σ′

z in (11.31)through the bunch length in the beam frame, σ′

z = γσz . Second, from (B.18) we seethat the potential in the lab frame is γ times larger than in the beam frame (note thatA′z = 0 in the beam frame). Third, we need to transform the coordinates x ′, y′, z′

in (11.31) to the lab frame. The resulting expression is:

φ(x, y, z, t) = 1

4πε0

√2

π

Q

σxσyσz

∫ ∞

0

dλ[(λ2 + σ−2

x )(λ2 + σ−2y )(λ2 + γ−2σ−2

z )]1/2

× exp

[

− x2λ2

2(λ2σ2x + 1)

− y2λ2

2(λ2σ2y + 1)

− (z − vt)2λ2

2(λ2σ2z + γ−2)

]

. (11.32)

According to (B.18), in addition to the electrostatic potential, in the lab frame thereis also a vector potential Az responsible for the magnetic field of the moving bunch.It is equal to Az = vφ/c2 with φ given by (11.32).

Examples of using Eq. (11.31) for the calculation of fields are given in Prob-lems 11.6 and 11.7.

Worked Examples

Problem 11.1 Verify by direct calculation that Eqs. (A.5) applied to the potentials(11.9) give the fields (11.4) and (11.6).

Solution: WithR = √(z − vt)2 + (x2 + y2)/γ2, we differentiate directly to find

E = q

4πε0

z(z − vt) + (xx + yy)/γ2

R3+

(Z0βq

) −vz(z − vt)

R3

= q

4πε0

z(1 − β2)(z − vt) + (xx + yy)/γ2

R3

= q

4πε0

rγ2R3

,

using Z0 = 1/ε0c and r = (x, y, z − vt). Similarly,

B = Z0βq

(xy − yx)/γ2

R3= qv

4πε0c2xy − yxγ2R3

= 1

c2v × E .

Page 151: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

11.4 Electric Field of a 3D Gaussian Distribution 147

Problem 11.2 For a fixed transverse offset ρ, calculate the dependence of the mag-nitude of E in Eq. (11.3) versus polar angle θ, where θ is the angle between thevector r , pointing from particle to the observation point, and the z-axis. What formdoes this take in the regime γ � 1?

Solution: Setting for simplicity t = 0, with ρ = (x, y, 0), we have

Eρ = 1

4πε0

qγρ(ρ2 + γ2z2

)3/2 = 1

4πε0

ρ2(1 + γ2 cot2 θ

)3/2 ,

Ez = 1

4πε0

qγz(ρ2 + γ2z2

)3/2 = 1

4πε0

qγ cot θ

ρ2(1 + γ2 cot2 θ

)3/2 ,

where tan θ = ρ/z. The magnitude is then

E =√E2

ρ + E2z = q

4πε0ρ2γ sin2 θ

(sin2 θ + γ2 cos2 θ

)3/2 .

For γ � 1, this becomes very small unless cos θ � 1, so we approximate sin θ �1 and express θ = π/2 + δθ, yielding

E � qγ

4πε0ρ2[1 + γ2(δθ)2

]−3/2.

The fields are localized around |θ − π/2| � 1/γ and the peak magnitude scaleswith γ.

Problem 11.3 UsingEq. (11.4) show that in the limitγ → ∞ the following relationshold:

∫ ∞

−∞Exdz = 1

4πε0

2qx

ρ2,

∫ ∞

−∞Eydz = 1

4πε0

2qy

ρ2.

Solution: Starting from Eq. (11.4) we evaluate

∫ ∞

−∞Exdz = qx

4πε0γ2

∫ ∞

−∞dz

[(z − vt)2 + (x2 + y2)/γ2

]3/2

= qx

4πε0

z − vt

(x2 + y2)[(z − vt)2 + (x2 + y2)/γ2

]1/2

∣∣∣∣∣

z=−∞= qx

2πε0(x2 + y2)

= qx

2πε0ρ2.

The formula for Ey follows identically.

Page 152: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

148 11 Self Field of a Relativistic Beam

Problem 11.4 Derive Eq. (11.21) for dEz and analyze it considering the limits|z − z′| � a/γ and |z − z′| � a/γ, using cylindrical coordinates.

Solution: Consider a disc of radius a located at position z′. A thin ring of thedisk within the radii ρ and ρ + dρ has a charge dq = q/(πa2) × 2πρdρ, where q isthe charge of the disk. The ring creates the longitudinal field on axis at coordinate zequal to

dEz = dq

4πε0γ2

z − z′[(z − z′)2 + ρ2/γ2

]3/2 = 2qρ dρ

4πε0γ2a2z − z′

[(z − z′)2 + ρ2/γ2

]3/2 .

To obtain the total field we need to integrate this expression from ρ = 0 to ρ = a:

Ez = 2q

a21

4πε0γ2

∫ a

0ρdρ

z − z′[(z − z′)2 + ρ2/γ2

]3/2

= 2q

a21

4πε0

−(z − z′)[(z − z′)2 + ρ2/γ2

]1/2

∣∣∣∣∣

a

ρ=0

= 2q

a2(z − z′)4πε0

(−1

[(z − z′)2 + a2/γ2

]1/2 + 1

|z − z′|

)

.

When |z − z′| � a/γ and the disc looks like an infinite sheet to the test charge, thesecond term dominates, and the field reduces to

Ez = 2q

a2sgn(z − z′)

4πε0.

The result is independent of z and γ, as we expect from an infinite sheet of charge,and depends on the sign of z − z′. In the opposite limit when the disc looks like apoint charge, |z − z′| � a/γ, we can approximate

Ez = −2q

a2(z − z′)4πε0

[1

|z − z′|(1 − a2

2γ2(z − z′)2

)− 1

|z − z′|]

= 2q

a2sgn(z − z′)

4πε0

a2

2γ2(z − z′)2,

and we recover the 1/z2 field we expect from a point charge.

Problem 11.5 A bunch of electrons in a future linear collider will have a charge ofabout 1 nC, bunch length σz ≈ 200 μm, and will be accelerated in the linac from 5GeV to 250 GeV over the length of L = 10 km. Estimate the energy spread in thebeam induced by the space charge, assuming the bunch radius of 50 μm.

Solution: With a/γσz = 50/(200γ) � 1, we can estimate the longitudinal fieldas having the form of Eq. (11.23),

Page 153: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

11.4 Electric Field of a 3D Gaussian Distribution 149

Ez ∼ 1

4πε0

Q

γ2σ2z

logγσz

a.

We assume linear acceleration, γ = γ0 + (γ f − γ0)z/L , with γ0 ≈ 104 and γ f ≈5 × 105, and an accelerating gradient of 25 MeV/m (L ≈ 10 km). We can approx-imate log(γσz/a) = log(γ f σz/a) = 14 as a constant, since the log term does notchange much when γ varies from γ0 to γ f . The energy spread from some particlesseeing this maximum electric field and others not being accelerated at all would beroughly

W = e∫ L

0Ezdz

≈ − 1

4πε0

eQ

σ2z

log(γ f σz

a

) L/(γ f − γ0)

γ0 + (γ f − γ0)z/L

∣∣∣∣

L

z=0

≈ 6.5 keV .

Problem 11.6 Derive an expression for the field Ez(z) on the beam axis for a Gaus-sian bunch using the result of Sect.11.4. Assume σx = σy .

Solution: We will carry out calculation of the longitudinal field in the beam frameusing Eq. (11.31) for the potential in which we set σ′

z = γσz and σx = σy = σ⊥,

φ′(ρ′, z′) = 1

4πε0

√2

π

Q

σ2⊥γσz

∫ ∞

0dλ

e− ρ′2λ2

2(λ2σ2⊥+1) e− z′2λ2

2(λ2γ2σ2z +1)

(λ2 + σ−2

⊥) √

λ2 + γ−2σ−2z

.

From E ′z = −∂φ′/∂z′ we then find on axis (ρ = 0)

E ′z = 1

4πε0

√2

π

Q

σ2⊥σzγ

∫ ∞

0dλ

z′λ2

γ2σ2z

e− z′2λ2

2(λ2γ2σ2z +1)

(λ2 + σ−2

⊥) (

λ2 + γ−2σ−2z

)3/2 .

The Lorentz transformation does not change the longitudinal field, Ez = E ′z . Sub-

stituting z′ = γ(z − βct) we get the Ez field in the lab frame,

Ez = 1

4πε0

√2

π

Q

σ2⊥σ3

zγ2(z − βct)

∫ ∞

0dλλ2 e

− γ2(z−βct)2λ2

2(λ2γ2σ2z +1)

(λ2 + σ−2

⊥) (

λ2 + γ−2σ−2z

)3/2 .

We could also have transformed the potentials into the lab frame to find the Ez

field, but then we would have to consider a nonzero vector potential as well.

Page 154: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

150 11 Self Field of a Relativistic Beam

Problem 11.7 Show that at large distances from the center, Eq. (11.31) reduces to

φ′ = Q

4πε0√x ′2 + y′2 + z′2 .

Solution: We would like to evaluate Eq. (11.31),

φ′ = 1

4πε0

√2

π

Q

σxσyσ′z

∫ ∞

0dλ

e− λ2x ′2

2(λ2σ2x+1) e− λ2 y′2

2(λ2σ2y+1) e− λ2 z′2

2(λ2σ′2z +1)

√λ2 + σ−2

x

√λ2 + σ−2

y

√λ2 + σ′−2

z

,

for the case of x � σx (and likewise for y, z). In this case, the exponentials arenon-vanishing only when λ2x2 � 1 =⇒ λ2σ2

x � 1, so in the denominator we can

approximate√

λ2 + σ−2x ≈ 1/σx , and in the exponential function we can neglect

λ2σ2x in comparison with 1. The full integral becomes

φ′ = 1

4πε0

√2

πQ

∫ ∞

0dλe−λ2(x ′2+y′2+z′2)/2

= 1

4πε0

Q√x ′2 + y′2 + z′2 ,

where we used∫ ∞0 e−ax2/2 dx = √

π/2a to do the integral.

References

1. J.D. Jackson, Classical Electrodynamics, 3rd edn. (Wiley, New York, 1999)2. L.D. Landau, E.M. Lifshitz, The Classical Theory of Fields, volume 2 of Course of Theoretical

Physics, 4th edn. (Elsevier, Butterworth-Heinemann, Burlington MA, 1980). (translated fromRussian)

3. L.D. Landau, E.M. Lifshitz, L.P. Pitaevskii, Electrodynamics of Continuous Media, volume 8of Course of Theoretical Physics, 2nd edn. (Elsevier, Butterworth-Heinemann, Burlington, MA,1984). (translated from Russian)

Page 155: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

Chapter 12Effect of Environmenton Electromagnetic Field of a Beam

In the previous chapter we calculated the electromagnetic field of a beam movingwith constant velocity in free space. In reality, beams propagate inside a vacuumchamber, and one has to take into account the effect of the material walls of thechamber on the field. In this chapter, we study several important effects that charac-terize the interactionwithmetal walls. First, we discuss the skin effect that defines thepenetration of the electromagnetic field into the metal. We then discuss an approxi-mation that treats the wall as a perfectly conducting medium. Finally, we calculatethe longitudinal field inside a round metal pipe excited by a relativistic point charge.

12.1 Skin Effect and the Leontovich Boundary Condition

Metals are good conductors and an electromagnetic field that rapidly varies with timeonly penetrates into a thin surface layer of the metal— this is the so-called skin effec-t. If the skin depth is much smaller than the characteristic dimension of the problemunder consideration, the metal surface can be represented by an effective boundarycondition for theMaxwell equations. In this sectionwewill derive this boundary con-dition and, in Sect. 12.3, wewill apply it to the calculation of the electromagnetic fieldexcited by a relativistic charge moving along the axis of a cylindrical metal pipe.

Consider a metal that has a constant conductivity σ and magnetic permeabilityμ. The equations that describe the electromagnetic field inside the metal are theMaxwell equations (A.1) in which we neglect the displacement current ∂D/∂t incomparison with the current density j :

∇ × H = j , ∇ · B = 0 , ∇ × E + ∂B∂t

= 0 , (12.1)

where B = μH . Dropping the term ∂D/∂t from the Maxwell equations is justifiedfor metals with high conductivity σ where the current density j inside the metal,

© Springer International Publishing AG, part of Springer Nature 2018G. Stupakov and G. Penn, Classical Mechanics and Electromagnetismin Accelerator Physics, Graduate Texts in Physics,https://doi.org/10.1007/978-3-319-90188-6_12

151

Page 156: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

152 12 Effect of Environment on Electromagnetic Field of a Beam

Fig. 12.1 Geometry usedfor deriving the boundarycondition on the surface of ametal

j = σE , (12.2)

is dominant. CombiningEqs. (12.1) and (12.2)weobtain an equation for themagneticinduction B,

∂B∂t

= −∇ × E

= −σ−1∇ × j

= −σ−1∇ × ∇ × H

= σ−1[∇2H − ∇(∇ · H)]= (σμ)−1∇2B , (12.3)

where we have used the relation ∇ · H = μ−1∇ · B = 0. We have derived the diffu-sion equation for the magnetic field B in the metal.1

We will now apply this equation to the case of an electromagnetic field thatpenetrates into a metal occupying the half space z > 0, as shown in Fig. 12.1. Letus assume that the field has the time dependence E, H ∝ e−iωt . We choose thecoordinate x so that the tangential component of the magnetic field on the metalsurface is directed along x , Hx = H0e−iωt . Due to the continuity of the tangentialcomponents of H , Hx is the same on both sides of the metal boundary, at z = +0and z = −0. We seek a solution inside the metal in the form Hx = h(z)e−iωt . Thex-component of Eq. (12.3) then reduces to

d2h

dz2+ iμσωh = 0 , (12.4)

which has two solutions, h = H0eikz , where k = ±√iωμσ. Only one of these so-

lutions, with Im k > 0, should be selected as it corresponds to the magnetic fieldvanishing far from the boundary, at z → ∞. The complex wavenumber k can bewritten as

k = √iωμσ = [sgn(ω) + i]

√μσ|ω|2

. (12.5)

1In the Gaussian system of units, the diffusion coefficient is c2/4πσμ, and the conductivity hasdimensions of inverse seconds.

Page 157: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

12.1 Skin Effect and the Leontovich Boundary Condition 153

In this equation we allow ω to take both positive and negative values. The skin depthδ, defined as the quantity

δ(ω) =√

2

μσ|ω| , (12.6)

has dimension of length and it characterizes how deeply the electromagnetic fieldpenetrates into the metal. If the magnetic properties of the metal can be neglected,then μ = μ0, and the skin depth can be written as

δ(ω) =√

2c

Z0σ|ω| , (12.7)

where Z0 = μ0c � 377 Ohm.The electric field inside the metal has only a y component; it can be found from

Eq. (12.2) in which the current is expressed through the magnetic field using the firstequation in (12.1),

Ey = jyσ

= 1

σ

dHx

dz= ik

σHx . (12.8)

We can now understand the mechanism that prevents the penetration of the magneticfield deep inside the metal. The time varying magnetic field in the metal, throughthe third equation in (12.1), generates a tangential electric field that in turn drives alarge current in the skin layer. The magnetic field induced by this current is directedopposite to the magnetic field outside of the metal and shields it from penetratingdeeper than several skin depths.

The relation (12.8) can also be written in the vector notation:

Et = ζH t × n , (12.9)

where n is the unit vector normal to the surface and directed into the metal, Et andH t are the components of the fields oriented tangential to the surface, and the surfaceimpedance ζ is given by

ζ(ω) = − ik

σ= [1 − i sgn(ω)]

√Z0|ω|2cσ

. (12.10)

Because both Et and H t are continuous on the surface of the metal, Eq. (12.9) alsoapplies to the vacuum components of the fields on the wall. With this understanding,Eq. (12.9) becomes a boundary condition for Maxwell’s equations on the metal walland is often called the Leontovich boundary condition. We need to emphasize thatEq. (12.9) is valid in the frequency domain, as indicated by the frequency dependence

Page 158: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

154 12 Effect of Environment on Electromagnetic Field of a Beam

of the parameter ζ. To be able to use it in the time domain, one has tomake the Fouriertransformation from ω to t .

We have derived the boundary condition (12.9) for a flat metal surface. One canalso apply it to curved surfaces if the radius of curvature is much larger than theskin depth δ: in this case one can locally approximate the surface by a plane whensolving the diffusion equation (12.3). Because Eq. (12.9) was derived assuming thatthe metal extends to infinity, it can only be used if the thickness of the metal is muchlarger than the skin depth.

12.2 Perfectly Conducting Boundary Conditions

It follows from Eq. (12.10) that in the limit σ → ∞ the surface impedance van-ishes, ζ → 0, and the boundary condition (12.9) reduces to the requirement for thetangential electric field to be equal to zero,

Et = 0 . (12.11)

We refer to this boundary condition as the perfect conductivity limit.It follows from Eq. (12.11) that the time dependent normal component of the

magnetic field, Bn , also vanishes on the metal wall. To prove this, we consider asmall piece of the wall which can locally be approximated by a flat surface. We thenintroduce a Cartesian coordinate system with the origin on the surface, with the axesx and y located in the plane of the surface and the axis z normal to the surface.According to Eq. (12.11), Ex (x, y, z = 0) = Ey(x, y, z = 0) = 0. It follows thenfrom the Maxwell equations that

∂Bz

∂t

∣∣∣∣z=0

= −(

∂Ey

∂x− ∂Ex

∂y

)∣∣∣∣z=0

= 0 . (12.12)

Hence, unless there is a static magnetic field in the system, Bz = 0, or using the moregeneral notation Bn for the component perpendicular to the metal surface,

Bn = 0 . (12.13)

In applications, one often encounters problems where one needs to calculate theelectromagnetic field of a beam of relativistic particles traveling along the axis of astraight cylindrical pipe of a given cross section. In some cases it may be easier to docalculations in the beam frame and then transform it to the laboratory frame usingthe Lorentz transformations. Of course, the Maxwell equations are the same in bothframes, but in general the boundary condition on the wall is expressed differently inthe beam frame compared to the boundary condition in the lab frame. A remarkableproperty of the perfectly conducting boundary condition (12.11) is that it is the samein both frames. Indeed, using the prime to denote the field components in the beam

Page 159: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

12.2 Perfectly Conducting Boundary Conditions 155

frame, and from the Lorentz transformations for the fields (B.17), we can write thetangential components of the electric field in the beam frame as:

E ′t z = Etz , E′

t⊥ = γ[Et⊥ + v( z × B)t

], (12.14)

where we have assumed that the velocity is directed along the z axis, v = v z, andused the subscript t⊥ to indicate the component of the field that is both tangentialto the metal surface and perpendicular to the velocity. It is now easy to see thatthe components of the vector product z × B that are tangential to the metal wallinvolve only themagnetic field component normal to thewall, ( z × B)t = ( z × Bn)t .Using the boundary conditions in the laboratory frame, Eqs. (12.11) and (12.13), weconclude that E ′

t z = 0 and E′t⊥ = 0 and hence the whole tangential component of the

electric field in the beam frame vanishes, E′t = 0. We thus proved that the perfectly

conducting boundary condition holds in any reference frame moving with a constantvelocity directed parallel to the surface of the metal.

12.3 Round Pipe with Resistive Walls

As an example of the applications of the Leontovich boundary condition, in thissection we will calculate how a round, conducting metal pipe influences the forcebetween two charges, q1 and q2, moving along the axis of a pipe of radius a withrelativistic velocity v, as shown in Fig. 12.2. We assume that the metal walls of thepipe have conductivity σ.

An important simplification in this problem comes from the fact that when thevelocity v is close to the speed of light we can take the limit v → c and set v = cin Maxwell’s equations. In this limit, the electromagnetic field of charge q2 does notpropagate ahead of it and hence charge q2 does not exert any force on charge q1. Onthe other hand, charge q1 generates a longitudinal electric field Ez behind it that doesact on charge q2. Our goal now is to find the field Ez at the location of charge q2.

Assuming that the conductivity of the pipe is sufficiently large, we will calculatethe fields using perturbation theory in which the solution is obtained as a Taylorexpansion in the small parameter σ−1. In the lowest approximation, we consider thepipe as being a perfect conductor. In this case, the electromagnetic field of charge

Fig. 12.2 Point charges q1and q2 moving in a roundpipe of radius a

Page 160: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

156 12 Effect of Environment on Electromagnetic Field of a Beam

q1 is the same as the free space solution given by Eq. (11.10), where it is assumedthat at time t the charge is located at z = ct . Indeed, it follows from Eq. (11.10) thatthe electric field is perpendicular to the wall of the pipe as required by the boundarycondition (12.11) in the limit of perfect conductivity. For what follows, we will onlyneed the magnetic field Bθ,

Bθ(ρ, z, t) = 1

4πε0

2q1cρ

δ(z − ct), (12.15)

where ρ is the radial coordinate in the cylindrical coordinate system. Note that thismagnetic field also satisfies the boundary condition (12.13).

Using the mathematical identity

δ(z − ct) = 1

2πc

∫ ∞

−∞dωe−iω(t−z/c), (12.16)

we Fourier transform Bθ and represent it as

Bθ(ρ, z, t) =∫ ∞

−∞dω Bθ(ρ)e−iωt+iωz/c, (12.17)

where

Bθ(ρ) = 1

4πε0c2q1πρ

= μ0q14π2ρ

. (12.18)

We can now make the next step in the perturbation approach and take into accountthe finite conductivity σ. In this approximation, the tangential electric field on themetal wall is not zero. To find the Fourier component Ez of the longitudinal electricfield on the surface of the metal we will substitute into the Leontovich boundarycondition (12.9) the magnetic field (12.18) computed in the limit σ → ∞,

Ez|ρ=a = −ζBθ(a)

μ0= −[1 − i sgn(ω)]

√Z0|ω|2cσ

q14π2a

. (12.19)

Having found the electric field on the wall, we now need an equation that determinesthe radial dependence of Ez . This equation is provided by the z-component of thewave equation (A.3) which in the cylindrical coordinate system has the form

1

c2∂2Ez

∂t2− ∂2Ez

∂z2− 1

ρ

∂ρρ∂Ez

∂ρ= 0 . (12.20)

It is clear that together with the magnetic field (12.15), the electric field depends onz and t in the same combination z − ct . With this dependence, the first two terms

Page 161: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

12.3 Round Pipe with Resistive Walls 157

in Eq. (12.20) cancel out leaving only the radial term, which we now write for theFourier transform Ez ,

1

ρ

∂ρρ∂ Ez

∂ρ= 0 . (12.21)

This equation has a general solution Ez(ρ) = A + B ln ρ, where A and B are arbitraryconstants. We do not expect Ez to have a singularity on the axis, hence B = 0, whichmeans that the electric field does not depend on ρ, and the electric field on the axisis given by the same equation (12.19),

Ez|ρ=0 = Ez|ρ=a . (12.22)

We have also proven that in the ultrarelativistic case the longitudinal electric fieldinside the pipe is constant throughout the pipe cross section at any given z and t .

To find Ez(z, t) we substitute Eq. (12.19) into the Fourier integral,

Ez(z, t) =∫ ∞

−∞dω Eze

−iω(t−z/c) , (12.23)

which gives

Ez(z, t) =√

Z0

2cσ

q14π2a

∫ ∞

−∞dω[i sgn(ω) − 1]√|ω|e−iω(t−z/c) . (12.24)

Strictly speaking, the integral in this equation diverges as |ω| → ∞. This happensbecause, as we will show below, our approach is not valid in the limit of very highfrequencies. To regularize the integral, we will introduce a convergence factor e−ε|ω|with ε > 0 into the integrand and then take the limit ε → 0 in the result. Using themathematical identity2:

limε→0

∫ ∞

−∞dω[i sgn(ω) − 1]√|ω|e−iωξ−ε|ω| =

√2π

ξ3/2h(ξ) , (12.25)

where h is the step function, we obtain,

Ez(z, t) =√

Z0

2cσ

q14π2a

√2π

(t − z/c)3/2h(t − z/c) = q1c

4π3/2a

√Z0

σs3h(s) . (12.26)

Here s = ct − z is understood as the distance between the charges q1 and q2, andthe step function indicates that the field does not propagate ahead of charge q1. Thepositive sign of Ez means that a trailing charge (if it has the same sign as q1) will be

2This identity can be established with the help of the following formula:∫ ∞0 dx

√xe±x−εx =

12

√π(ε ∓ i)−3/2.

Page 162: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

158 12 Effect of Environment on Electromagnetic Field of a Beam

accelerated in thewake.Recall fromSect. (11.1) that the longitudinal fields in vacuumscale as 1/γ2. Thus, for highly relativistic beams the impact of the resistive wall canbe greater than the space charge fields that have been ignored in this derivation.

In our derivation we took the magnetic field on the wall to be the same as in thecase of perfect conductivity which is the same as the magnetic field in free space,Eq. (12.15), where it is generated by the current associatedwith themoving charge q1.We are now in a position to estimatewhen this approximation is valid. Accounting forthe finite conductivity, the magnetic field changes because in addition to the movingcharge there is an additional source — the displacement current ∂D/∂t (see theMaxwell equations (A.1d)). The z-component of the displacement current density is

jdispz = ε0∂Ez

∂t. (12.27)

To be able to neglect the contribution of jdispz to Bθ, we require the total displacementcurrent πa2 jdispz to be much smaller than the beam current. In the Fourier represen-tation, the time derivative ∂/∂t is replaced by the multiplication factor −iω, and therequirement reduces to

ε0πa2ω|Ez| � | I | , (12.28)

where I is the Fourier component of the beam current. The latter is calculated bymaking the Fourier transformation of the equation I = q1cδ(z − ct) for the currentof a point charge, yielding | I | = q1/2π. Substituting this into Eq. (12.28) and usingEq. (12.19) for Ez we obtain

ω

c�

(Z0σ

a2

)1/3

. (12.29)

As was pointed out above, we find that indeed our solution in the frequency domainis limited to the range of frequencies defined by this inequality. In the space-timedomain, a high spectral frequency ω corresponds to a small distance s behind thecharge q1, s ∼ c/ω. The inequality (12.29)means that our solution (12.26) is valid fordistances s s0 where s0 is of the order of the inverse right-hand side of Eq. (12.29):

s0 ≡(2a2

Z0σ

)1/3

(12.30)

(the extra factor of 2 which has been added inside the cubic root is conventionallyused in the literature). Values of s0 are typically small for good conductors: for a pipeof radius a = 5 cmmade of copper (σ = 5.8 × 107 1/Ohm·m) we have s0 = 60µm.

The solution Ez for small values of s, s � s0, can be found in the literature,see [1]. In Fig. 12.3 we show a plot of Ez for s comparable to s0. Note that thesingularity Ez ∼ s−3/2 in Eq. (12.26) actually saturates for s � s0, and the electric

Page 163: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

12.3 Round Pipe with Resistive Walls 159

Fig. 12.3 Longitudinalelectric field as a function ofdistance s from the particle.The field is normalized byq1/4πε0a2, and the distanceis normalized by s0. Thevalue of the normalized fieldat the origin is equal to -4

field normalized by q1 changes sign here and is negative at s = 0. The negative signof Ez/q1 when s → 0 is expected: it corresponds to deceleration of the charge q1and is explained by the fact that this charge loses energy which goes into the heatingof the metal walls.

Worked Examples

Problem 12.1 Calculate the skin depth in copper (σ = 5.8 × 107 1/Ohm·m) andstainless steel (σ = 1.4 × 106 1/Ohm·m) at the frequency of 5 GHz.

Solution: The skin depth is given by δ = √2/μ0σω, with μ0 = 1.3 × 10−6 H/m.

At ω = 2π f = 10π GHz, copper has δ = 0.9 μm. For stainless steel, δ = 6 μm.

Problem 12.2 Given the tangential component B0e−iωt of the magnetic field on thesurface, use the Poynting vector to find the averaged over time energy absorbed inthe metal per unit time per unit area.

Solution: To find the energy of the EMwave going into the metal, we will use thePoynting vector, S = E × H , which defines the energy flow per unit time per unitarea. We have a tangential component of the magnetic field, H = B0e−iωt/μ, andE = ζH × n, with ζ = (1 − i)/σδ(ω) (here we assume ω > 0). Before substitutingthe fields in the Poynting vector, we need to take their real parts,

S = Re [(ζH × n)] × Re [H]

= 1

σδ(ω)Re [(1 − i)H] · Re [H] z

= B20

μ2σδ(ω)[(cosωt − sinωt) cosωt] z .

Page 164: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

160 12 Effect of Environment on Electromagnetic Field of a Beam

This can be thought of as the instantaneous rate of energy flow in and out of thesurface. Averaging cos2 ωt over time gives 1/2, while (sinωt cosωt) averages to 0.Taking δ(ω) = √

2/μσω, we find

〈Sz〉 = B20

2μ2σδ(ω)

= ωδ(ω)B20

4μ∝ √

ω .

For comparison, the electromagnetic energy density just outside the surface isB20/2μ0.

Problem 12.3 An ultra-relativistic bunch with charge Q propagates along the axisof a round metal pipe of radius a. Using Eq. (12.26) calculate the energy loss of thebunch per unit length. Assume a Gaussian distribution with the rms bunch length σz .

Solution: The electric field Ez at longitudinal coordinate s1, with s1 = ct − zindicating the relative position inside thebunch, canbe computed as a superpositionoffields generated by all preceding elementary charges. For this, wewill use Eq. (12.26)in which q1 is replaced by Qλ(s2)ds2, the distance is set to s = s2 − s1, and theformula is integrated over s2. We obtain

Ez(s1) = Qc

4π3/2a

√Z0

σ

∫ ∞

s1

ds2 λ(s2)(s2 − s1)−3/2 . (12.31)

This integral actually diverges as s2 → s1. The reason for this divergence is that,as we discussed in Sect. 12.3, Eq. (12.26) becomes invalid in the limit s → 0. Thisdivergence can be eliminated if we express the field’s dependence on s through apartial derivative using

(s2 − s1)−3/2 = −2

∂s2(s2 − s1)

−1/2 ,

which is valid everywhere except for s2 = s1. Integrating equation (12.31) by partsand discarding the boundary term we obtain

Ez(s1) = Qc

2π3/2a

√Z0

σ

∫ ∞

s1

ds2 λ′(s2)(s2 − s1)−1/2 ,

where the prime denotes derivative with the respect to the argument.3

3Getting rid of the singularity through the integration by parts is a subtle point. A careful analysisshows that it requires the point charge field Ez(s) to have the property

∫ ∞0 ds Ez(s) = 0. This is

indeed satisfied by the field plotted in Fig. 12.3.

Page 165: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

12.3 Round Pipe with Resistive Walls 161

The energy loss per unit length is obtained by integrating eEz(s1) with the distri-bution function over s1,

dEdz

= e∫ ∞

−∞ds1Ez(s1)λ(s1)

= eQc

2π3/2a

√Z0

σ

∫ ∞

−∞ds1 λ(s1)

∫ ∞

s1

ds2 λ′(s2)(s2 − s1)−1/2 .

The two-dimensional integral can be simplified by replacing the second integrationvariable and changing the order of integration,

∫ ∞

−∞ds1 λ(s1)

∫ ∞

s1

ds2 λ′(s2)(s2 − s1)−1/2

=∫ ∞

−∞ds1 λ(s1)

∫ ∞

0dξ λ′(s1 + ξ)ξ−1/2

=∫ ∞

0dξ ξ−1/2

∫ ∞

−∞ds1 λ(s1)λ

′(s1 + ξ) .

For the Gaussian distribution function λ(s) = (2π)−1/2σ−1z e−s2/2σ2

z the inner integralcan be calculated analytically,

∫ ∞

−∞ds1 λ(s1)λ

′(s1 + ξ) = − 1

4√

πσ3z

ξe−ξ2/4σ2z .

Using the mathematical identity

∫ ∞

0dt t1/2e−t2/4 = √

2�

(3

4

),

where � is the gamma function, we arrive at the following result:

dEdz

= − eQc

25/2π2aσ3/2z

√Z0

σ�

(3

4

).

The negative sign in this formula indicates that the bunch is losing energy.This result is valid if the bunch length ismuch longer than the parameter s0 defined

by Eq. (12.30), σz s0.

Reference

1. K.L.F. Bane, M. Sands, The short-range resistive wall wakefields. AIP Conf. Proc. 367, 131(1996)

Page 166: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

Chapter 13Plane Electromagnetic Wavesand Gaussian Beams

Plane electromagnetic waves are solutions of the Maxwell equations that areunbounded in the plane perpendicular to the direction of propagation. They approx-imately describe local properties of the real field far from the source of radiation.They can also be used as building blocks fromwhich a general solution ofMaxwell’sequations in free space can be constructed.An important practical example of electro-magnetic radiation that finds many applications in accelerator physics and elsewhereis a focused laser beam. The distribution of the electromagnetic field in such lightis characterized by Gaussian modes which can be considered as a superposition ofplanewaves propagating at small angles to the direction of the beam.Gaussian beamscorrectly describe the field structure near the focus and the diffraction of the beam asit propagates away from the focal region. In this chapter, we will briefly summarizethe main properties of plane electromagnetic waves, and then derive the field in aGaussian beam.

13.1 Plane Electromagnetic Waves

A plane electromagnetic wave is a solution of Maxwell’s equations that describespropagation of electromagnetic fields in free space. In this solution, all componentsof the field depend only on one variable ξ = z − ct ,

E(r, t) = F(ξ) , B(r, t) = G(ξ) , (13.1)

where F and G are vector functions of ξ. From the equation ∇ · E = 0 it followsthat ∂Fz/∂ξ = 0, from which we conclude that Fz = 0 (a more general solutionFz = const describes a constant longitudinal electric field and is not related to thewave). Similarly, the equation ∇ · B = 0 yields Gz = 0. We conclude that a planewave is transverse: the electromagnetic field is perpendicular to the direction ofpropagation z.

© Springer International Publishing AG, part of Springer Nature 2018G. Stupakov and G. Penn, Classical Mechanics and Electromagnetismin Accelerator Physics, Graduate Texts in Physics,https://doi.org/10.1007/978-3-319-90188-6_13

163

Page 167: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

164 13 Plane Electromagnetic Waves and Gaussian Beams

Applying the Maxwell equation ∂B/∂t = −∇ × E to the fields (13.1) we find

F ′x = c G ′

y , F ′y = −c G ′

x , (13.2)

where the prime denotes differentiation with respect to the argument ξ. From theserelations we conclude that Fx = c Gy and Fy = −c Gx (again, we neglect possi-ble constants of integration that describe constant fields). In vector notation, theserelations can be written as F = −cn × G, or

E = −c n × B , (13.3)

where n is a unit vector in the direction of propagation (in our case n is directedalong the z axis). Multiplying vectorially Eq. (13.3) by n, we also obtain

B = 1

cn × E . (13.4)

The potentials φ and A for a plane wave also depend on ξ only: φ = φ(ξ), A =A(ξ). A useful formula that we will use later expresses the magnetic field in a planewave through the time derivative of the vector potential A:

B = ∇ × A

= −xA′y + yA′

x

= n × A′

= −1

cn × ∂A

∂t, (13.5)

where on the last step we replaced the derivative d/dξ by the partial derivative−c−1∂/∂t . After the magnetic field is found, the electric field can be obtained fromEq. (13.3).

In applications, a plane wave often has a sinusoidal time dependence with somefrequency ω. For such waves, it is convenient to use the complex notation:

E = Re(E0e−iωt+ik·r+iφ0) , B = Re(B0e

−iωt+ik·r+iφ0) , (13.6)

where E0 and B0 are the amplitudes of the wave with E0 = cB0, k = nω/c is thewavenumber that defines the direction of propagation, and φ0 is the phase. In general,E0 and B0 can be complex vectors orthogonal to k, i.e., E0 = E(r)

0 + iE(i)0 with E(r)

0

and E(i)0 real. When E(r)

0 and E(i)0 are collinear, the wave has a linear polarization;

when they are orthogonal and equal in length, the wave is circularly polarized; inthe most general case, the polarization is elliptical.

The energy flow in the plane wave is given by the Poynting vector,

S = E × H = n√

ε0

μ0E20 cos

2(−ωt + k · r + φ0) , (13.7)

Page 168: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

13.1 Plane Electromagnetic Waves 165

Fig. 13.1 Approximation oflocal plane waves far from aradiative system of chargesshown by the blue blob.Solid lines indicate the phasefronts of the waves. Note thatin different regions thedirections n of the wavepropagation are different

(in this formula E0 is assumed real). As follows from this formula, the energy flowsalong n, in the direction of propagation of the wave. Denoting by S the energy flowaveraged over time, we find

S = n2

√ε0

μ0E20 = n

E20

2Z0= n

c2B20

2Z0. (13.8)

In reality one never dealswith planewaves that occupy thewhole space as is indicatedby Eq. (13.6). Usually, a planewave provides a useful approximation to a propagatingelectromagnetic field in a limited region of space. We will use this approximationin Chaps. 16–18 to study the radiation field of moving charges at large distances.Figure13.1 illustrates how a plane-wave approximation is applied in different regionsof the radiation field.

Another important aspect of sinusoidal plane waves is that an arbitrary solutionof Maxwell’s equations in free space (without charges) can be represented as asuperposition of plane waves with various amplitudes and directions of propagation.This is illustrated by Problem 13.1.

13.2 Gaussian Beams

The plane-wave approximation from the previous section is not sufficient for describ-ing a laser beam propagating towards its focus and then diffracting away from thefocal point. A better representation of laser beams is provided by Gaussian modes in

Page 169: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

166 13 Plane Electromagnetic Waves and Gaussian Beams

free space. In this section, we will derive the field in an axisymmetric fundamentalGaussian mode using the paraxial approximation, which assumes that the field isrepresented by a plane wave with an amplitude that slowly varies in space.

We start from the wave equation (A.3) for the x component of the electric fieldassuming a linear polarization of the laser light,

∂2Ex

∂x2+ ∂2Ex

∂y2+ ∂2Ex

∂z2− 1

c2∂2Ex

∂t2= 0 . (13.9)

Let us seek a solution to this equation in the following form,

Ex (x, y, z, t) = u(x, y, z)e−iωt+ikz , (13.10)

where ω = ck and u is a slow varying function of its arguments mathematicallyexpressed in the requirement

∣∣∣∣1u∂u

∂z

∣∣∣∣ � k . (13.11)

As we will see from the result, u also varies slowly in the transverse directions x andy. Note that a constant u corresponds to a sinusoidal plane electromagnetic wavepropagating along the z axis; allowing u to vary in space takes us beyond the limitsof the plane-wave approximation.

Putting Eq. (13.10) into (13.9) and neglecting the second derivative ∂2u/∂z2 incomparison with k∂u/∂z yields

∂2u

∂x2+ ∂2u

∂y2+ 2ik

∂u

∂z= 0 . (13.12)

We are looking for an axisymmetric solution to this equation that depends on ρ =√x2 + y2, so that u = u(ρ, z). Eq. (13.12) then becomes

1

ρ

∂ρρ∂u

∂ρ+ 2ik

∂u

∂z= 0 . (13.13)

The fundamental Gaussianmode has the following particular dependence on ρ and z:

u = A(z)eQ(z)ρ2 , (13.14)

where A(z) and Q(z) are as yet unknown functions.Wewill see that Q has a negativereal part so that u exponentially decays in the radial direction.

Substituting Eq. (13.14) into (13.13) yields

4Q2ρ2u + 4Qu + 2ik

(A′

A+ Q′ρ2

)= 0 , (13.15)

Page 170: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

13.2 Gaussian Beams 167

where the prime denotes differentiation with respect to z. In order for this equationto be valid for arbitrary ρ, both the coefficient in front of ρ2 and the terms that do notcontain ρ should vanish,

2Q2 + ikQ′ = 0 ,

2Q + ikA′

A= 0 . (13.16)

The solution of the first equation can be written as

Q(z) = − 1/w20

1 + 2i z/kw20

, (13.17)

wherew0 is a constant of integration which has dimension of length and is called thebeam waist. Substituting Q into the second of Eq. (13.16) and integrating it yields

A(z) = E0

1 + 2i z/kw20

, (13.18)

where E0 is another constant of integration that has a meaning of the field amplitude.We now introduce two important geometrical parameters: the Rayleigh length ZR

and the angle θ:

ZR = kw20

2, θ = w0

ZR= 2

kw0. (13.19)

The Rayleigh length and the waist can be expressed through the angle θ and thereduced wavelength λ = k−1 = c/ω:

ZR = 2λ

θ2, w0 = 2

λ

θ. (13.20)

Rewriting Q and A as Q = −w−20 /(1 + i z/ZR) and A = E0/(1 + i z/ZR) and sub-

stituting them into Eq. (13.14), we arrive at the following expression for the electricfield:

u = E0

1 + i z/ZRe−ρ2/w2

0(1+i z/ZR) . (13.21)

At a given coordinate z, the radial dependence of |u| is ∝ e−ρ2/w20(1+z2/Z2

R), whichmeans that the quantityw0 gives the minimal transverse size of the mode at the focal

spot z = 0. On the axis of the beam, ρ = 0, we have |u| = E0/

√1 + z2/Z2

R , andhence ZR is the characteristic length of the focal region along the z axis. The angle θcharacterizes the divergence of the beam far from the focus. These geometricalcharacteristics of the mode are illustrated in Fig. 13.2.

Page 171: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

168 13 Plane Electromagnetic Waves and Gaussian Beams

Fig. 13.2 Geometricparameters of a Gaussianbeam

We can now derive conditions for the validity of the paraxial approximation.Evaluating ∂2u/∂z2 ∼ u/Z2

R and k∂u/∂z ∼ ku/ZR we see that in order to be ableto neglect the second derivative in Eq. (13.12) we need to require ZR � λ or, equiv-alently, θ � 1. The latter condition means that the waves that constitute a Gaussianbeam propagate at small angles to the axis z, which is the origin of the term paraxial.A small θ also means that λ � w0 � ZR which implies that the size of the focal spotw0 is much larger than the reduced wavelength λ. Locally, on the scale of several λ,a Gaussian beam can be considered as a plane wave, but on a larger scale∼ w0 in thetransverse direction and ∼ ZR in the longitudinal direction, the field substantiallydeviates from the plane-wave approximation.

The magnetic field in a Gaussian beam can be found in the lowest order usingEq. (13.4). In our case n is directed along z, hence the magnetic field is directedalong y,

By = 1

cEx . (13.22)

Let us now look at the field at a large distance from the focus. At |z| � ZR we canapproximate Q and A as:

A(z) ≈ − i E0kw20

2z, Q(z) ≈ ik

2z− k2w2

0

4z2. (13.23)

From these equations it follows that the amplitude of the field on the axis decaysas 1/z, and the radial profile at coordinate z as given by the absolute value of theexponential factor is e−k2w2

0ρ2/4z2 . The region occupied by the field in the radial

direction can be estimated as ρ ∼ 2z/kw0 = θz, it expands as the field propagatesaway from the focus. There is also a correction Im Qρ2 = kρ2/2z to the plane-wavephase kz, so that the total phase is

φ = kz + kρ2

2z. (13.24)

Page 172: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

13.2 Gaussian Beams 169

One can give this phase the following physical interpretation. Let us introducethe distance R from the focal point at the origin of the coordinate system to theobservation point, R = √

z2 + ρ2. In the limit of large z, assuming z > 0, we haveR ≈ z + ρ2/2z, andwe see that the phase (13.24) is approximately equal to kR. Sucha phase dependence is characteristic for a spherical wave, and we conclude that at alarge distance from the focus a Gaussian beam represents a spherical wave divergingaway from the focus. The electric and magnetic fields in this wave are not constanton the wave front — they are localized within the angle θ defined by Eq. (13.19).

Worked Examples

Problem 13.1 At time t = 0 the electromagnetic field in free space is given by func-tions E0(r) and B0(r) such that ∇ · E0 = ∇ · B0 = 0. Find the field at time t byrepresenting E(r, t) and B(r, t) as integrals over plane waves.

Solution: We represent the time-dependent electromagnetic field as a combinationof waves propagating in both directions, along k and −k,

E(t, r) =∫

d3k(e−ir·k+iωtE(+)

k + e−ir·k−iωtE(−)

k ) ,

B(t, r) =∫

d3k(e−ir·k+iωtB(+)

k + e−ir·k−iωtB(−)

k ) , (13.25)

with ω = c|k|. From Eq. (13.4) applied to each wave it follows that

B(+)

k = 1

ωk × E(+)

k ,

B(−)

k = − 1

ωk × E(−)

k . (13.26)

At time t = 0 we have

E(0, r) =∫

d3ke−ir·k(E(+)

k + E(−)

k ) = E0(r) ,

B(0, r) =∫

d3ke−ir·k(B(+)

k + B(−)

k ) = B0(r) ,

which means that the sumsE(+)

k + E(−)

k andB(+)

k + B(−)

k are equal to the correspond-ing Fourier transforms of the initial fields:

E(+)

k + E(−)

k = E0,k ≡ 1

(2π)3

∫d3reir·kE0(r),

B(+)

k + B(−)

k = B0,k ≡ 1

(2π)3

∫d3reir·kB0(r) . (13.27)

Page 173: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

170 13 Plane Electromagnetic Waves and Gaussian Beams

Note that E0,k · k = B0,k · k = 0. Solving this equation in combination withEq. (13.26) we find

E(−)

k = 1

2

(E0,k + c2

ωk × B0,k

),

E(+)

k = 1

2

(E0,k − c2

ωk × B0,k

).

Together with Eq. (13.26) this solves the problem.

Problem 13.2 A plane electromagnetic wave propagates at some angle in a framemoving with velocity βc along the z axis. The magnitude of the Poynting vector atsome location in the wave is equal to S′. Show that in the lab frame the magnitudeof the Poynting vector at this location is given by the following equation:

S = S′

γ2(1 − β cos θ)2, (13.28)

where θ is the angle between the direction of propagation in the lab frame and the zaxis.

Solution: First, consider the plane wave in the lab frame where it is moving at anangle θ to the z axis. Assume the E field is in the x-z plane, so that |B| = By = B0.The amplitude of the electric field is E0 = cB0. We have Ex = E0 cos θ and Ez =−E0 sin θ. Transforming into the moving frame,

E ′z = Ez = −E0 sin θ ,

E ′x = γ[Ex + (υ × B)x ] = γE0(cos θ − β) ,

where (υ × B)x = −βE0 was used in the final line. According to Eq. (13.8) thePoynting vector in the moving frame is S′ = E ′2/2Z0. We have

S′ = E ′2

2Z0= E ′2

x + E ′2z

2Z0

= E20

2Z0

(sin2 θ + γ2(cos2 θ − 2β cos θ + β2)

)

= E20

2Z0γ2 (1 − β cos θ)2

= Sγ2 (1 − β cos θ)2 ,

where S = E20/2Z0 is the Poynting vector in the lab frame and we have used the

identity 1 + γ2β2 = γ2. Note that it is necessary to start in the lab frame since we

are given the angle, θ, in the lab frame. Alternatively, if we start in the moving frame,

Page 174: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

13.2 Gaussian Beams 171

we find

S = S′γ2(1 + β cos θ′)2 .

Then using cos θ′ = (cos θ − β)/(1 − β cos θ) we find

S = S′γ2

(1 + β cos θ − β2

1 − β cos θ

)2

= S′ 1

γ2(1 − β cos θ)2.

For a plane wave in which the electric field is polarized along y, one obtains thesame result considering the Lorentz transformation for themagnetic field. Finally, foran arbitrary polarization, thePoyntingvector is a sumof twoorthogonal polarizations,and hence is also transformed according to Eq. (13.28).

Problem 13.3 Calculate the longitudinal electric field Ez in the laser beam usingthe equation ∇ · E = 0.

Solution: From∇ · E = 0, for aGaussian beampolarized along x wehave∂Ez/∂z =−∂Ex/∂x . With Ex = A(z)eQ(z)(x2+y2)e−iωt+ikz , we have

∂Ex

∂x= 2xQ(z)Ex .

If we assume Ez = υ(x, y, z)e−iωt+ikz with slowly varying function υ, ∂υ/∂z �ikυ, then ∂Ez/∂z = (υ−1∂υ/∂z + ik)Ez ≈ ikEz . We can then put together the twoderivatives to find

Ez ≈ − 1

ik

∂Ex

∂x

= −2xQ(z)Ex

ik

= 2x

ikw20 − 2z

Ex .

For z small and x ∼ w0, this reduces to Ez ∼ 2Ex/ ikw0 ∼ θEx � Ex , so weconclude Ez � Ex .

Problem 13.4 Show that the energy flux in the laser beam (the Poynting vectorintegrated over the cross section of the beam) is equal to

π

4Z0E20w

20 .

Page 175: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

172 13 Plane Electromagnetic Waves and Gaussian Beams

Solution: We need to integrate over the beam cross section the time averaged Poynt-ing vector S = E2

x/2Z0. At z = 0, we have u = E0e−ρ2/w20 , so

S = E20

2Z0

∫2πρ dρ e−2ρ2/w2

0 = π

4Z0E20w

20 .

Problem 13.5 Expand the Gaussian laser field over plane waves.

Solution: The exact solution of Maxwell’s equations corresponding to a Gaussianmode with the frequency ω can be represented as a superposition of plane electro-magnetic waves with different wavenumbers kx and ky :

E = e−iωt∫

E(kx , ky)eikx x+iky y+ikz z

(x − kx

kzz)

dkx dky ,

where kz(kx , ky) ≡ (k2 − k2x − k2y)1/2 with k = ω/c and E(kx , ky) is the wave ampli-

tude with the wavenumbers kx and ky . Note that due to the particular form of theexpression in the brackets in the integrand this field satisfies the equation ∇ · E = 0.Tofind thewaveamplitudes E(kx , ky)we take the x components of thefield, set z = 0,

Ex |z=0 = e−iωt∫

dkx dky E(kx , ky)eikx x+iky y ,

and equate it to the Gaussian mode at z = 0,

E0e−iωt e−(x2+y2)/w2

0 .

From this equality, E(kx , ky) can be found through the inverse Fourier transform:

E(kx , ky) = E0

(2π)2

∫dx dy e−ikx x−iky ye−(x2+y2)/w2

0 .

The integral on the right-hand side factors out into a product of one-dimensionalintegrals which are easily integrated,

∫ ∞

−∞dx e−ikx x−x2/w2

0 = √πw0e

−k2xw20 ,

and we find

E(kx , ky) = E0w20

πe−(k2x+k2y)w

20 .

Note that the typical value for the angles relative to the z axis is kx/k ∼ ky/k ∼1/w0k ∼ θ.

Page 176: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

Chapter 14Waveguides and RF Cavities

In the previous chapter we studied propagation of electromagnetic waves in freespace. Free space is an idealization which can be used when the waves propagatefar from material boundaries. If this is not the case, one has to take into accountthe interaction of the electromagnetic field with the medium. In this chapter we willstudy one example that is especially important for accelerator applications, when themedium can be modeled as a perfectly conducting metal whose boundaries reflectthe electromagnetic field without losses. The impact of resistive losses can oftenbe treated perturbatively, as we have seen in Chap. 12. We will consider cylindricalwaveguides and resonant cavities. We will also discuss how cavity eigenmodes canbe excited by relativistic beams of charged particles.

14.1 TMModes in Cylindrical Waveguides

We consider a cylindrical waveguide of radius a with perfectly conducting wallsand choose a cylindrical coordinate system (r,φ, z) with the z-axis directed alongthe cylinder axis. Such a waveguide can serve as a conduit for the transportation ofelectromagnetic energy along its axis in the form of eigenmodes that have particulartransverse and longitudinal field distributions. We will first focus our attention onthe so-called transverse magnetic modes, or TM modes, that have zero longitudinalcomponent of the magnetic field, Bz = 0, but nonzero Ez . To find Ez in TM modes,we will assume that

Ez(r,φ, z, t) = E(r)e−iωt−imφ+iκz , (14.1)

where ω is the mode frequency, m is an integer number that defines the azimuthalvariation of the field, andκ is the longitudinal wavenumber. The function Ez satisfiesthe wave equation (A.3) which in the cylindrical coordinate system r,φ, z reads

© Springer International Publishing AG, part of Springer Nature 2018G. Stupakov and G. Penn, Classical Mechanics and Electromagnetismin Accelerator Physics, Graduate Texts in Physics,https://doi.org/10.1007/978-3-319-90188-6_14

173

Page 177: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

174 14 Waveguides and RF Cavities

1

r

d

drr

dEdr

− m2

r2E +

(ω2

c2− κ

2

)E = 0 . (14.2)

The solution to this equation is

E = E0 Jm (k⊥r) , (14.3)

where Jm is the Bessel function of mth order and k⊥ = c−1√

ω2 − c2κ2. The bound-ary condition Ez = 0 at r = a requires that Jm (k⊥a) be equal to zero. For eachfunction Jm , there is an infinite sequence of zeros jm,n with n = 1, 2, . . ., such thatJm

(jm,n

) = 0. Hence k⊥ = jm,n/a for a given choice of m, n, and we can expressthe longitudinal wavenumber κ of that mode with the indices m, n in terms of thefrequency as

κm,n = ±(

ω2

c2− j2m,n

a2

)1/2

. (14.4)

We see from this equation that in order for a mode to have a real value of κm,n ,its frequency should be larger than the cut-off frequency cjm,n/a. The term ‘cut-offfrequency’ is generally used to refer to frequencies where the wave number goes tozero. The positive values ofκm,n describe modes propagating in the direction of the zaxis (assumingω > 0), and the negative ones correspond tomodes propagating in theopposite direction. For ω < cjm,n/a we deal with evanescent modes that, dependingon the sign of Imκm,n exponentially decay or grow along the z-axis. Such modesplay an important role in the formation of localized fields around obstacles inside awaveguide.

Given Ez(r,φ, z, t) defined by Eqs. (14.1) and (14.3) we can find all other com-ponents of the electric and magnetic fields using Maxwell’s equations. They will allhave the same dependence e−iωt−imφ+iκz versus time, angle and z. The radial distri-bution of the four unknown components Eφ, Er , Bφ and Br (we recall that Bz = 0)are found from the four algebraic equations, which are the r and φ components ofthe two vector equations ∇ × E = iωB and c2∇ × B = −iωE. A straightforwardcalculation yields:

Er = E0iκm,na

jm,nJ ′

m

(jm,n

r

a

)e−iωt−imφ+iκm,n z , (14.5a)

Eφ = −E0mκm,na2

r j2m,n

Jm

(jm,n

r

a

)e−iωt−imφ+iκm,n z , (14.5b)

Br = E0mωa2

c2r j2m,n

Jm

(jm,n

r

a

)e−iωt−imφ+iκm,n z , (14.5c)

Bφ = E0iωa

c2 jm,nJ ′

m

(jm,n

r

a

)e−iωt−imφ+iκm,n z , (14.5d)

Page 178: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

14.1 TM Modes in Cylindrical Waveguides 175

where J ′m is the derivative of the Bessel function of order m with respect to its

argument. These modes are designated as TMmn . Note that in addition to vanishingEz on the wall, which we have satisfied by choosing k⊥ = jm,n/a, we should alsomake sure that the other tangential component, Eφ, is equal to zero at r = a. Thisis indeed satisfied because the radial dependence of Eφ in (14.5b) is the same as Ez

in (14.3).Of course, only the real parts of Eqs. (14.1) and (14.5) have physical meaning.

Since the longitudinal wavenumbers (14.4) do not depend on the sign of m, themodes with positive and negative values of m are degenerate—they have the samevalues of κm,n . Often the sum and difference of m and −m modes, which converteimφ and e−imφ into cosmφ and sinmφ, are used as an alternative choice for the setof fundamental eigenmodes in cylindrical waveguides.

14.2 TE Modes in Cylindrical Waveguides

Transverse electric modes, or TE modes, have Ez = 0 and nonzero longitudinalmagnetic field Bz . They can be derived following closely the derivation of TMmodesin the previous section. However, a simple observation of a special symmetry ofMaxwell’s equations allows one to obtain thefields inTEmodeswithout an additionalderivation.

Assuming the time dependence ∝ e−iωt for all fields, Maxwell’s equations in freespace take the following form:

∇ × E = iωB, c2∇ × B = −iωE, ∇ · E = 0, ∇ · B = 0 . (14.6)

Note that the transformation

(E,B) → (cB,−E/c) (14.7)

does not change Eq. (14.6), or in other words Maxwell’s equations are invariantwith respect to this transformation. This means that having found any solution ofMaxwell’s equation one can obtain a second solution by means of the simple substi-tution (14.7), if the new solution satisfies proper boundary conditions. In the deriva-tion of TM modes we satisfied the boundary condition of zero tangential electricfield on the wall by choosing k⊥ = jm,n/a. The boundary conditions for TE modesis different, so we replace jm,n in Eq. (14.5) by yet unknown numbers j ′

m,n and carryout the transformation (14.7) (also replacing E0 by cB0),

Page 179: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

176 14 Waveguides and RF Cavities

Bz = B0 Jm

(j ′m,n

r

a

)e−iωt−imφ+iκm,n z ,

Br = B0iκm,na

jm,nJ ′

m

(j ′m,n

r

a

)e−iωt−imφ+iκm,n z ,

Bφ = −B0mκm,na2

r j2m,n

Jm

(j ′m,n

r

a

)e−iωt−imφ+iκm,n z ,

Er = −B0mωa2

r j2m,n

Jm

(j ′m,n

r

a

)e−iωt−imφ+iκm,n z ,

Eφ = −B0iωa

jm,nJ ′

m

(j ′m,n

r

a

)e−iωt−imφ+iκm,n z . (14.8)

These modes are designated by TEmn . The only tangential component of the electricfield on the wall in TE modes is Eφ and in order for it to be equal to zero at r = awe require

J ′m

(j ′m,n

) = 0 , (14.9)

which means that j ′m,n are the roots of the derivative J ′

m of the Bessel function.

14.3 RF Modes in Cylindrical Resonators

A cylindrical resonator is a cylindrical waveguide with the ends closed by metallicwalls perpendicular to its axis. The waveguide modes described in the previoussection are now reflected by the end walls and become trapped in the resonator.The resonator modes can be obtained as linear combinations of waveguide modespropagating in the opposite directions.

In comparison with waveguides, a resonator has two additional boundaryconditions—the vanishing tangential electric field on the end walls. Let us assumethat the left wall is located at z = 0, and the right wall is at z = L , where L isthe resonator length. We start with TM modes. To satisfy the boundary conditionEr = Eφ = 0 at z = 0 we choose two waveguide modes with the same frequencyand the same m and n indices but opposite values of κm,n (that is two identical wavespropagating in the opposite directions), add them and divide the result by 2. Withthe help of the relations

1

2(eiκm,n z + e−iκm,n z) = cos(κm,nz), (14.10)

1

2(κm,neiκm,n z − κm,ne−iκm,n z) = iκm,n sin(κm,nz) ,

it is easy to see that both Er and Eφ acquire the factor sin(κm,nz) and hence satisfythe boundary condition at z = 0. To satisfy the boundary condition at the opposite

Page 180: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

14.3 RF Modes in Cylindrical Resonators 177

wall, z = L , we require κm,n L = lπ, where l = 0, 1, 2, . . ., is an integer number. Asa result we obtain the following expressions for the fields:

Ez = E0 Jm

(jm,n

r

a

)cos

(lπz

L

)e−iωt−imφ , (14.11)

Er = −E0lπa

L jm,nJ ′

m

(jm,n

r

a

)sin

(lπz

L

)e−iωt−imφ ,

Eφ = −E0imlπa2

Lr j2m,n

Jm

(jm,n

r

a

)sin

(lπz

L

)e−iωt−imφ ,

Br = E0mωa2

c2r j2m,n

Jm

(jm,n

r

a

)cos

(lπz

L

)e−iωt−imφ ,

Bφ = E0iωa

c2 jm,nJ ′

m

(jm,n

r

a

)cos

(lπz

L

)e−iωt−imφ .

Because κm,n is now determined by the boundary conditions on the end walls,Eq. (14.4) should now be solved for ω: replacing κm,n by lπ/L , we find for thefrequency ω of the mode:

ω

c= ±

[(lπ

L

)2

+ j2m,n

a2

]1/2

. (14.12)

The frequency depends on the indices m, n, and l as well as on the waveguidedimensions. The modes defined by Eqs. (14.11) and (14.12) are called the TMmnl

modes.A similar derivation can be carried out with the TE waveguide modes, but instead

of adding, we need to subtract the mode with a negative κm,n from the mode withthe positive κm,n and divide the result by 2i . This gives the following expressionsfor the fields,

Bz = B0 Jm

(j ′m,n

r

a

)sin

(lπz

L

)e−iωt−imφ , (14.13)

Br = B0lπa

L jm,nJ ′

m

(j ′m,n

r

a

)cos

(lπz

L

)e−iωt−imφ ,

Bφ = B0imlπa2

Lr j2m,n

Jm

(j ′m,n

r

a

)cos

(lπz

L

)e−iωt−imφ ,

Er = −B0mωa2

r j2m,n

Jm

(j ′m,n

r

a

)sin

(lπz

L

)e−iωt−imφ ,

Eφ = −B0iωa

jm,nJ ′

m

(j ′m,n

r

a

)sin

(lπz

L

)e−iωt−imφ .

Page 181: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

178 14 Waveguides and RF Cavities

The frequency of the mode is

ω

c= ±

[(lπ

L

)2

+(

j ′m,n

a

)2]1/2

. (14.14)

The modes given by these equations are called the TEmnl modes. Note that in TEmodes l can not be equal to zero.

An important mode characteristic is the energy W of the electromagnetic field inthe mode of a given amplitude. This energy is obtained by integrating (ε0/2)(E2 +c2B2) over the volume of the cavity, where one has to take the real parts of the fieldsbefore squaring them.

For illustration, let us calculate the energy of the TM010 mode. In this modem = l = 0, and as follows from Eq. (14.11) only Ez and Bφ are not equal to zero.The calculation can be simplified if one notices that although Ez and Bφ depend ontime, the energy W does not. Because there is a phase shift π/2 between these fields,at a moment when Bφ = 0 the electric field reaches its maximum value equal to theabsolute value of the complex expression |Ez|, and

W = ε0

2

∫dV |Ez|2 = ε0

2

∫ a

02πr dr

∫ L

0dz|Ez|2

= ε0

2πE2

0a2L J 21 ( j0,1) , (14.15)

where we have used the integral∫ 10 xdx J 2

0 (bx) = 12 [J 2

0 (b) + J 21 (b)] and we note

that J0( j0,1) = 0.In the analysis above we have assumed the perfect conductivity of the walls. In

this approximation, there are no energy losses, and the modes have real frequencies.If one takes into account the finite wall conductivity, one finds that an initially excitedmode decays with time because its energy is absorbed in the walls. This damping isreflected in the imaginary part γ in the mode frequency, ω = ω′ − iγ, where ω′ andγ are real and positive. A consistent derivation of the imaginary part of the frequencycan be carried out by imposing the Leontovich boundary conditions on the surfaceof the metal instead of zero tangential electric field, as has been done in our analysis.Related to the damping is the quality factor Q of the cavity mode,

Q = ω′

2γ. (14.16)

Without derivation, we will give here the quality factor for the fundamental modeTM010 of the cylindrical cavity,

Q = aL

δ(a + L), (14.17)

Page 182: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

14.3 RF Modes in Cylindrical Resonators 179

where δ is the skin depth calculated at the frequency of the mode. For a crudeestimate of the quality factor one can use Q ∼ l/δ, where l is a characteristic size ofthe cavity (assuming that all dimensions of the cavity are of the same order). Typicalcopper cavities used in accelerators have Q ∼ 104; superconducting cavities mayhave quality factors as high as Q ∼ 1010.

14.4 Electromagnetic Field Pressure

An electromagnetic field confined by material surfaces exerts a force on these sur-faces. In the most general formulation, this force can be derived from the so-calledMaxwell stress tensor, see [1], Sect. 6.7. In this section, we will give a simplifiedderivation of this force for the case of a metal boundary.

If electric field lines are terminated on a metal plate as shown in Fig. 14.1, thereare image charges on the surface of the metal with a surface charge density equalto ε0En , where the subscript n indicates that the field is normal to the metal wall.To calculate the force, we need to consider in more detail the distribution of theelectric field inside the metal. Let us assume that z = 0 corresponds to the surfaceof the metal and the metal occupies the region z > 0. Denote by ρ(z) the chargedensity and Ez(z) the electric field inside the metal. From the Maxwell equation∇ · E = ρ/ε0 we obtain

d Ez

dz= ρ(z)

ε0. (14.18)

The force per unit area of the metal is given by the integral

f (E)z =

∫ ∞

0dz ρEz . (14.19)

To take this integral, we express ρ from Eq. (14.18) to obtain

Fig. 14.1 Electric field linesare normal to the metalsurface where they areterminated

Page 183: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

180 14 Waveguides and RF Cavities

Fig. 14.2 Magnetic fieldlines are parallel to thelateral metal surface

f (E)z = ε0

∫ ∞

0dz Ez

d Ez

dz= −ε0

2E2

n , (14.20)

where we integrated by parts noting that Ez d Ez/dz = (d/dz)E2z /2 and took into

account that deep inside the metal the field vanishes, Ez(∞) = 0, while on thesurface Ez(0) = En . The minus sign in this equation means that the electric field hasa “negative pressure”—it pulls the surface in the direction of free space. Remarkably,the details of the charge distribution in Eq. (14.18) dropped out from the final result.

Similarly to the electric field, a magnetic field near a metal wall also exerts aforce on the metal, but the field is tangential to the surface as shown in Fig. 14.2. Tocompute the force, we assume that the magnetic field By(z) is directed along y, andvaries along z due to the current jx (z) flowing in the x direction inside the metal.From the Maxwell equation

d Hy

dz= − jx , (14.21)

with the expression for the force per unit area

f (M)z =

∫ ∞

0dz jx By , (14.22)

we find

f (M)z = −

∫ ∞

0dz By

d Hy

dz= 1

2μ0B2

t , (14.23)

where we took into account that By(∞) = 0, and on the surface By(0) = Bt (weneglect the magnetic properties of the metal and assume μ = μ0). We see that f (M)

zis positive—it acts as a real pressure applied to the surface.

Page 184: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

14.4 Electromagnetic Field Pressure 181

The effect of the electromagnetic pressure is usually small, however it causesa so-called Lorentz detuning in modern superconducting cavities as the resultingdeformation increases with increasing amplitude of the mode, which should be com-pensated by a special control system (see [2], p. 580).

14.5 Slater’s Formula

In practice, resonant cavities have complicated shapes and their resonant frequenciesare calculated numerically using computer codes. In many cases it is important tobe able to estimate how small deformations of the cavity shape, either due to manu-facturing errors or from temporary distortions created, for example, by temperaturechanges, affect the frequency of the cavity mode. The resonant frequency can alsobe tuned by small amounts by applying an external force to shift the cavity walls.The electromagnetic forces derived in the previous section allow us to calculate thechange in the frequency of a mode if the cavity shape is slightly distorted as shownin Fig. 14.3.

Let us assume that a mode with frequency ω is excited in the cavity, and the cavitywalls are being slowly moved with the electromagnetic field oscillating inside. Dueto the pressure of the electromagnetic field, to change the cavity shape requires somework, d A, (which can be positive or negative) to act against the pressure forces.From the energy balance, this work changes the electromagnetic energy of the mode,δW = d A. Since the distortions of the shape are assumed small, to calculate theforce we can take the unperturbed distribution of the electric and magnetic fields onthe walls, compute the sum of the electric and magnetic pressures f (E)

z + f (M)z and

average them over the period of oscillations. This averaging introduces a factor 12 in

Eqs. (14.20) and (14.23). We then multiply the force by the displacement h of thesurface, and integrate it over the area of the dent,

δW = 1

2

∫S

hd S

(1

2μ0|Bt |2 − ε0

2|En|2

), (14.24)

Fig. 14.3 Initial (black line)and distorted (blue line withshaded interior) cavityshapes

Page 185: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

182 14 Waveguides and RF Cavities

where h is assumed positive in the case when the volume of the cavity decreases,and negative otherwise. The quantities Bt and En are understood as the amplitudevalues of the fields on the surface. A positive (negative) value of δW means that theelectromagnetic energy in the mode increases (decreases).

The easiest way to relate the change of the frequency of the mode to the energychange δW , is to think about the modes as a collection of photons of frequency ω inthe cavity. The number of quanta of the electromagnetic field in the cavity does notchange if the cavity reshaping occurs adiabatically slow. Because the photon energyis �ω, this number is proportional to the ratio of the electromagnetic energy to thefrequency, W/ω, and by requiring W/ω = const we conclude that

δω

ω= δW

W. (14.25)

Using Eq. (14.24) we obtain

δω

ω= ε0

4W

∫�V

dV(c2|Bt |2 − |En|2

), (14.26)

where the integration in the numerator goes over the volume of the dent. Note thatthe right-hand side of this formula does not depend on the amplitude of the mode:the integral is proportional to the amplitude of the field squared, but this dependenceis canceled by the energy in the denominator that also scales as the square of thefield. Equation (14.26) is often called Slater’s formula.

As was emphasized in the derivation, for Eq. (14.26) to be valid, the change in theshape of the cavity should be small. In addition to this requirement, we also add thatthe perturbation should be smooth—otherwise the field variation near sharp edgesof the dent would be large, and one could not use the unperturbed fields of the modein the calculation of the energy (14.24).

14.6 Excitation of a Cavity Mode by a Beam

In accelerators, cavity resonators are predominantly used to accelerate beams ofcharged particles. The accelerating mode is excited by an external source that feedsradio-frequency power into the cavity through a coupling in the cavity wall. In addi-tion to this controlled external excitation, a beam of accelerating particles passingthrough the resonator also contributes to the excitation of the mode. This effect,besides presenting a challenge in controlling the fields in accelerating cavities, canalso be exploited to form the basis of cavity beam positionmonitors, which are highlyaccurate beam diagnostic tools [3].

Page 186: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

14.6 Excitation of a Cavity Mode by a Beam 183

In this section we will calculate the mode excitation by a point charge using amethod originally proposed by P. Wilson in Ref. [4]. We will assume that the chargemoves with velocity close to the speed of light and use the approximation v = c. Ourderivation is based on the principles of superposition and conservation of energy.Assume that a point charge q enters an empty cavity at time t = 0, and is movingalong the z axis, z = ct . We represent the real longitudinal component of the electricfield of the mode under consideration as

Ez(z) = E0e(z) , (14.27)

where E0 is the amplitude and e(z) gives the distribution of the field in the modealong the z axis at a given time. Note that the function e(z), being a solution of thehomogeneous wave equation, is defined with an arbitrary normalization factor. Wewill choose this normalization factor so that the total electromagnetic energy W ofthe mode is equal to E2

0 ,

W = E20 . (14.28)

When the particle passes through the cavity, the amplitude of the mode E0 varieswith time, E0(t), starting from an initial zero value, E0(0) = 0. As the particle movesfrom z = ct to z + dz = ct + cdt due to the interaction with the field of the mode itchanges its amplitude by an infinitesimal value d E0. We can find d E0 from energyconservation: the change in the mode energy dW is equal in magnitude, but oppositein sign, to the work of the electric field of the mode on the charge,

dW = −q Ezdz = −q E0e(ct)c dt . (14.29)

On the other hand, from Eq. (14.28) we have dW = 2E0d E0. Equating this expres-sion to Eq. (14.29) we find the change in the mode amplitude,

d E0 = −q

2e(ct)c dt . (14.30)

To find the final electric field E0 at the moment when the charge exits the cavity atz = L , t0 = L/c, we need to add all the infinitesimal contributions (14.30) along theparticle orbit. It is important to take into account that after an infinitesimal amplituded E0 is excited at time t it oscillates with the frequency of the mode ω and at timet0 evolves to d E0 exp[−iω(t0 − t)] (in this formula we implicitly assume that theexcited field starts the oscillations with a zero phase). Integrating equation (14.30)with proper phase factors we find

E0(t0) = −qc

2

∫ L/c

0dt e−iω(t0−t)e(ct) = −qV

2e−iωt0 , (14.31)

Page 187: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

184 14 Waveguides and RF Cavities

where

V =∫ L

0dzeiωz/ce(z) . (14.32)

An important characteristic of the mode excitation is the total energy deposited bythe charge to the cavity. This energy, normalized per unit charge, is called the lossfactor. Since the final amplitude of the field (14.31) is a complex number, for themode energywe need to replace the square of the field in Eq. (14.28)with the absolutevalue of the field, W = |E0(t0)|2, which gives for the loss factor kloss

kloss = 1

q2|E0(t0)|2 = |V 2|

4. (14.33)

Equation (14.33) was derived assuming a special normalization of the mode (14.28).We can drop this constraint if we redefine V by replacing e(z) on the right-hand sideof Eq. (14.32) by Ez(z). This introduces an extra factor E0 into V which we cancelby putting E2

0 = W into the denominator of the right-hand side of Eq. (14.33),

kloss = 1

4W

∣∣∣∣∫ L

0dzeiωz/c Ez(z)

∣∣∣∣2

. (14.34)

The advantage of this formula is that now kloss does not depend on the normalizationof Ez(z): multiplying Ez(z) by a factor of λ adds a factor of λ2 to both the numeratorand denominator of the fraction and cancel each other.

Worked Examples

Problem 14.1 Calculate the TM modes in a rectangular waveguide with crosssection a × b.

Solution: Choosing the coordinate system so that the waveguide area occupiesthe region 0 < x < a, 0 < y < b, we require Ez = 0 on the surfaces x = 0, x = a,y = 0, and y = b. We can use simple plane waves, and choose the horizontal andvertical terms to be proportional to sin functions with appropriate kx and ky :

Ez(x, y, z, t) = E0 sin(mπx

a

)sin

(nπy

b

)e−iωt+iκz .

The dispersion relation is

κm,n = ±(

ω2

c2− m2π2

a2− n2π2

b2

)1/2

.

Page 188: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

14.6 Excitation of a Cavity Mode by a Beam 185

We calculate the horizontal electric field, first by taking the horizontal componentof c2∇ × B = −iωE, and recalling that Bz ≡ 0:

−iωEx = c2(

∂Bz

∂y− ∂By

∂z

)= −c2

∂By

∂z.

Then, from the vertical component of ∇ × E = iωB,

iωBy =(

∂Ex

∂z− ∂Ez

∂x

).

Combining these expressions yields

(ω2

c2+ ∂2

∂z2

)Ex = ∂2Ez

∂z ∂x.

Using the initial expression for Ez and assuming that the horizontal componenthas the same z-dependence, Ex ∝ eiκm,n z , we find that

(ω2

c2− κ

2m,n

)Ex = ∂2Ez

∂x∂z= iκm,n

aE0 cos

(mπx

a

)sin

(nπy

b

).

A similar calculation applies for Ey , or we can simply use ∂Ey/∂x − ∂Ex/∂y =iωBz = 0. The result for all of the other components is

Ex = i E0κm,nmπ/a

(mπ/a)2 + (nπ/b)2cos

(mπx

a

)sin

(nπy

b

)e−iωt+iκz ,

Ey = i E0κm,nnπ/b

(mπ/a)2 + (nπ/b)2sin

(mπx

a

)cos

(nπy

b

)e−iωt+iκz ,

Bx = −i E0ω

c2nπ/b

(mπ/a)2 + (nπ/b)2sin

(mπx

a

)cos

(nπy

b

)e−iωt+iκz ,

By = i E0ω

c2mπ/a

(mπ/a)2 + (nπ/b)2cos

(mπx

a

)sin

(nπy

b

)e−iωt+iκz .

Problem 14.2 Follow up on Problem 14.1 and derive TE modes in a rectangularwaveguide by applying the transformation of Eq. (14.7) to TM modes and satisfyingthe boundary conditions on the wall.

Solution: We will write Bz in the same form as Ez , but add as-yet unknown phaseterms φx and φy . Together with the transformations (14.7) we find

Page 189: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

186 14 Waveguides and RF Cavities

Bz = B0 sin(mπx

a+ φx

)sin

(nπy

b+ φy

)e−iωt+iκz ,

Bx = i B0κm,nmπ/a

(mπ/a)2 + (nπ/b)2cos

(mπx

a+ φx

)sin

(nπy

b+ φy

)e−iωt+iκz ,

By = i B0κm,nnπ/b

(mπ/a)2 + (nπ/b)2sin

(mπx

a+ φx

)cos

(nπy

b+ φy

)e−iωt+iκz ,

Ex = i B0ωnπ/b

(mπ/a)2 + (nπ/b)2sin

(mπx

a+ φx

)cos

(nπy

b+ φy

)e−iωt+iκz ,

Ey = −i B0ωmπ/a

(mπ/a)2 + (nπ/b)2cos

(mπx

a+ φx

)sin

(nπy

b+ φy

)e−iωt+iκz .

The boundary conditions for the electric field are: Ex = 0 when y = 0 or y = b, andEy = 0 when x = 0 or x = a. This requires both φx and φy to be π/2 (or −π/2 butthat just changes the overall sign of all the fields). The resulting field terms are:

Bz = B0 cos(mπx

a

)cos

(nπy

b

)e−iωt+iκz ,

Bx = −i B0κm,nmπ/a

(mπ/a)2 + (nπ/b)2sin

(mπx

a

)cos

(nπy

b

)e−iωt+iκz ,

By = −i B0κm,nnπ/b

(mπ/a)2 + (nπ/b)2cos

(mπx

a

)sin

(nπy

b

)e−iωt+iκz ,

Ex = −i B0ωnπ/b

(mπ/a)2 + (nπ/b)2cos

(mπx

a

)sin

(nπy

b

)e−iωt+iκz ,

Ey = i B0ωmπ/a

(mπ/a)2 + (nπ/b)2sin

(mπx

a

)cos

(nπy

b

)e−iωt+iκz .

Problem 14.3 Consider a point charge passing through a cylindrical cavity wherethe fundamental mode is excited with amplitude E0. Calculate the maximum energygain for the charge.

Solution: The fundamental mode for a cylindrical cavity is the TM010 mode,which has a physical electric field given by

Ez = E0 J0(ωr

c

)cos(ωt − φ) ,

where the field amplitude E0 is assumed to be positive. This field is localized insidethe cavity, 0 < z < L . The charged particle moves on a trajectory z = v(t − t0),where t0 is the time when the particle enters the cavity. We set r = 0 where theBessel function is unity and assume that the velocity does not change significantlyduring the acceleration.

The instantaneous force on the particle is q Ez , so the total energy gained acrossthe whole cavity is W = ∫

dz q Ez . Thus we obtain

Page 190: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

14.6 Excitation of a Cavity Mode by a Beam 187

W = q∫ L

0dz Ez = q E0

∫ L

0dz cos(ωz/v + ωt0 − φ)

= 2q E0v

ωsin

(Lω

2v

)cosψ ,

where we defined ψ = ωt0 − φ + ωL/2v.As a function of ψ, the maximum energy gain occurs when ψ = 0 for positive q,

or when ψ = π for negative q. As a function of the cavity length L , the maximumpossible energy gain in one cavity is 2|q|E0v/ω and occurs if L = πv/ω. The mosteffective rate of acceleration comes fromusingmultiple cavities,with phases adjustedfrom cavity to cavity.

Problem 14.4 The radius of a cylindrical cavity is changed by a small quantity δa,and the length is changed by δL. Consider this as a deformation of the cavity shapeand find the frequency change of the fundamental mode TM010 in the cavity usingSlater’s formula. Verify that the result agrees with Eq. (14.12).

Solution: As discussed in this chapter, during an adiabatic change to a cavity thenumber of photons should remain constant, so δω/ω = δW/W . We calculate δWresulting from expanding the volume of a cylinder by δa (radius) and δL (length).The radial expansion has no contribution from the E field, because there is no Efield normal to the radial walls. So for the E field we get only a contribution from thelongitudinal term, given by the integral over the surface area, multiplied by the lengthof expansion, h = −δL . We may assume that only one side of cylinder, with surfaceS1 at a fixed longitudinal position, is shifted by the full amount δL to make the cavitylonger. By definition, h is negative for an expansion (because we are calculatingworkdone to the field), so we find

δWE = −ε0

4(−δL)

∫S1

d S |Ez(r)|2

= ε0

4δL

∫ a

02πrdr E2

0 J 20 (ωr/c)

= ε0

4δLπE2

0a2 J 21 (ωa/c) ,

following the derivation of Eq. (14.15) and setting the frequency ω = j0,1c/a. Thenfor the B field, Bφ = E0(i/c)J1(ωr/c), we have contributions from both δL and δabecause the B field is tangential to the surface everywhere:

δWB = ε0c2

4

[(−δL)

∫S1

d S |Bφ(r)|2]

+ (−δa)

∫S2

d S |Bφ(r)|2

= −ε0c2

4

[δL

∫ a

02πrdr E2

0 J 21 (ωr/c)/c2 + 2πaδaL E2

0 J 21 (ωa/c)/c2

]

= −ε0

4δLπE2

0a2 J 21 (ωa/c) − ε0

42πaδaL E2

0 J 21 (ωa/c) ,

Page 191: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

188 14 Waveguides and RF Cavities

where S2 is the radial boundary of the cavity, r = a, so themagnetic fieldmagnitude isconstant everywhere along S2.We also used the identity

∫ 10 xdx J 2

1 (bx) = 12 [J 2

1 (b) −J0(b)J2(b)].

We can see that the effect of the E field cancels the first term from the B field, soin total

δW = δWE + δWB = −ε0

42πaδaL E2

0 J 21 ( j1) .

Using for the mode energy Eq. (14.15) we find

δω = δW

Wω = −c δa

a2j0,1 .

This result agrees with the explicit formula for the frequency of the TM010 modeω = cj0,1/a from which it follows that δω = −cj0,1δa/a2.

Problem 14.5 Find the loss factor for the fundamental mode TM010 of the cylindricalcavity. Assume r = 0.

Solution: The longitudinal electric field Ez in the TM010 mode does not dependon z, Ez = E0. The loss factor is given by k = |V 2|/4W , with

V =∫ L

0eiωz/c Ez(z)dz

= E0

∫ L

0eiωz/cdz

= E01

ik(eikL − 1) .

So |V |2 = 4E20 sin

2(kL/2)/k2 = E20 L2sinc2(kL/2). The energy in the mode is

given by Eq. (14.15), W = (ε0/2)πE20a2L J 2

1 ( j0,1), so we find

k = L sinc2(kL/2)

2πε0a2 J 21 ( j0,1)

.

References

1. J.D. Jackson, Classical Electrodynamics, 3rd edn. (Wiley, New York, 1999)2. A.W.Chao,M.Tigner,Handbook of Accelerator Physics and Engineering, Third printing edition

(World Scientific Publishing, Singapore, 2006)

Page 192: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

References 189

3. Sean Walston, Stewart Boogert, Carl Chung, Pete Fitsos, Joe Frisch, Jeff Gronberg, HitoshiHayano, Yosuke Honda, Yury Kolomensky, Alexey Lyapin, Stephen Malton, Justin May, Dou-glas McCormick, Robert Meller, David Miller, Toyoko Orimoto, Marc Ross, Mark Slater, SteveSmith, Tonee Smith, Nobuhiro Terunuma, Mark Thomson, Junji Urakawa, Vladimir Vogel,David Ward, Glen White, Performance of a high resolution cavity beam position monitor sys-tem. Nucl. Instrum. Methods Phys. Res. A 578, 1–22 (2007)

4. P. Wilson. High energy electron linacs: Applications to storage ring rf systems and linear col-liders. Report SLAC-AP-2884 (Rev.), SLAC (1991)

Page 193: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

Chapter 15Radiation and Retarded Potentials

In this chapter, based on simple intuitive arguments we derive the Liénard–Wiechertpotentials that solve the problemof the electromagnetic field of a point chargemovingin free space.

15.1 Radiation Field

Let us consider a point charge that has been at rest for t < 0. At t = 0 it is rapidlyaccelerated to velocity υ0 andmoveswith this velocity from then on. The dependenceof the charge velocity versus time is shown in Fig. 15.1a. Before the acceleration,the charge has a spherically symmetric Coulomb electric field with straight fieldlines emanating from the origin. We would like to understand how the field linesof the electric field look like after the acceleration. From the wave equation for theelectromagnetic field we know that the field perturbations propagate with the speedof light. Hence, outside of the sphere of radius r = ct , the field “does not know” thatthe charge has been set in motion, and the field remains the same Coulomb field thatit had been before the acceleration. Inside of the sphere, the electromagnetic fieldchanges into the field of a moving charge described by Eq. (11.4). In a thin sphericalsheath in the vicinity of the radius r = ct the Coulomb field transitions from thefield of a charge at rest to the field of a moving charge, as shown in Fig. 15.2a. Thistransitional layer expanding with the speed of light is the radiation electromagneticfield in this case.

If the charge is accelerated two times, at t = 0 and then at t = t1, as shownin Fig. 15.1b, the field lines at time t > t1 would look like shown in Fig. 15.2b.In addition to the original spherical shell r = ct , there is another one of radiusr = c(t − t1) with the center at z = υ0t1, that is at the location of the charge at themoment of the second acceleration.We now have two spherical shells of the radiationfield.

© Springer International Publishing AG, part of Springer Nature 2018G. Stupakov and G. Penn, Classical Mechanics and Electromagnetismin Accelerator Physics, Graduate Texts in Physics,https://doi.org/10.1007/978-3-319-90188-6_15

191

Page 194: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

192 15 Radiation and Retarded Potentials

(a) (b)

Fig. 15.1 The velocity versus time of an abruptly accelerated charge

(a) (b)

Fig. 15.2 Field lines corresponding to the acceleration shown in Fig. 15.1a, b, respectively. Thespherical shell in figure a and the bigger sphere in figure b have their origin at z = 0. The smallersphere in figure b is centered at z = υ0t1

One can now easily imagine that a continuously accelerating charge is constantlyradiating spherical shells at every moment, and the shells are expanding with thespeed of light eventually filling the whole space with radiation.

For a quantitative description of the radiation process, we first need to relate apoint on a given expanding sphere to the time and location of the charge when thisparticular sphere was emitted. This time, tret, is called the retarded time and theposition of the charge at this moment — see Fig. 15.3 — the retarded position. Ifthe trajectory of the charge is given by a vector function r0(t), and we make anobservation at time t at point r, then the retarded time tret is determined from theequation

c(t − tret) = |r − r0(tret)| , (15.1)

which equates the radius of the sphere |r − r0(tret)| to the time of expansion t − tretfrom the moment of the emission multiplied by the speed of light. The retardedposition is r0(tret). Note that both tret and r0(tret), for a given path (determined by thefunction r0(t)), are functions of variables t and r.

Page 195: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

15.2 Retarded Time and Position of a Particle Moving with Constant Velocity 193

Fig. 15.3 Illustrationcharacterizing the definitionof the retarded time andretarded position. Themagenta dot shows thecurrent position of thecharge, and the green dot isits retarded position. The redcircle is the radiation sphereat time t corresponding tothe same tret

Fig. 15.4 Point chargemoving with constantvelocity along the z-axis.Point 1 (magenta)corresponds to the chargeposition at time t and point 2(green) is its position at theretarded time. Vector Rconnects point 2 with theobservation point

15.2 Retarded Time and Position of a Particle Movingwith Constant Velocity

If a charged particle emits spherical shells of electromagnetic field in the process ofacceleration, what happens if the acceleration is zero? The answer to this questionis that even in this case we can formally consider the field of a particle moving witha constant velocity as a result of emission of spherical shells. We will now give amathematical representation of this process using the expressions for the field of amoving charge from Sect. 11.1.

Consider a point charge moving with constant velocity υ along the z axis. Itstrajectory is given by

r0(t) = (0, 0, υt) . (15.2)

Let the observation point at time t have coordinates x, y, z.We denote byR the vectorconnecting the retarded position with the observation point, as shown in Fig. 15.4,R = r − r0(tret), and n = R/R. We first need to find the retarded time tret. SquaringEquation (15.1) yields

c2(t − tret)2 = (z − υtret)

2 + x2 + y2 . (15.3)

This is a quadratic equation for tret which can easily be solved. It has two solutions,one of which is an advanced solution with t < tret, the other one is our retardedsolution with t > tret. The advanced solution should be discarded because it does not

Page 196: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

194 15 Radiation and Retarded Potentials

satisfy the original Equation (15.1) which requires t > tret; for the retarded one wehave

c(t − tret) = γ2[β(z − υt) +

√(z − υt)2 + γ−2(x2 + y2)

], (15.4)

with β = υ/c. Noticing that the square root in this expression is the same as in Rin Eq. (11.5) and that c(t − tret) = R, we can try to expressR through the vector R.Let us show that

R = R(1 − β · n) = R − β · R , (15.5)

where vector β is β = (0, 0, υ/c). To prove it we take the square of Eq. (15.5),

R2 = (R − β · R)2 , (15.6)

or using coordinates:

(z − υt)2 + x2 + y2

γ2= [c(t − tret) − β(z − υtret)]2 , (15.7)

wherewehave used R = c(t − tret) andβ · R = β(z − υtret). Substituting x2 + y2 =c2(t − tret)2 − (z − υtret)2 from Eq. (15.3) we arrive at

(z − υt)2 + c2(t − tret)2 − (z − υtret)2

γ2= [c(t − tret) − β(z − υtret)]2 . (15.8)

Taking into account that γ−2 = 1 − β2 it is now a straightforward calculation tocheck that the above equation is an identity, which proves Eq. (15.5).

The potentials (11.9) for a particle moving with a constant velocity can now bewritten as

φ = 1

4πε0

q

R(1 − β · n) , A = Z0

4πβ

q

R(1 − β · n) . (15.9)

Remembering that R involves the retarded position of the particle, and noting thatβ does not depend on time, we can formally consider β as also taken at the retardedtime. With this new interpretation of Eq. (15.9), not only do we maintain consistencywith the previous solution for a charged particle moving with constant velocity, but itturns out that these expressions are valid for arbitrary motion of a point charge, evenwhen the charge is being accelerated. They are known in classical electrodynamicsunder the name of Liénard–Wiechert potentials.

Page 197: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

15.3 Liénard–Wiechert Potentials 195

15.3 Liénard–Wiechert Potentials

The Liénard–Wiechert potentials define the electrostatic potential φ and the vectorpotentialA at the observation point r at time t for arbitrarymotion of the point charge:

φ(r, t) = 1

4πε0

q

R(1 − βret · n),

A(r , t) = Z0

qβret

R(1 − βret · n). (15.10)

Here the particle velocity β should be taken at the retarded time, βret = β(tret), andwe remind the reader thatR = r − r0(tret) is a vector drawn from the retarded positionof the particle to the observation point, and n is a unit vector in the direction of R.

Differentiating the potentials (15.10) we can derive expressions for the electricand magnetic fields, E = −∇φ − ∂A/∂t , B = ∇ × A. When taking the derivativesit is important to remember that tret = tret(r, t) is a function of r and t and shouldbe differentiated over these variables. As an example, we will calculate the partialderivative ∂tret/∂t . We begin by squaring both sides of Eq. (15.1) and taking the timederivative of both sides of the equality:

∂tc2(t − tret)

2 = ∂

∂t[r − r0(tret)]2 , (15.11)

which reduces to

−2c2(t − tret)

(∂tret∂t

− 1

)= −2(r − r0(tret)) · ∂r0

∂tret

∂tret∂t

. (15.12)

Noting that ∂r0/∂tret = βret we find

∂tret∂t

= 1

1 − βret · n. (15.13)

In a similar fashion, one can calculate the spatial derivatives of tret, c∇tret = −n/(1 −βret · n). It is then straightforward to calculate the electric and magnetic fields fromthe Liénard–Wiechert potentials:

E = q

4πε0

n − βret

γ2R2(1 − βret · n)3+ q

4πε0c

n × {(n − βret) × βret}R(1 − βret · n)3

,

B = 1

cn × E , (15.14)

where βret is the acceleration (normalized by the speed of light) taken at the retardedtime, and γ = (1 − β2

ret)−1/2.

Page 198: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

196 15 Radiation and Retarded Potentials

The two terms on the right-hand side of Eq. (15.14) scale differently with distanceR. The first one decays as R−2 and is usually associated with the Coulomb fieldmodified by the relativistic motion of the particle. It is the same field as given byEq. (11.4), but expressed in terms of the vector R and the retarded velocity. Thesecond one scales as inverse distance, ∝ R−1; this is the radiation field. This term isonly present when a particle is being accelerated, β �= 0.

15.4 Radiation in the Ultra-Relativistic Limit

In the limit γ � 1 the behavior of the radiation from Eq. (15.14) is dominated by the(1 − βret · n) term in the denominator. Denoting the angle between n and βret as θ,we can write for small angles

1 − βret · n � 1 − β

(1 − θ2

2

)= 1 − β + 1

2βθ2 � 1

2

(1

γ2+ θ2

). (15.15)

The factor (1 − βret · n)−3 makes the radiation far away from the charges localizedwithin an angle 1/γ around the particle velocity, as shown in Fig. 15.5.

When the acceleration is perpendicular to the motion, βret ⊥ βret, and the obser-vation point is in the forward direction, θ = 0, the radiation field is directed oppositeto the acceleration and is equal to

Erad = − q

4πε0c4βret

γ4

R. (15.16)

The radiation fields at large angles, or traveling backwards relative to the particlevelocity, are suppressed by a factor of 1/γ4 in comparison with Eq. (15.16).

Fig. 15.5 Expandingradiation sphere of arelativistic particle. The redarea around the direction ofvelocity υ shows highintensity of the radiationfield. The magenta dotindicate the current positionof the particle (1), and thegreen dot its retardedposition (2)

Page 199: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

15.5 Retarded Potentials for an Ensemble of Particles 197

15.5 Retarded Potentials for an Ensemble of Particles

The Liénard–Wiechert potentials given by Eq. (15.10) are convenient for calculationof fields of a moving point charge. What if we are given continuous, time-dependentcurrent and charge distributions ρ(r, t) and j(r, t)? Can we integrate the Liénard–Wiechert potentials over all space to obtain the result for such a distribution?

Naively, one might think that to obtain the potential for a continuous distribu-tion one has to replace the charge q by the infinitesimal charge ρ(r′, t)d3r ′ in theelementary volume d3r ′ and integrate over the space with vector R directed from r′to the observation point r, R = r − r′. The expression q[R(1 − βret · n)]−1 on theright-hand side of Eq. (15.10) would be replaced by

∫ρ(r′, tret)d3r ′

|r − r′|(1 − βret · n), (15.17)

where n = (r − r′)/|r − r′| and the retarded time is now defined by the equation

c(t − tret) = |r − r ′| . (15.18)

This logic, however, is flawed. Indeed, if we want, using Eq. (15.17), to recover theoriginal Liénard–Wiechert potentials for a point charge we need to do the integralwith ρ(r, t) = qδ(r − r0(t)):

∫δ(r′ − r0(tret(r, r′, t)))

|r − r′|(1 − βret · n)d3r ′ = 1

(1 − υret · ∇r′ tret)

1

|r − r′|(1 − βret · n), (15.19)

where we have explicitly indicated the arguments of the retarded time and usedstandard rules for integration of the delta functions1. The gradient of the retardedtime with respect to the vector r′ can be found from Eq. (15.18), and is equal to∇r′ tret = n/c. Hence, we conclude that

∫δ(r′ − r0(tret))d3r ′

|r − r′|(1 − βret · n)= 1

|r − r0(tret)|(1 − βret · n)2, (15.20)

and we get an extra factor (1 − βret · n) in the denominator in comparison withEq. (15.10). The origin of this discrepancy can be traced to the fact that because thesecond argument in ρ is tret, the integral of the charge density over the volume is notequal to the total charge q,

∫ρ(r, tret)d3r �= q. Hence, the replacement of charge q

in the Liénard–Wiechert potentials by ρ(r′, tret)d3r ′ with the subsequent integrationover the space cannot be justified.

1An integral of a three-dimensional delta function∫

δ( f (x, y, z), g(x, y, z), h(x, y, z))dxdydz isequal to |J |−1 where J is a Jacobian of the transformation from x, y, z to f, g, h taken at the pointwhere f = g = h = 0.

Page 200: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

198 15 Radiation and Retarded Potentials

It is now clear, however, how we can correct for this extra factor in the transitionfrom the point charge to a continuous distribution. We just need to use the rule

q

1 − βret · n→ ρ(r′, tret)d

3r ′,

which gives

φ(r, t) = 1

4πε0

∫ρ(r′, tret)

|r − r′| d3r ′ ,

A(r, t) = Z0

4πc

∫j(r′, tret)

|r − r′| d3r ′ . (15.21)

These integrals are called the retarded potentials. They descrive the radiation fieldin free space of a system of charges represented by continuous distributions of thecharge density ρ and the current density j.

Worked Examples

Problem 15.1 Show that

∇tret = − nc(1 − n · βret)

.

Find ∂R/∂t and ∇ R. The operator ∇ here is understood as x∂/∂x + y∂/∂y +z∂/∂z.

Solution: To simplify notation, for ∇tret, we will only calculate the x componentfor now, using x · ∇ = ∂/∂x . We square both sides of Eq. (15.1) and differentiate itwith respect to x ,

−2Rc∂tret∂x

= 2 [r − r0(tret)] ·(x − dr0

dtret

∂tret∂x

)

= 2

(R · x − R · cβret

∂tret∂x

),

where we used the notation R = r − r0(tret) and R = |R| = c(t − tret). Solving thisequation for ∂tret/∂x we find

∂tret∂x

= −x · nc(1 − β · n) .

Page 201: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

15.5 Retarded Potentials for an Ensemble of Particles 199

The vertical and longitudinal gradients are calculated analogously but with x · nreplaced with y · n and z · n respectively. Therefore

∇tret = − nc(1 − βret · n)

.

The derivatives ∂R/∂t and ∇ R are now easily found:

∂R

∂t= c

∂t(t − tret) = c

(1 − ∂tret

∂t

)

= c

(1 − 1

1 − βret · n)

= cβret · n

1 − βret · n,

and

∇ R = −c∇tret = n1 − βret · n

.

Problem 15.2 Find the angular dependence of the radiation electric field Erad fora relativistic particle assuming that βret is collinear to βret. Consider small anglesbetween n and βret.

Solution: The radiation field is given by the second term in the expression for theelectric field in Eq. (15.14),

Erad = q

4πε0c

n × {(n − β) × β}R(1 − β · n)3 , (15.22)

where to simplify the notation we dropped the subscript ret. Choosing the coordinatesystem with axis z directed along the velocity we have β = zβ, β = zβ and n =xnx + yny + znz . In the spherical coordinate system with the angles θ and φ we canalso write nx = sin(θ) cos(φ), ny = sin(θ) sin(φ) and nz = cos(θ). For a relativisticparticles we can use β ≈ 1 − 1/2γ2, and for a small angle θ we replace sin(θ) ≈ θand cos(θ) ≈ 1 − θ2/2 in the expressions for n. Using these approximations, thenumerator and the denominator in Eq. (15.22) become

n × {(n − β) × β} ≈ β(xnx + yny)

(1 − β · n)3 ≈ 1

8(θ2 + γ−2)3 ,

and we obtain for the electric field

Erad = 2qβ

πε0cR

θ

(θ2 + γ−2)3[x cos(φ) + y sin(φ)] .

This field has a radial polarization and vanishes in the forward direction, θ = 0.

Page 202: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

Chapter 16Dipole Radiation and Scatteringof Electromagnetic Waves

In this chapter we consider scattering of an electromagnetic wave on a free chargedparticle — a process often referred to as Thomson scattering. This interaction inducesa damping force on the charged particle called the radiation reaction force, whichmaintains the energy balance. Additionally, the radiation exerts another force in thedirection of propagation called the light pressure, which is responsible for conservingthe combined momentum of the particles and fields. We also briefly discuss inverseCompton scattering off a point charge moving with a relativistic velocity.

16.1 Dipole Radiation of a Linear Oscillator

The general formula (15.14) for the electromagnetic field of a point charge canbe considerably simplified if the charge is moving with a non-relativistic velocity,β � 1. Neglecting in (15.14) βret in comparison with vectors of unit length, for theradiation field we obtain

E = q

4πε0cRn × (

n × βret

),

B = 1

cn × E = − Z0q

4πcRn × βret . (16.1)

The distance R in this equation is defined as the distance from the retarded positionof the particle to the observation point r, R = | r − r0(tret)|, and n is the unit vectorin this direction.

An important example of non-relativistic radiation is the harmonic oscillator forwhich v(t) = Re(v0e−iωt ), where ω is the oscillation frequency and v0 � c is thevelocity amplitude. Due to the linearity of Eq. (16.1) involved in the calculation ofthe fields, it is convenient to work with complex numbers omitting the symbol forthe real value, with the understanding that all physically meaningful quantities are

© Springer International Publishing AG, part of Springer Nature 2018G. Stupakov and G. Penn, Classical Mechanics and Electromagnetismin Accelerator Physics, Graduate Texts in Physics,https://doi.org/10.1007/978-3-319-90188-6_16

201

Page 203: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

202 16 Dipole Radiation and Scattering of Electromagnetic Waves

obtained by taking the real parts of these complex quantities. The equation for thevelocity then reads

v(t) = v0e−iωt . (16.2)

Integrating the velocity over time gives the position of the oscillator r0(t),

r0(t) = i

ωv0e

−iωt , (16.3)

where we have chosen the constant of integration so that the averaged over timeposition of the particle is at the origin of the coordinate system. From the conditionv0 � c it follows that r0ω/c � 1, or

r0 � λ , (16.4)

where the reduced wavelength λ = c/ω = λ/2π, with λ being the wavelength of theelectromagnetic radiation with frequency ω.

Having calculated the particle motion, we can now find its radiation. The radiationfield becomes dominant at distances larger than a wavelength and the size of thesource of the radiation, so we have

r0 � λ � r . (16.5)

This leads us to the dipole approximation, where we can approximate R ≈ r inEq. (16.1) by neglecting r0(t) in comparison with r. Correspondingly, for the retardedtime we have

tret = t − r

c. (16.6)

By evaluating the acceleration at this value for tret, we obtain for the magnetic fieldof the radiation

B = Z0qiω

4πcrn × β0e

−iωt+ikr , (16.7)

where k = ω/c. This is a spherical electromagnetic wave, whose amplitude decayswith distance as 1/r .

The intensity of the radiation is the energy flow given by the Poynting vector (A.7).This vector is directed along n and the time average of this component, which wedenote by S, is given by the same equations as the Poynting vector in the planeelectromagnetic wave (13.8):

Page 204: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

16.1 Dipole Radiation of a Linear Oscillator 203

Fig. 16.1 The three-dimensional angulardistribution of the dipoleradiation. The diagram isaxially symmetric withrespect to the vector β0. Thelength of the blue vectors inthe cutout is proportional tothe radiation intensity in thedirection of the vectors

S = 1

2Z0|E |2 = c2

2Z0|B|2

= Z0

32π2

q2ω2

r2|n × β0|2 . (16.8)

The Poynting vector gives the energy flow through a unit area. Another convenientquantity is Sr2 which describes the power radiated per unit solid angle. We denotethis quantity by dP/d�. We have

dPd�

= Sr2 = Z0

32π2q2ω2β2

0 sin2 ψ , (16.9)

where ψ is the angle between the direction of the observation n and β0. The angulardistribution of the dipole radiation is shown in Fig. 16.1. Integrating dP/d� overthe solid angle gives the total power of radiation:

P =∫

dPd�

d�

= Z0

32π2q2ω2β2

0

∫ π

0sin2 ψ · 2π sin ψ dψ

= Z0

32π2

(8π

3

)q2ω2β2

0 . (16.10)

16.2 Radiation Reaction Force

Because the charge is losing energy to radiation, it should feel a force that, onaverage, works against the velocity to keep the energy balance in the system. Indeed,such a force exists, and is called the radiation reaction force. If the charge executes

Page 205: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

204 16 Dipole Radiation and Scattering of Electromagnetic Waves

free oscillations with the frequency ω, this force would cause the amplitude of theoscillations to decay with time and would eventually terminate the oscillations withall of the kinetic energy converted into radiation. If the oscillations are driven by anexternal force at a constant amplitude, then this force is responsible for taking theenergy from the external source and supplying it to the radiated waves.

Let us calculate this force, f r , equating the radiated power P to the work of theforce with the negative sign,

P = −〈f r · v〉 , (16.11)

where the angular brackets indicate averaging over time. We take the expression forthe velocity Eq. (16.2) in real form

v = v0 cos ωt , (16.12)

and assume that f r acts in the direction opposite to the velocity,

f r = −Av . (16.13)

We then have

〈f r · υ〉 = −1

2Av2

0 . (16.14)

Equating this expression to P given by Eq. (16.10), we find A and the force,

f r = − Z0

q2ω2

c2v . (16.15)

Because for an oscillator ω2v = −v we can also write

f r = 1

6πε0

q2

c3v = 2

3

rcm

cv , (16.16)

where rc is the classical radius,

rc = 1

4πε0

q2

mc2; (16.17)

for electrons, rc = 2.8 × 10−15 m. The expression (16.16), as it turns out, is moregeneral than our derivation assumes — it is valid for arbitrary nonrelativistic motionof a point charge.

As we emphasized above, the radiation reaction force is responsible for the energybalance in the radiation process. There is one subtle issue with the application of thisforce [1, 2] which is clearly seen if we equate this force to the particle acceleration,

Page 206: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

16.2 Radiation Reaction Force 205

(a) (b)

Fig. 16.2 a In Thomson scattering a plane electromagnetic wave illuminates a charge at rest. Dueto oscillations of the charge in the incident wave, secondary waves will be radiated shown by bluewiggly arrows.bWhen the charge is moving towards the wave with relativistic velocity, the radiationis localized in a narrow angle ∼ 1/γ

υ = f r/m, in the absence of external fields. It is easy to solve this differentialequation and find an exponentially growing solution v ∝ exp(2ct/3rc), which isclearly unphysical. The origin of this solution is in the infinite Coulomb energy ofa point charge. To avoid such kinds of singularities one has to use Eq. (16.16) onlyin situations where a particle is accelerated by external electromagnetic fields andsubstitute for υ on the right-hand side of this equation the time derivative of theacceleration caused by these fields. In other words, it cannot be applied in a fullyself-consistent manner, but has to be evaluated on the basis of all forces excludingthe radiation reaction force itself1.

In Chap. 22 we will consider the implications of this force for the case of syn-chrotron radiation of a relativistic particle.

16.3 Thomson Scattering

As an application of the results of Sect. 16.1, we now consider Thomson scattering— radiation of a point charge caused by motion in a plane electromagnetic wave.We assume that an electron that was initially at rest is illuminated by a plane elec-tromagnetic wave with frequency ω. This electron starts to oscillate in the wave andto radiate secondary electromagnetic waves (Fig. 16.2).

We first need to find the electron motion in the incident wave. We will assume thatthis wave is not too strong so that the electron velocity in the wave is nonrelativistic,v � c. The field in the wave is given by Eq. (13.6) and the equation of motion forthe electron in complex number notation is

mdv

dt= qE0e

−iωt+ik·r , (16.18)

1This holds true even though in some cases the radiation reaction force for a relativistic particle canbe larger than the Lorentz force from an external electromagnetic field [3].

Page 207: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

206 16 Dipole Radiation and Scattering of Electromagnetic Waves

(here we have assumed the phase φ0 = 0). The magnetic force in this equation isomitted because in the limit v � c it is much smaller than the electric one. Assumingthat the electron is located near the origin of the coordinate system, r ≈ 0, we dropthe term ik · r on the right-hand side of Eq. (16.18),

mdv

dt= qE0e

−iωt . (16.19)

The smallness of the parameter k · r is justified by the inequality (16.4) valid foroscillations with a non-relativistic velocity v. Integration over time gives

v = iq

mωE0e

−iωt . (16.20)

Comparing this expression with Eq. (16.2) we relate v0 to the electric field in theplane wave,

v0 = iq

mωE0 . (16.21)

Using this equation, all the results of Sect. 16.1 can now be expressed in terms of theamplitude of the electric field, E0. For examples, the total radiation power (16.10)becomes

P = Z0

32π2

(8π

3

)q4E2

0

m2c2. (16.22)

From the condition v � c it follows that

a ≡ qE0

mωc� 1 , (16.23)

where we have introduced the dimensionless parameter a, which characterizes thestrength of the electromagnetic field. This is a very important condition, which wewill meet again, in a different context, in Chap. 19.

If we divide the radiated power by the average energy flow in the incident wave,we obtain a quantity that has dimension of length squared. This quantity can beinterpreted as a scattering cross section, and is called the Thomson cross section:

σT = PE2

0/2Z0=

(1

4πε0

)2 8πq4

3m2c4= 8π

3r2c . (16.24)

In Chap. 19 we will need the intensity of the radiation written in terms of sphericalcoordinates, in which the wave propagates in the z direction and the electric field isdirected along x . We introduce the polar angle θ measured relative to the z axis and

Page 208: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

16.3 Thomson Scattering 207

Fig. 16.3 The sphericalcoordinate system. Vector nshown in red is defined by itstwo angles, θ and φ. Thewave propagates along the zaxis, and the electric field inthe wave is directed alongthe x axis

the azimuthal angle φ measured in the x − y plane, see Fig. 16.3. In terms of thesevariables the (x, y, z) components of the propagation direction n are given by

n = (sin θ cos φ, sin θ sin φ, cos θ) , (16.25)

and with E0 = (E0, 0, 0) we find

|n × E0|2 = E20(1 − sin2 θ cos2 φ) . (16.26)

Returning to the definition dP/d� = Sr2 and using Eq. (16.8) with Eqs. (16.21)and (16.26) we obtain

dPd�

= Z0

32π2

q4E20

m2c2(1 − sin2 θ cos2 φ) . (16.27)

While Thomson scattering is a classical phenomenon and as such is perfectlywell described by classical electrodynamics2, it is instructive to convert the radiationpower (16.22) into quantum language. In quantum theory the electromagnetic fieldis represented by photons, so we want to calculate the number of photons emittedby the charge per unit time. This number, Np, is obtained by dividing the radiationpower P by the photon energy �ω,

Np = P�ω

= 1

12πε0

q2E20

m2c2ω2ωq2

�c

= 1

3a2ωα , (16.28)

where a is defined by Eq. (16.23) and α is the fine structure constant,

2We will estimate the conditions when quantum effects start to play a role in electromagneticscattering in the next section.

Page 209: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

208 16 Dipole Radiation and Scattering of Electromagnetic Waves

α = 1

4πε0

q2

�c≈ 1

137. (16.29)

If we multiply Eq. (16.28) by 2π/ω, we find that the number of photons scattered byan electron in one period of oscillations is approximately 2αa2.

16.4 Light Pressure

The quantum picture of Thomson scattering helps to understand the important phys-ical effect of light pressure. The photons of the incident wave within the Thomsoncross section are scattered off in different directions. The incident photons carrymomentum in the direction of k, that is the direction of the wave propagation.Because the scattered photons are equally distributed between the forward and back-ward directions, the averaged momentum of the scattered radiation is zero. Hencethe initial momentum of the incident photons is transferred to the scatterer, whichmeans that there is a force, f p, exerted on the charge in the direction of the incidentwave propagation.

It takes only one more step to calculate this force from the quantum viewpoint.We need to multiply the number of photons scattered per unit time (16.28) by thephoton momentum �k,

f p = �k Np = Pc

= σTE2

0

2cZ0. (16.30)

It is clear that this force is directed along k. Note that although we used quantumtheory in the derivation, our final result does not involve the Planck constant andhence is classical, and can also be derived in classical electromagnetism.

In the classical derivation of f p we first need to modify the equation of motion(16.19) by adding on the right-hand side the radiation reaction force (16.16),

mdv

dt= qE0e

−iωt + 2

3

rcm

c

d2v

dt2. (16.31)

With this modification, the solution Eq. (16.20) acquires an additional term υ1,

v = iq

mωE0e

−iωt + v1 . (16.32)

We expect that the correction v1 is small and substituting this expression intoEq. (16.31) we neglect d2v1/dt2 on the right-hand side, which then gives

mdv1

dt= −2

3

iqrcω

cE0e

−iωt . (16.33)

Page 210: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

16.4 Light Pressure 209

The solution of this equation is

v1 = 2

3

qrcmc

E0e−iωt . (16.34)

The final step in the derivation is to calculate the average over time magnetic forceqv1 × B which arises from the cross product of the velocity v1 and the magneticfield in the incident wave, B = (k/ck) × E0e−iωt . To average the force, we have totake the real parts of v1 and B, with the result

q〈Re υ1 × ReB〉 = 1

3

q2rcmc2k

E0 × (k × E0) = 1

3

q2rcE20

mc2

kk

. (16.35)

We see that this force acts in the direction of the wave propagation. The net forceis associated with a small change in the phase offset between the velocity and theelectromagnetic fields, from π/2 to π/2 − 2rcω/3c. It is easy to check that themagnitude of the force is exactly equal to the expression (16.30) obtained in thequantum derivation.

We conclude this section with an estimate of when the classical theory cannot beapplied to the scattering. If the photon energy in the incident electromagnetic wave�ω is so large that it is comparable to the rest energy mc2, �ω ∼ mc2, the classicaldescription breaks down. In this case, a part of the incident photon energy goes intothe recoil momentum of the charge, and the frequency of the scattered photons ω′becomes smaller than the incident frequency, ω′ < ω. The process in which scatteringleads to a change of frequency is usually referred to as Compton scattering.

16.5 Inverse Compton Scattering

Thomson scattering, analyzed in Sect. 16.3, can be generalized to the case of a mov-ing scatterer. In practice, the most interesting situation occurs when the scatterer isrelativistic and is moving towards the wave as shown in Fig. 16.2b. This is usually ref-ered to as inverse Compton scattering. The process is most easily calculated througha Lorentz transformation from the lab frame to the particle frame of reference wherethe average velocity of the scatterer is equal to zero and we can use the results ofSect. 16.3. The radiation in that reference frame should then be transformed back tothe lab frame using the Lorentz transformations for the fields. In this section, we willlimit our analysis to the frequency of the scattered radiation in the lab frame when vis close to the speed of light, γ 1.

Let us choose the coordinate system with the z axis directed along the velocityv and denote the incident wave frequency in the lab frame by ω0. The wave prop-agates in the negative-z direction. In the particle frame of reference, the frequencyω′ of the incident wave is different from ω0 — it can be obtained from the Lorentztransformation (B.15) setting θ = π,

Page 211: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

210 16 Dipole Radiation and Scattering of Electromagnetic Waves

ω′ = 2γω0 . (16.36)

This is also the frequency of the scattered radiation in the particle frame. To transformit back to the lab frame we will assume small angles θ (θ is the angle between ascattered photon and the z axis) and use Eq. (B.16),

ω = 4γ2ω0

1 + γ2θ2. (16.37)

We see that the frequency of the scattered radiation now depends on the direction ofobservation, and the frequncy spectrum has a maximum frequency equal to 4γ2ω0,which for γ 1 is much larger than the incident wave frequency ω0. It follows fromEq. (16.37) that the high frequencies, ω ∼ γ2ω0, are localized within a narrow rangeof angles θ ∼ 1/γ around the direction of the velocity v.

We know that in the beam frame the photons are radiated more or less in alldirections. The angles in the lab frame θ are related to the angles in the beam framethrough Eq. (B.12). In the limit γ 1, this equation maps most of the interval 0 <

θ′ < π into a narrow range of angles θ ∼ 1/γ around the velocity v (to be moreprecise, for θ to be ∼ 1/γ, the angle θ′ should not be very close to π, π − θ′ 1/γ).This means that almost all photons in the lab frame propagate within the angle 1/γin the direction of the beam motion. The number of photons will be the same as inthe beam frame, but their energy in the lab frame is ∼ γ2 times larger than the energyof the incident photons. The fact that the dominant part of the scattering is localizedwithin the angle ∼ 1/γ is a manifestation of the general properties of the radiationof relativistic particles discussed in Sect. 15.4.

Worked Examples

Problem 16.1 Prove that the ratio qE0/mωc is a Lorentz invariant — it does notchange under the Lorentz transformation (in other words, it is the same in any coor-dinate system moving relative to the laboratory reference frame). Assume that theframes move in the direction of propagation of the wave.

Solution: It is sufficient to establish the invariance of the ratio E0/ω. FromAppendix B we know that when performing a Lorentz boost in the same directionas the radiation propagation the frequency becomes ω′ = γ(1 − β)ω.

For a radiation field, where B = (1/ω)k × E, the electric field transforms as

E = γ(E′ − v × B′) = γ(1 + β)E′ .

This gives E′ = E/γ(1 + β) = γ(1 − β)E. So we find E ′0/ω

′ = E/ω, showing thatthe ratio is Lorentz invariant.

Page 212: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

16.5 Inverse Compton Scattering 211

Problem 16.2 Consider scattering of an electromagnetic wave on a charge q that isattached to an immobile point through a spring, and can oscillate with the frequencyω0. Find the scattering cross section as a function of frequency of the incident waveω assuming that the electric field in the wave is directed along the line of oscillationsof the charge.

Solution: The charge motion is determined by

x + ω20x = qE0

me−iωt+ik·r ≈ qE0

me−iωt .

We will look for solutions of the form x = Ae−iωt . Substituting into the above equa-tion we find

A = qE0

m(ω20 − ω2)

.

The charge velocity is

υx = −iωx = iqE0

ω2

(ω2 − ω20)e−iωt .

The power radiated, P , is proportional to υ2x . Comparing our expression for υx to

that of a free particle, Eq. (16.20), we see that the spring modifies the velocity by theterm ω2/(ω2 − ω2

0), so our power is modified by ω4/(ω2 − ω20)

2, and the new crosssection is

σ = PE2

0/2Z0= σT

(ω2

ω2 − ω20

)2

.

When the frequency ω approaches ω0, the cross section becomes large, σ σT .

Problem 16.3 In the derivation of the light pressure we neglected the term qv × Bwhere v is given by the real part of Eq. (16.20). Show that 〈v × B〉 = 0.

Solution: The velocity v ∝ Re (iE0e−iωt ) and B ∝ Re[n × E0e−iωt ]. When weaverage the product over time, we first take the real part of υ and B to find

〈v × B〉 ∝ 〈sin(ωt) cos(ωt)〉 = 0 .

Problem 16.4 It has been proposed to use Thomson scattering in a compact electronring as a source of intense X-ray radiation (e.g., Phys. Rev. Lett., 80, 976, (1998)).Consider the following parameters of such a system: the electron energy in the ringis 8 MeV, the number of electrons in the bunch is Ne = 1.1 × 1010, the laser energyis 20 mJ, the laser pulse length is 1 mm, and the laser is focused to the spot size 25micron. Estimate the number of photons from a single collision of the laser pulsewith the electron beam.

Page 213: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

212 16 Dipole Radiation and Scattering of Electromagnetic Waves

Solution: The photon flux of the laser beam at 800 nm (hν = 2.5 × 10−19 J), withlaser energy EL = 20 mJ and waist w = 25 µm is

EL

πw2Lhν

= 2 × 10−2

π × 252 × 10−12 × 2.5 × 10−19photons/m2

= 4 × 1025 photons/m2 .

The total scattering cross section of all electrons in the bunch is NeσT = 6.7 × 10−19

m2. So a single interaction scatters 3 × 107 photons.

Problem 16.5 An electromagnetic wave with frequency ω and amplitude of theelectric field E0 occupies a volume with dimensions Lx × Ly × Lz . It propagatesalong the z axis with fields Ex = cBy . Using the results of Problem 13.2, find theelectromagnetic energy W of the wave in the lab frame and the energy W ′ in a frameK ′ moving with velocity v along the z axis relative to the lab frame. Show thatW/ω = W ′/ω′, where ω′ is the frequency of the wave in K ′.

Solution: The averaged energy density w in a plane wave is equal to w = ε0E20/2.

Comparing this expression with the Poynting vector (13.8) we see that w = S/c. Wecan now write the energy in the lab frame as W = wLx L yLz = SLx L yLz/c. To findthe energy in the moving frame we need to transform S and Lz .

The electromagnetic field has the same number of periods in each frame, soLz must transform as the wavelength λ, which transforms as the inverse of ω′ =γ(1 − β)ω as in (B.15). For the radiation field, we know from Eq. (B.17) that Ex =γ(1 + β)E ′

x and By = γ(1 + β)B ′y . We then find

W ′

ω′ = S′L ′x L

′y L

′z

cω′

= SLx L yLz

cγ4(1 − β)2(1 + β)2ω

= SLx L yLz

(1 − β2)2

(1 − β)2(1 + β)2

= W

ω.

Since the ratio W/ω is proportional to the number of photons, its invariance tells usthat the number of photons in both frames is the same.

References

1. W. Panofsky, M. Phillips.Classical Electricity andMagnetism, 2nd edn. (Addison-Wesley 1962)2. J.D. Jackson, Classical Electrodynamics, 3rd edn. (Wiley, New York, 1999)3. L.D. Landau, E.M. Lifshitz. The classical theory of fields, volume 2 of Course of Theoretical

Physics, 4th edn. (Elsevier Butterworth-Heinemann, Burlington MA, 1980) (translated fromRussian)

Page 214: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

Chapter 17Transition and Diffraction Radiation

Transition radiation occurs when a moving charged particle crosses a boundary be-tween two media with different electrodynamic properties. In its simplest form, whichis commonly used in experiments, transition radiation is generated by sending a beamthrough a metallic foil. In this chapter, we will derive the spectrum and angular dis-tribution of the transition radiation when a particle crosses a foil at normal incidence.We will also discuss radiation generated by the beam when it passes through a holein a metal foil — the so-called diffraction radiation.

17.1 Transition Radiation

We will calculate the transition radiation generated when a plane metal surface is hitby a point charge moving with a constant velocity v in the direction perpendicular tothe surface, as shown in Fig. 17.1a. We choose the coordinate system with the originlocated at the entrance to the metal in such a way that the particle is moving alongthe z axis in the positive direction. The metal occupies the region z > 0 with z = 0being the metal boundary.

To find the electromagnetic field at z < 0 one has to solve Maxwell’s equationwith the charge and current density corresponding to the moving point charge q,ρ = qδ(z − vt), j = vρ, at t < 0. At time t = 0 the charge q enters the metal andgets shielded by the charges inside the metal; there are no charges at t > 0 in theupper half space. Assuming the perfect conductivity of the metal, we need to solveMaxwell’s equations with the boundary condition of zero tangential electric field atz = 0. The solution is greatly simplified if one uses the method of image charges. Inthis method, one replaces the metal with an image charge of the opposite sign, movingwith velocity v in the opposite direction, as shown in Fig. 17.1b. In what follows,we will refer to the original charge q by index 1, and to the image charge −q byindex 2. Using Eq. (11.4) from Chap. 11, it is easy to verify that in the plane z = 0, thecomponents of the electric field tangential to the metal surface (that is, perpendicular

© Springer International Publishing AG, part of Springer Nature 2018G. Stupakov and G. Penn, Classical Mechanics and Electromagnetismin Accelerator Physics, Graduate Texts in Physics,https://doi.org/10.1007/978-3-319-90188-6_17

213

Page 215: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

214 17 Transition and Diffraction Radiation

(a) (b)

Fig. 17.1 A charge moving perpendicular to the metal surface (a), and the two charges in themethod of image charges (b). The observation point O is a distance r from where the charge willstrike the surface, and the line connecting these two points forms an angle θ with the directionnormal to the surface. Panel (b) also shows the unit vectors n1 and n2 from Eq. (17.1)

to the z direction) of the two charges cancel each other and the boundary conditionsEx = Ey = 0 are automatically satisfied. At time t = 0 the two charges meet at pointz = 0 where they annihilate so that at time t > 0 there are no charges in the system.

While we assume that the metal occupies the half space z > 0, we would obtain thesame electromagnetic field in the region z < 0 in the case of a metal foil occupyingthe region 0 < z < h, where h is the thickness of the foil. This follows from the factthat the boundary condition of zero tangential field at z = 0 does not change for thefoil (again, assuming the perfect conductivity of the metal). Hence our results forthe transition radiation will be also valid for the metal foil.1 In either case, the fieldsfrom the real particle are shielded by the metal for t > 0.

To calculate the radiation field, we need to find the vector potential A(r, t) at theobservation point r at time t . The trajectories of the two charges, 1 and 2, for t < 0are given by the equations r1(t) = (0, 0, vt) and r2(t) = (0, 0,−vt) respectively.Because the charges are moving with a constant velocity, we can use Eq. (15.9) for thepotentials. The retarded times for both particles, t (1)

ret (r, t) and t (2)ret (r, t), satisfy the

equations c(t − t (1)ret ) = |r − r1(t

(1)ret )| and c(t − t (2)

ret ) = |r − r2(t(2)ret )|, respectively

(see Eq. (15.1)), again for t (1)ret < 0 and t (2)

ret < 0. Note that the moment t (1)ret = t (2)

ret = 0corresponds to t = r/c — this is a spherical shell in the upper half space expandingwith the speed of light, r = ct . Because at t (1)

ret = t (2)ret = 0 the charges collide and

disappear, the potentials are zero for t (1)ret , t (2)

ret > 0, or equivalently for t > r/c. Hencethe vector potential A can be obtained as a combination of expressions (15.9) withthe step function h(r/c − t) which guarantees the zero value of A for t > r/c:

1If the charge passes through the foil and exits to the region z > h, there is transition radiation inthis region of space as well. It can be solved by the same method of image charges.

Page 216: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

17.1 Transition Radiation 215

A = Z0

q

R1(t(1)ret )(1 − β · n1)

+ (−β)(−q)

R2(t(2)ret )(1 + β · n2)

)h

(rc

− t)

,

(17.1)

where

R1(t) =√

(z − vt)2 + x2 + y2 ,

R2(t) =√

(z + vt)2 + x2 + y2 , (17.2)

and h is equal to one for a positive argument and zero otherwise.Let us now assume that the radiation is observed at a large distance from the

origin: we can then neglect the difference between n1 and n2 and assume that theyboth are equal to the unit vector n directed from the origin of the coordinate sys-tem to the observation point. The magnetic field of the radiation can be calculatedusing Eq. (13.5):

B = −1

cn × ∂A

∂t. (17.3)

When we differentiate (17.1) with respect to time, we only need to differentiate thefunction h — differentiating R1 and R2 would give a field that decays faster than 1/r ,which is simply the static field of each moving charge. The result of the differentiationis

B = Z0

q

cδ(rc

− t)(

1

R1(0)(1 + β cos θ)+ 1

R2(0)(1 − β cos θ)

)n × β ,

(17.4)

where θ is the angle between n and the normal to the metal, see Fig. 17.1. Becauseof the delta function factor, the values of R1 and R2 in this equation are taken at theretarded times t (1)

ret = t (2)ret = 0:

R1(0) = R2(0) =√z2 + x2 + y2 = r , (17.5)

which simplifies the formula for the magnetic field to yield

B = Z0

2q

rcδ(rc

− t) n × β

1 − β2 cos2 θ. (17.6)

We see that the radiation field consists of an infinitely thin spherical shell expandingaway from the point where the charge enters the metal.

Page 217: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

216 17 Transition and Diffraction Radiation

17.2 Fourier Transformation of the Radiation Fieldand the Radiated Power

When a radiation field is a pulse of finite duration, as derived in the preceding section,it makes sense to try to calculate the total radiation energy and its spectral distribution,or spectrum. Let us denote by B(t) the radiation magnetic field as a function of timeat a distance r from the radiation source. This field also depends on the directionof the vector n and decays with distance as 1/r . The energy radiated per unit solidangle, dW/d�, is given by the time integral of the product of the Poynting vectorS = |E × H| and the squared distance r2:

dWd�

= r2∫ ∞

−∞dt S(t) = c2r2

Z0

∫ ∞

−∞dt B(t)2 . (17.7)

In the last equation we have used the fact that in the radiation zone the field can beapproximated locally by a plane wave in which E = cB and the electric and magneticfields are perpendicular to each other.

The spectrum is obtained by taking the Fourier transform of the field and repre-senting the radiated energy as an integral over the frequencies ω. Using Parseval’stheorem from Fourier analysis we express the time integral of the square of the mag-netic field through the frequency integral of the square of the absolute value of itsFourier transform, ∫ ∞

−∞dt B(t)2 = 1

∫ ∞

−∞dω |B(ω)|2

= 1

π

∫ ∞

0dω |B(ω)|2 , (17.8)

where

B(ω) =∫ ∞

−∞dt B(t)eiωt . (17.9)

At this point, we introduce the energy radiated per unit frequency interval per unitsolid angle, d2W/dωd�, as

d2Wdωd�

= c2r2

πZ0|B(ω)|2 , (17.10)

so that the total energy radiated per unit solid angle can be represented as

dWd�

=∫ ∞

0dω

d2Wdωd�

. (17.11)

Page 218: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

17.2 Fourier Transformation of the Radiation Field and the Radiated Power 217

Fig. 17.2 Angulardistribution of the transitionradiation for a relativisticparticle

To find the spectrum of the transition radiation with the magnetic field given byEq. (17.6) we first make the Fourier transform,

B(ω) =∫ ∞

−∞dt B(t)eiωt = Z0

2qeiωr/c

rc

n × β

1 − β2 cos2 θ. (17.12)

Since the Fourier transform of the delta function is equal to one, the spectrum does notdepend on the frequency. Again, this is one of the consequences of our assumptionof the perfect conductivity of the metal — in reality at very high frequencies theapproximation of the perfect conductivity fails and the intensity of the radiationdecreases. For the angular distribution of the spectral power we have

d2Wdωd�

= c2r2

πZ0|B(ω)|2 = Z0q2

4π3

β2 sin2 θ

(1 − β2 cos2 θ)2. (17.13)

It follows from this equation that for a relativistic particle the dominant part of theradiation goes in the backward direction, θ ≈ 0. Using β2 = 1 − γ−2 and approxi-mating sin θ ≈ θ and cos2 θ ≈ 1 − θ2 we find

d2Wdωd�

≈ Z0q2

4π3

θ2

(γ−2 + θ2)2. (17.14)

The plot of this function is shown in Fig. 17.2 — the maximum intensity of theradiation is emitted at angle θ = 1/γ, while the intensity is zero at θ = 0.

One can integrate Eq. (17.13) over all angles to find the spectrum of the transitionradiation,

dWdω

= 2π

∫ π/2

0sin θdθ

d2Wdωd�

= Z0q2

4π2

[(1

β+ β

)arctanh(β) − 1

]. (17.15)

Page 219: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

218 17 Transition and Diffraction Radiation

Because the spectrum of the radiation does not depend on frequency, the integralover ω from zero to infinity diverges which formally means that the total radiatedenergy is infinite. This is in agreement with the fact that for the magnetic fieldproportional to the delta function of time, the integral (17.7) diverges as well. Inreality, as was already mentioned above, the energy is finite because at the very highfrequencies metals lose their capability of being almost perfect conductors, and thespectrum eventually tends to zero. In the time domain, this means that the particledoes not instantaneously become shielded by the metal as it crosses z = 0, but insteadtakes some time for the shielding to be complete.

The transition radiation is often used in accelerators for measuring the transversesize and position of the beam by observing an image of the radiation produced whenit strikes a metal foil in its path.

17.3 Diffraction Radiation

An interception of a beam with a metal foil to generate the transition radiation eitherdestroys the beam or deteriorates its properties, making this technique unsuitablefor many applications. In some cases one would like to generate radiation withoutstrongly perturbing the beam. This can be achieved if the beam passes through asmall hole in a metal foil — it then generates the so-called diffraction radiation,which can serve for diagnostic purposes while avoiding significant disruption of theelectron beam.

A complete electromagnetic solution of the diffraction radiation in the generalcase is extremely complicated and is beyond the scope of this book. A simplifiedtreatment [1] that uses methods from the theory of diffraction can be carried out for around hole in the limit when the radiation wavelength is much smaller than the holeradius a. In the limit γ � 1, and for small observation angles, θ � 1, this analysisresults in the following formula for the angular spectral distribution:

d2Wdωd�

≈ Z0q2

4π3

θ2

(γ−2 + θ2)2F

(ωaθ

c,ωa

), (17.16)

where

F(x, y) =[y J2(x)K1(y) − y2

xJ1(x)K2(y)

]2

, (17.17)

with J1,2 the Bessel functions and K1,2 the modified Bessel functions. In the limitx → 0 and y → 0, corresponding to the zero size of the hole a, the function F(x, y)tends to 1 and we recover the result for transition radiation (17.14). This also meansthat, at a given frequency ω, the hole has a small effect on the transition radiation ifa � cγ/ω. In Fig. 17.3 we plot the spectral intensity of the radiation as a functionof the angle θ for several values of the parameter aω/cγ.

Page 220: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

17.3 Diffraction Radiation 219

Fig. 17.3 Angulardistribution of the diffractionradiation for various valuesof the parameter aω/cγ. Thedashed line shows the limita → 0, corresponding to thecase of transition radiation

Fig. 17.4 Transitionradiation with foil tilted at45◦

In contrast to transition radiation, the total energy in diffraction radiation is finiteeven in the perfect conducting limit. It can be obtained by integrating Eq. (17.16)over the frequency ω and the solid angle θ.

Worked Examples

Problem 17.1 The usual experimental setup for an optical transition radiation (OTR)diagnostic is shown in Fig. 17.4: the beam passes through a metal foil tilted at theangle 45 degrees relative to the beam orbit. Show that in this case the radiationpropagates predominantly in the direction perpendicular to the orbit. How can thisproblem be solved using the method of image charges?

Solution: The image charge for t < 0 (before the charge hits the foil) has charge−q and moves downwards vertically. At the point of contact, it looks like the im-age charge annihilates with the real charge. The disappearance of the image charge

Page 221: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

220 17 Transition and Diffraction Radiation

produces the radiation propagating vertically in the direction that the image particlewas moving in.

Instead of appealing to image charges, one can equivalently consider the transitionradiation as being reflected off of the foil. Both descriptions are only accurate forfrequencies sufficiently low that the foil can be treated as a good conductor.

If the particle passes completely through the foil then there will be another burstof radiation on the other side, this time moving horizontally.

Problem 17.2 Calculate the total energy of the diffraction radiation by integratingEq. (17.16) over the angle and the frequency.

Solution: The diffraction radiation is localized at small angles, so we can ap-proximate sin θ ≈ θ and extend the integration range in the integral over the angleto infinity:

W = 2π

∫ ∞

0θdθ

∫ ∞

0dω

d2Wdωd�

= Z0q2

2π2

∫ ∞

0

θ3dθ

(γ−2 + θ2)2

∫ ∞

0dωF

(ωaθ

c,ωa

).

Introducing the integration variables u = γθ and v = aω/cγ we obtain

W = AZ0q2cγ

2π2a,

where the numerical factor A is

A =∫ ∞

0

u3du

(1 + u2)2

∫ ∞

0dvF (uv, v) .

The two-dimensional integral can be calculated numerically and gives

A ≈ 0.90 .

Reference

1. Z. Huang, G. Stupakov, M. Zolotorev, Calculation and optimization of laser acceleration invacuum. Phys. Rev. ST Accel. Beams 7, 011302 (2004)

Page 222: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

Chapter 18Synchrotron Radiation

A relativistic charge moving along a circular orbit emits synchrotron radiation. Inthis chapter we will calculate the intensity of the radiation in the limit γ � 1. Usingthe Liénard–Wiechert potentials we first derive the fields at a large distance fromthe charge in the plane of the orbit and find the radiation spectrum. We then discussangular and spectral distributions of the synchrotron radiation for arbitrary angles.

18.1 Synchrotron Radiation Pulse in the Plane of the Orbit

To simplify calculations, we begin our analysis with the particular case of the radi-ation emitted in the plane of the orbit. The layout for our calculation is shown inFig. 18.1. A point charge q moves on a circular orbit of radius ρ with constant mag-nitude of the velocity v. When the orbit is the result of a transverse magnetic field,ρ is given by Eq. (5.2). An observer is located at point O in the plane of the orbit farfrom the origin of the radiation. The observer will see a periodic sequence of pulsesof electromagnetic radiation repeating with the revolution period of the particle inthe ring. Each pulse originates from the region x ≈ z ≈ 0 in the coordinate systemshown in the figure (the origin of the coordinate system is located at the point wherethe observation line touches the circle).

To calculate the radiation electromagnetic field, wewill use the Liénard–Wiechertpotentials (15.10).Because the observer is located at a large distance from theparticle,we can replace R in the denominator of Eq. (15.10) by r , the distance from theobservation point to the origin of the coordinate system:

A(r, t) = Z0q

4πr

β(tret)

1 − β(tret) · n . (18.1)

© Springer International Publishing AG, part of Springer Nature 2018G. Stupakov and G. Penn, Classical Mechanics and Electromagnetismin Accelerator Physics, Graduate Texts in Physics,https://doi.org/10.1007/978-3-319-90188-6_18

221

Page 223: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

222 18 Synchrotron Radiation

Fig. 18.1 A schematicshowing the particle’s orbitand the observation point.The y axis is directed out ofthe page

We will also use the fact that locally the radiation field can be represented by a planewave for which, according to Eq. (13.5), the magnetic field can be found from thevector potential,

B = −1

cz × ∂A

∂t, (18.2)

where we have replaced n in Eq. (13.5) by the unit vector in the z direction.We set the time t = 0 at the moment when the particle passes through the origin

of the coordinate system; the position of the particle on the circle at time tret is givenby the angle ωr tret, as shown in Fig. 18.1, with the angular revolution frequency ωr =βc/ρ. As we will see later, the main contribution to the radiation pulse comes from asmall part of the particle trajectory where the angle ωr tret is small, ωr tret � 1. Due tothe smallness of this angle we can use an approximation R(tret) ≈ r − ρ sin(ωr tret)which upon substitution into the equation R(tret) = c(t − tret) gives

r − ρ sin(ωr tret) = c(t − tret) . (18.3)

In what follows, we will replace tret by the dimensionless variable ξ, ξ = tretc/ρ ≈ωr tret. We see that ξ is approximately equal to the angle between the position of theparticle on the circle and the x-axis, which, by assumption, is much smaller than one,|ξ| � 1. Equation (18.3) can now be written as

ct − r = ρ[ξ − sin(βξ)] . (18.4)

For the particle velocity at the retarded time we have

βx (tret) = −β sin(ωr tret) = −β sin(βξ) ,

βz(tret) = β cos(ωr tret) = β cos(βξ) . (18.5)

Because the vector potentialA has only x and z components, it follows fromEq. (18.2)that B is directed along y,

cBy = −∂Ax

∂t= −∂Ax/∂ξ

∂t/∂ξ. (18.6)

The x-component of the vector potential (18.1) as a function of ξ is given by

Page 224: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

18.1 Synchrotron Radiation Pulse in the Plane of the Orbit 223

Fig. 18.2 The radiationpulse of the electromagneticfield in dimensionlessvariables

Ax = Z0q

4πr

βx (tret)

1 − βz(tret)= − Z0q

4πr

β sin(βξ)

1 − β cos(βξ), (18.7)

and the function t (ξ), the derivative ofwhich appears in the denominator ofEq. (18.6),is defined by Eq. (18.4). Using the smallness of ξ we expand the trigonometric func-tions,

ξ − sin(βξ) ≈ ξ(1 − β) + 1

6ξ3 ≈ 1

2γ2ξ + 1

6ξ3

1 − β cos(βξ) ≈ 1 − β + 1

2ξ2 ≈ 1

2γ2+ 1

2ξ2 . (18.8)

Substituting these expressions into Eqs. (18.4) and (18.7) we find

Ax = − Z0q

4πr

γ−2 + ξ2, t = r

c+ ρ

c

(1

2γ2ξ + 1

6ξ3

), (18.9)

with the magnetic field

By = Z0q

πrρ

γ−2 − ξ2

(ξ2 + γ−2)3. (18.10)

The equation for the magnetic field is further simplified if we use the dimen-sionless time variable t = (cγ3/ρ)(t − r/c) and the dimensionless magnetic fieldB = (πrρ/Z0qγ4)By . We then have

B = 1 − ζ2

(ζ2 + 1)3, t = 1

2ζ + 1

6ζ3 , (18.11)

Page 225: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

224 18 Synchrotron Radiation

where ζ = γξ. These two equations implicitly define the function B(t) whose plotis shown in Fig. 18.2. We see from this plot that the characteristic width of the pulse�t ∼ 1, which means that the duration of the pulse in physical units is

�t ∼ ρ

cγ3. (18.12)

If wemake the Fourier transformation of themagnetic field, thewidth of the spectrumcan be estimated as �ω ∼ 1/�t ∼ cγ3/ρ. In the next section we will study thisspectrum in more detail.

18.2 Radiation Spectrum in the Plane of the Orbit

Using the results of the previous section we can now calculate the energy radiatedin a unit solid angle d� along the x-z plane of the orbit and the spectrum of theradiation, as discussed in Sect. 17.2. This energy is given by the time integral (17.7)of the Poynting vector, and the angular distribution of the spectral intensity and thespectrum are defined by Eqs. (17.10) and (17.11), respectively. The magnetic field Bin these equations now has only a By component.

To calculate By(ω) it is convenient to start from Eq. (18.6) which in the Fourierrepresentation becomes

cBy(ω) = iω Ax (ω) . (18.13)

We then use Eq. (18.9) to find the Fourier transform of Ax ,

Ax (ω) =∫ ∞

−∞Ax (t)e

iωt dt

=∫ ∞

−∞Ax (ξ)e

iωt (ξ) dt

dξdξ

= − Z0q

4πr

ρ

ceiωr/c

∫ ∞

−∞ξei(ωρ/2c)(γ−2ξ+ξ3/3)dξ . (18.14)

Introducing the new integration variable ζ = γξ and the critical frequency

ωc = 3cγ3

2ρ, (18.15)

we find

Ax (ω) = −iZ0q

4πr

ρ

cγ2eiωr/cF

(3ω

4ωc

), (18.16)

Page 226: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

18.2 Radiation Spectrum in the Plane of the Orbit 225

where

F(x) = Im∫ ∞

−∞ζeix(ζ+ζ3/3)dζ = 2√

3K2/3

(2x

3

), (18.17)

with K2/3 the modified Bessel function of the second type of order 23 . Note that the

real part of the integral in Eq. (18.17) is equal to zero because of the symmetry ofthe integrand. This gives for the spectrum the following expression:

d2Wdωd�

= Z0q2

12π3

(ωρ

c

)2(

1

γ2

)2

K 22/3

2ωc

). (18.18)

18.3 Synchrotron Radiation Out of the Orbit Plane

Our analysis in the previous two sections was carried out for the radiation in the planeof the orbit. Amore general consideration is needed to treat the radiation propagatingout of plane, at an angle ψ �= 0, as shown in Fig. 18.3. While conceptually similar towhat has been done above, these calculations are more involved, and we will not tryto reproduce them here giving only the final result for the radiation spectrum.

A formula that generalizes Eq. (18.18) for the case ψ �= 0 is

d2Wdωd�

= Z0q2

12π3

(ωρ

c

)2(

1

γ2+ ψ2

)2 [K 2

2/3(χ) + ψ2

1/γ2 + ψ2K 2

1/3(χ)

], (18.19)

where

χ = ωρ

3c

(1

γ2+ ψ2

)3/2

= ω

2ωc

(1 + γ2ψ2

)3/2. (18.20)

It is easy to see that setting ψ = 0 reduces this expression to Eq. (18.18).

Fig. 18.3 The particle’sorbit and the coordinatesystem for the off-axisradiation shown by the redwiggling line

Page 227: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

226 18 Synchrotron Radiation

Fig. 18.4 Intensity of σ and π modes in arbitrary units

The modified Bessel functions of the second type K2/3(χ) and K1/3(χ) decayexponentially when their argument becomes much larger than unity. This means thatin the rangeof frequenciesω ∼ ωc the angular spreadof the radiation canbe estimatedfrom χ ∼ 1 as |ψ| ∼ 1/γ. This estimate is in agreement with our general conclusionin Sect. 15.4 that the bulk of the radiation of a relativistic particle is localized withinan angle 1/γ. Radiation at lower frequencies, ω � ωc, has a larger angular spreadψ ∼ (ωc/ω)1/3/γ ∼ (c/ωρ)1/3.

The two terms in the square brackets of Eq. (18.19) correspond to different polar-izations of the radiation. The first one is the so-called σ mode which has nonzeroEx and By ; this is the only polarization that survives in the limit ψ = 0. The secondterm has fields with nonzero Ey and Bx ; this polarization is called the π mode. Theangular distribution of the spectral intensity, d2W/dωd�, for these two modes isshown in Fig. 18.4. Note that the intensity of the π mode is zero in the plane of theorbit, ψ = 0.

18.4 Integral Characteristics of the Synchrotron Radiation

To find the total spectral energy dW/dω radiated in one revolution into all angles,we need to integrate (18.19) over the solid angle d�. This integration consists of twosteps. First, we need to integrate over the angle ψ. Due to the fast convergence ofthe integral over ψ in the region |ψ| � 1, the limits of integration can be extendedfrom minus to plus infinity. Second, we need to take into account that the radiationintensity does not depend on the angle in the plane of the orbit, which we will denoteby θ. The integration over the angle θ extends from 0 to 2π covering all possibledirections into which a particle radiates in one turn on the circle,

dWdω

=∫

d�d2Wdωd�

=∫ 2π

0dθ

∫ ∞

−∞dψ

d2Wdωd�

= 2π∫ ∞

−∞dψ

d2Wdωd�

. (18.21)

The result can be written in the following form:

Page 228: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

18.4 Integral Characteristics of the Synchrotron Radiation 227

Fig. 18.5 S function

dWdω

= 2πρ

c

Z0q2cγ

9πρS

ωc

), (18.22)

where, substituting τ = γψ,

S(x) = 27x2

16π2

∫ ∞

−∞dτ

(1 + τ 2)2

×[K 2

2/3

( x2(1 + τ 2)3/2

)+ τ 2

1 + τ 2K 2

1/3

( x2(1 + τ 2)3/2

)]. (18.23)

The function S is normalized to one,∫ ∞0 dx S(x) = 1; it is plotted in Fig. 18.5. One

can show (with some difficulty) that the function S can also be written as

S(x) = 9√3

8πx

∫ ∞

xK5/3(y)dy . (18.24)

This form of S is the most common definition used and is convenient for establishingits asymptotic expressions in the limits of small and large values of the argument x :

S(x) 27

√3

21/3�

(5

3

)x1/3 , x � 1 , (18.25a)

S(x) 9

8

√3

√xe−x , x � 1 . (18.25b)

Integrating dW/dω over all frequencies, we find the total energy W radiated in onerevolution:

Page 229: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

228 18 Synchrotron Radiation

Wr =∫ ∞

0dω

dWdω

= 2πρ

c

Z0q2cγ

9πρωc . (18.26)

The radiation power (the energy radiated per unit time) by a point charge is

P = Wr

2πρ/c= Z0q2cγ

9πρωc = Z0q2c2γ4

6πρ2= 2rcmc3γ4

3ρ2, (18.27)

where rc = q2/(4πε0mc2) is the classical radius of the particle.In quantumelectrodynamics the radiation is representedbyphotons travellingwith

the speed of light, where each photon of a given frequency carries energy hν = �ω.Estimating the typical frequency as ωc, we can write the number of photons emittedper unit length for a relativistic particle as

dN

dz≈ P

�ωcc= Z0q2

9π�

γ

ρ= 4

γ

)−1

, (18.28)

taking q as the electron charge and where α = e2/4πε0�c is the fine structure con-stant, α ≈ 1/137. The length scale ρ/γ corresponds to a circular arc of 1/γ radians,which is the instantaneous angular spread of the radiated beam as it crosses a fixedlocation a large distance away.We can be a little more precise by noting that the typi-cal photon frequency weighted by photon number instead of power is about 0.31ωc.This gives for dN/dz about 1.44α/(ρ/γ).

Worked Examples

Problem 18.1 Find the asymptotic dependence By(t) for |t − r/c| � ρ/cγ3.

Solution: For |t − r/c| � ρ/cγ3, we have |t | = |t − r/c|/(ρ/cγ3) � 1. Theequation t = ζ/2 + ζ3/6 then implies |ζ| � 1 and t ≈ ζ3/6. For the dimensionlessmagnetic field B we obtain

B = 1 − ζ2

(ζ2 + 1)3≈ −ζ−4 ≈ −(6t)−4/3 .

We recall that By ∝ B.

Problem 18.2 Prove that the area under the curve By(t) is equal to zero (that is∫ ∞−∞ By(t)dt = 0).

Solution: The area under the curve By(t) is

∫ ∞

−∞By(t)dt ∝

∫ ∞

−∞Bdt =

∫ ∞

−∞Bdt

dζdζ

=∫ ∞

−∞(1 − ζ2)(1 + ζ2)

2(1 + ζ2)3dζ

Page 230: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

18.4 Integral Characteristics of the Synchrotron Radiation 229

= ζ/2

1 + ζ2

∣∣∣∣∞

−∞= 0 ,

where we used t = ζ/2 + ζ3/6 to find dt/dζ and we note that the integration is thereverse of the derivative of Ax used to calculate By in Eq. (18.10). We could haveseen this more simply by noting that By(t) ∝ ∂Ax/∂t , so

∫ ∞−∞ Bydt ∝ Ax |∞−∞ = 0

for Ax ∝ ξ/(γ−2 + ξ2).

Problem 18.3 SimplifyEq. (18.19) in the limitψ � 1/γ.Make a plot of the quantityω−2/3d2W/(dωd�) versus the quantity ωρψ3/c. Infer from these equations that theangular spread of the radiation at frequency ω � ωc is of order of (c/ωρ)1/3.

Solution: We are interested in the angular dependence of the radiation at largeangles (outside the main peak, ψ � 1/γ). Equation (18.19) becomes

d2Wdωd�

∝(ωρ

c

)2ψ4

[K 2

2/3(χ) + K 21/3(χ)

],

andχ ωρ

3cψ3 = ω

2ωcγ3ψ3 .

Then we find

d2Wdωd�

∝(ωρ

c

)2/3 (ωρ

c

)4/3ψ4

[K 2

2/3(χ) + K 21/3(χ)

]

∝(ωρ

c

)2/3χ4/3

[K 2

2/3(χ) + K 21/3(χ)

].

Figure18.6 shows the functionχ4/3[K 22/3(χ) + K 2

1/3(χ)]. It reaches a peak of roughly1.4 at χ 0.1, but even for χ � 1 it remains above unity in this approximation. Atχ = 2 this function is already below0.1. Sowe conclude that the radiation is localizedin the range of angles ωρψ3/c � 1 and the angular spread is

�ψ ∼(

λ

ρ

)1/3

.

This approximation only makes sense for frequencies well below ωc, otherwise therange of angles is set by 1/γ.

Problem 18.4 Calculate the RF power needed to compensate the synchrotronradiation in the Advanced Light Source. The parameters are: energy is 1.9GeV,I = 0.56A, circumference is 200m.

Solution: The power from synchrotron radiation of a single electron isP = 2rcmc2γ4c/3ρ2 = cCγE4/2πρ2, where rc is the classical electron radius andCγ ≡ 4πrc/3(mc2)3. With rc = 2.82 × 10−15 m and mec2 = 5.11 × 10−4 GeV, wefind Cγ = 8.9 × 10−5 m/GeV3. Thus the power radiated is 54GeV/s.

Page 231: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

230 18 Synchrotron Radiation

Fig. 18.6 Plot of

χ4/3[K 22/3(χ) + K 2

1/3(χ)].

The energy lost in one revolution is

�E =∫

Pdt = 2πρP

c= CγE4

ρ

= 36 keV .

The total power from all of the electrons in the ring is

PT = (�E)I/e = (36 kV)(0.56 A) = 20 kW .

Page 232: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

Chapter 19Undulator Radiation

Undulators and wigglers are widely used in modern accelerator-based light sources.The significant amount of radiated power they can emit also make them importantcomponents for electron or positron damping rings. We derive the properties ofthe undulator radiation using the solution of the Thomson scattering problem fromChap.16.

19.1 Undulators and Wigglers

Aplane undulator is shown inFig. 19.1. Themagnetic field in the undulator is directedalong the y axis and sinusoidally varies in the z direction,

By(z) = B0 cos kuz , (19.1)

where ku = 2π/λu withλu the undulator period and B0 the amplitude of themagneticfield. We will denote by Nu the number of periods in the undulator.

We assume that a relativistic beam propagates along the z axis with velocity v.The equation of motion in the horizontal (x–z) plane is obtained by equating theLorentz force in the x direction to the product of the relativistic mass mγ and theacceleration x :

mγ x = −qvB0 cos kuz

≈ −qcB0 cos kuct , (19.2)

where we have used the approximation z ≈ vt ≈ ct . The solution of this equation is

© Springer International Publishing AG, part of Springer Nature 2018G. Stupakov and G. Penn, Classical Mechanics and Electromagnetismin Accelerator Physics, Graduate Texts in Physics,https://doi.org/10.1007/978-3-319-90188-6_19

231

Page 233: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

232 19 Undulator Radiation

Fig. 19.1 Magnetic fieldand an electron trajectory ina plane undulator

x ≈ qB0

mγk2uccos kuct ≈ qB0

mγk2uccos kuz , (19.3)

where we have set two integration constants equal to zero (these constants can beeliminated by the proper choice of the offset and the direction of the z axis). Thesinusoidal orbit (19.3) is characterized by the maximum deflection angle

dx

dz

∣∣∣∣max

= qB0

mγkuc. (19.4)

Comparing this angle with the inverse Lorentz factor γ−1 we introduce an importantquantity, the undulator parameter K :

K = γdx

dz

∣∣∣∣max

= qB0

kumc� 0.934λu[cm] B0[Tesla] , (19.5)

where for m we have used the value of the electron mass. A device with K � 1 isusually referred to as an undulator, and the term wiggler is used in the case of largeK . We will use the term undulator in the general case when K is not specified.

With the particle orbit in the undulator found above, we could calculate its radi-ation using the retarded potential formalism as we did for the synchrotron radiationin Chap.18. Here, however, we will use a different approach, and calculate the spec-trum of the radiation by applying the Lorentz transformation to the solution of theThomson scattering problem studied in Chap. 16.

19.2 Undulator Radiation for K � 1

Let us consider a long undulator with K � 1 and a large number of periods, andneglect the effects associated with the entrance to and the exit from the undulator.To calculate the undulator radiation we will transform into the frame of reference inwhich the particle is on average at rest. We will use the primes to denote quantitiesin this reference frame.

In the particle frame, the undulator is moving in the negative direction of the zaxis with velocity v = (0, 0,−v). Using the Lorentz transformations for the fields

Page 234: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

19.2 Undulator Radiation for K � 1 233

and assuming γ � 1 we find that in this frame the undulator field has both electricand magnetic components perpendicular to the z axis:

E ′x = γvB0 cos kuz ≈ γcB0 cos [kuγ(z′ − ct ′)] ,

B ′y = γB0 cos kuz ≈ γB0 cos [kuγ(z′ − ct ′)] , (19.6)

where on the last step we have used the Lorentz transformation to express the coor-dinate z through the coordinate z′ in the particle frame and replaced β by 1,

z = γ(z′ − βct ′) ≈ γ(z′ − ct ′) . (19.7)

In Eq. (19.6) we recognize a plane electromagnetic wave propagating in the negativez direction with the frequencyω′ = γkuc and the field γ times larger than the lab fieldof the undulator. This field would cause oscillations of the particle and would resultin the Thomson scattering that we studied in Chap.16. To calculate the parametera defined by Eq. (16.23), we substitute for the electric field E0 → γcB0 and for thefrequency ω → ω′ = γkuc and find that a is exactly equal to the undulator parameter(19.5). Because the results of Chap. 16 are valid in the limit a � 1, we require thatK � 1. We shall also see in Sect. 19.4 that only in the limit K � 1 is the electronvelocity v a good approximation for the velocity used in the Lorentz transformation.

The incident electromagnetic wave is scattered by the electron and the intensityof the scattered radiation is given by Eq. (16.27), which we rewrite here using thenew notation,

dP ′

d�′ = Z0

32π2

q4γ2B20

m2(1 − sin2 θ′ cos2 φ) . (19.8)

Note that the angle φ′ in the x ′–y′ plane is the same as the angle φ. We now needto transform the quantities dP ′, d�′, θ′, as well as ω′ into the lab frame. For thetransformation of the angle θ we can use Eq. (B.13) from the Doppler effect,

sin θ′ = sin θ

γ(1 − β cos θ)≈ 2γθ

1 + γ2θ2, (19.9)

where we have assumed that θ � 1, expanded cos θ ≈ 1 − θ2/2, and used 1 − β ≈1/2γ2. The relativistic frequency transformation Eq. (B.16) gives

ω ≈ 2γω′

1 + γ2θ2= 2γ2kuc

1 + γ2θ2. (19.10)

Note that the frequency now varies with the angle θ and hence depends on thedirection of the radiation. We denote by ω0 the maximum frequency of the radiationwhich goes in the forward direction, θ = 0. This frequency is equal to

Page 235: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

234 19 Undulator Radiation

ω0 = 2γω′ = 2γ2kuc . (19.11)

For the differential of the elementary solid angle d�′ we write it as

d�′ = sin(θ′)dθ′dφ

= |d cos(θ′)|dφ , (19.12)

and then use Eq. (B.13) to obtain

d�′ = 1 − β2

(1 − β cos θ)2|d cos(θ)|dφ ≈ 4γ2

(1 + γ2θ2)2d�, (19.13)

where d� is the solid angle in the lab frame corresponding to d�′ in the beamframe. Finally, we need to transform into the lab frame the differential dP ′ whichhas the meaning of the radiated energy of the electromagnetic field per unit time,dP ′ = dE ′/dt ′. We know the transformation of time, dt ′ = dt/γ, which is the timedilation effect. The easiestway to figure out how to transform the energy is to considerradiation as a collection of photons. In quantum language, the photon energy isequal to �ω, and the number of photons Nph is the same in any reference frame,N ′ph = Nph (cf. Problem16.5). Hence the energy of an ensemble of photons with the

same frequency is transformed as the frequency, dE ′ = Nph�ω′ = dE(ω′/ω). Wenow have

dPd�

= dE

d�dt

= dE ′

d�′dt ′ω

ω′1

γ

4γ2

(1 + γ2θ2)2

= dP ′

d�′8γ2

(1 + γ2θ2)3

= Z0

32π2

q4γ2B20

m2

[

1 − 4γ2θ2

(1 + γ2θ2)2cos2 φ

]8γ2

(1 + γ2θ2)3

= Z0

4π2

q4γ4B20

m2

(1 + γ2θ2)2 − 4γ2θ2 cos2 φ

(1 + γ2θ2)5. (19.14)

From this equation it follows that the dominant part of the radiation energy is emittedwithin the angle θ ∼ 1/γ, in agreement with the general principles of the radiationof relativistic particles discussed in Sect. 15.4.

In Sect. 17.2 we introduced the angular distribution of the spectral intensity of theradiation, Eq. (17.11), so that the integral over the frequency gives the energy radiatedper unit solid angle. We can apply the same notion to the undulator radiation andintroduce the angular distribution of the spectral power of the radiation, dP/d�dω,such that

∫ ∞0 (dP/d�dω)dω = dP/d�. In the model of an infinitely long undulator

thatwe assumehere the frequency is uniquely related to the angle throughEq. (19.10).

Page 236: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

19.2 Undulator Radiation for K � 1 235

Mathematically this one-to-one correspondence between the frequency and the anglecan be expressed through a delta function that indicates an infinitely narrow spectrumat each given angle:

dPd�dω

= dPd�

δ

(

ω − 2γ2kuc

1 + γ2θ2

)

. (19.15)

It is clear that the integration of dP/d�dω over the frequency gives the angulardistribution of the power dP/d�. The infinitely narrow frequency spectrum here isdue to the assumption of an infinitely long undulator. Taking into account a finiteundulator length, and hence a finite time of flight through the undulator, introducesa nonzero width of the spectrum, as we will see in the next section.

To find the energy radiated per unit time in all directions we integrate equa-tion (19.14) over � using the approximation sin θ ≈ θ for small angles,

P0 =∫

dPd�

d� ≈∫ ∞

0θdθ

∫ 2π

0dφ

dPd�

= Z0

12π

q4γ2B20

m2, (19.16)

wherewe have used themathematical identity∫ ∞0 (1 + x2)(1 + x)−5dx = 1/3. If we

rewrite in this expression the square of the peak amplitude of the magnetic field B20

in terms of the square of the field averaged over the undulator period, B20 = 2〈B2〉,

and compare it with the intensity of the synchrotron radiation (18.27) (rememberingthat ρ = γmc/qB), we find that they are equal. Hence the radiated power from theundulator, per unit time, is equal to the radiated power fromabendingmagnetwith thesame averaged square of the magnetic field. In contrast to the synchrotron radiation,where the radiation spectrum is broad, in an undulator with a small K the radiationat a given angle θ is localized within a narrow range around the frequency (19.10).

19.3 Effects of Finite Length of the Undulator

Taking now into account the finite length of the undulator, we will assume that thenumber of periods in the undulator Nu is large, Nu � 1. As was pointed out in theprevious section, the finite number of periods in the undulator results in a nonzerowidth of the spectrum of the radiation. The shape of the spectrum can be rather easilyestablished if one looks at the time dependence of the electric field in the radiationpulse. In the particle reference frame an undulator with Nu periods is representedby the same plane wave (19.6) which now has Nu periods. The particle executesNu oscillations in the wave and emits a scattered radiation pulse with the same Nu

number of periods. An example of such pulse is shown in Fig. 19.2 for Nu = 10.One can see that the pulse in this case is a piece of a sinusoidal function with the

Page 237: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

236 19 Undulator Radiation

number of oscillations equal to the number of periods in the undulator; the oscillationfrequency is given by the same Eq. (19.10), which we now denote by ω1(θ),

ω1(θ) = 2γ2kuc

1 + γ2θ2. (19.17)

As is well known, the Fourier transform of a truncated sinusoidal pulse is given bythe sinc function, and is proportional to

sin(πNu�ω/ω1(θ))

�ω, (19.18)

where �ω = ω − ω1(θ), and we have assumed |�ω| � ω1(θ). It is not surprisingthat the delta function in (19.15) is now replaced by the square (because the poweris proportional to the electric field squared) of the sinc function:

dPd�dω

= dPd�

sin(πNu�ω/ω1(θ))2

π2Nu�ω2/ω1(θ), (19.19)

where dP/d� is given byEq. (19.14) and the additional factors on the right-hand sideare chosen from the requirement that the integration of dP/d�dω over the frequencyω should give dP/d�. It follows from this equation that the spectral width of theundulator radiation in a given direction θ is inversely proportional to the number ofperiods, �ω ∼ ω1(θ)/Nu .

For an undulator of finite length it makesmore sense to talk about the total radiatedenergy W and its angular spectrum dW/d�dω rather than the radiation power Pand the angular spectrum dP/d�dω. The total energy radiated from the length ofthe undulator is obtained by multiplying the power by the time of passage throughthe undulator, approximately equal to Lu/c where Lu = 2πNu/ku is the length ofthe undulator. In particular, dW/d�dω is given by the right-hand side of Eq. (19.19)multiplied by Lu/c.

19.4 Wiggler Radiation for K � 1

The undulator radiation in a wiggler with large K can be derived using the sameapproach as for K � 1 case—using the Lorentz transformation to the particle frameof reference. However, calculations become much more involved when K � 1, sowe will limit our analysis here to the derivation of the radiation frequency and adiscussion of some qualitative properties of the spectrum. We assume that K � γ.

First, we need to calculate the averaged velocity vz of the particle moving alongthe z-axis of the wiggler. The x velocity can be found from Eq. (19.3),

Page 238: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

19.4 Wiggler Radiation for K � 1 237

Fig. 19.2 Electric fieldversus time (in arbitraryunits) for undulator radiationin the forward direction withNu = 10, K = 0.1 andγ = 10. The time isnormalized by the periodT0 = 2π/ω1(0)

vx = dx

dt= −Kc

γsin(kuct) , (19.20)

which gives for vz

vz =√

v2 − v2x ≈ v

(

1 − v2x

2c2

)

= v

[

1 − K 2

2γ2sin2(kuct)

]

, (19.21)

where we have used the approximations |vx | � v and v ≈ c. Averaging Eq. (19.21)over time, we obtain

vz = v

(

1 − K 2

4γ2

)

. (19.22)

This is the velocity of the reference frame in which the particle, on average, is at rest.Given that K � γ, we see that this velocity is still close to the speed of light, but theeffective gamma factor corresponding to vz , which we denote by γz , is smaller thanthe original γ:

γz = 1√

1 − v2z /c

2≈

[

1 − v2

c2

(

1 − K 2

2γ2

)]−1/2

≈(

1 − v2

c2+ K 2

2γ2

)−1/2

=(

1

γ2+ K 2

2γ2

)−1/2

= γ√

1 + K 2/2. (19.23)

When we make the Lorentz transformation to the particle frame of reference, thefrequency of the electromagnetic plane wave Lorentz transformed from the wigglerfield will be ω′ = γzkuc; correspondingly γz should replace γ in Eq. (19.10) for theradiation frequency in the lab frame,

Page 239: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

238 19 Undulator Radiation

Fig. 19.3 Electric field versus time (in arbitrary units) for an undulator radiation with K = 1 (leftpanel) and K = 4 (right panel). The time is normalized by T0 = 2π/ω0. The undulator has Nu = 10periods and γ = 10

Fig. 19.4 Undulatorspectrum for K = 4 (in thelimit Nu � 1). Thefrequency ω0 is given byEq. (19.25)

ω = 2γ2z kuc

1 + γ2z θ

2= 2γ2kuc

1 + K 2/2 + γ2θ2. (19.24)

In particular, for θ = 0, that is the frequency in the forward direction, we obtain

ω0 = 2kucγ2

1 + K 2/2. (19.25)

This, however, is not the maximum frequency of the radiation as it was in the case ofK � 1. The reason for this is illustrated by Fig. 19.3 where we show the electric fieldof the wiggler radiation detected in the forward direction for the two cases K = 1and K = 4. The repetition rate of the pulses is equal to the frequency ω0, but forlarger values of K the spikes of the electric field become narrower which leads toa rich content of higher harmonics in the radiation spectrum. Indeed, as shown inFig. 19.4 for K = 4, the undulator spectrum (integrated over all angles �) extendsto frequencies much higher than ω0.

Page 240: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

19.4 Wiggler Radiation for K � 1 239

Fig. 19.5 The spectrum ofthe undulator radiation givenby Eq. (19.26)

Worked Examples

Problem 19.1 Integrate equation (19.15) over the solid angle and show that theintensity of the radiation per unit frequency is

dPdω

= 3P0

ω0

ω

ω0

[

1 − 2ω

ω0+ 2

ω0

)2]

(19.26)

for ω < ω0, and zero for ω > ω0. The plot of this function is shown in Fig. 19.5.

Solution: Expressing the elementary solid angle as d� = θdθdφ (where weapproximated sin θ ≈ θ) we can write dP/d�dω = θ−1dP/dθdφdω. We then firstintegrate it over φ to obtain

dPθdθdω

= Z0

4π2

q4γ4B20

m2δ

(

ω − 2γ2kuc

1 + γ2θ2

)∫ 2π

0dφ

(1 + γ2θ2)2 − 4γ2θ2 cos2 φ

(1 + γ2θ2)5

= Z0

q4γ4B20

m2

(1 + γ2θ2)2 − 2γ2θ2

(1 + γ2θ2)5δ

(

ω − 2γ2kuc

1 + γ2θ2

)

. (19.27)

In the second step we integrate this equation over the angle θ,

dPdω

=∫ ∞

0dθ

dPdθdω

,

where we have extended the upper limit of the integration range to infinity becausethe dominant contribution to the integral comes from small angles∼ 1/γ. Integrationof the delta function in Eq. (19.27) gives an inverse derivative of its argument with

Page 241: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

240 19 Undulator Radiation

respect to θ with the rule that the angle θ should be expressed through the frequencyω from the Eq. (19.10). We then obtain

dPdω

= Z0

q4γ4B20

m2

(1 + γ2θ2)2

2ω0γ2

(1 + γ2θ2)2 − 2γ2θ2

(1 + γ2θ2)5

= Z0

q4γ2B20

m2ω0

ω0

)3 [(ω0

ω

)2 − 2ω0

ω+ 2

]

= 3P0

ω0

ω

ω0

[

1 − 2ω

ω0+ 2

ω0

)2]

,

with P0 = Z0q4γ2B20/12πm

2. From the definition, ω0 is at θ = 0, when the fre-quency is maximized, so ω ≤ ω0.

Page 242: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

Chapter 20Formation Length of Radiationand Coherent Effects

The radiation process is not instantaneous — it requires some time and free spacearound the orbit for the radiation to be formed. In this chapter we estimate the longi-tudinal extent and transverse size of the free space volume needed for the synchrotronradiation. We then analyze the radiation of a bunch of many particles.

20.1 Longitudinal Formation Length

It takes some time and space for a moving charge to generate radiation. Let us takea closer look at the derivation in Sect. 18.1 and try to figure out what fraction of thelength of the orbit is involved in the formation of the synchrotron pulse.

In Eq. (18.10) the variable ξ = cτ/ρ is related to the retarded time τ . It followsfrom this equation that the characteristic width of the electromagnetic pulse in vari-able ξ is �ξ ∼ γ−1, which corresponds to the time duration τ ∼ ρ/cγ. Hence thelength of the orbit necessary for the formation of the radiation pulse, which we willcall the formation length l f , is

l f ∼ cτ ∼ ρ

γ. (20.1)

In quantum language, the formation length is needed to convert a virtual photoncarried by the electromagnetic field of a particle into a real photon.

How does this formation length agree with the duration of the radiation pulse∼ ρ/γ3c? Since the charge ismovingwith the velocity v ≈ c(1 − 1/2γ2), the relativevelocity with which the electromagnetic field propagating with the speed of lightovertakes the charge is �v ∼ c/γ2. During the formation time τ the radiated fieldpropagates ahead of the charge at the distance τ�v ∼ ρ/γ3, which explains theduration of the pulse (18.12).

© Springer International Publishing AG, part of Springer Nature 2018G. Stupakov and G. Penn, Classical Mechanics and Electromagnetismin Accelerator Physics, Graduate Texts in Physics,https://doi.org/10.1007/978-3-319-90188-6_20

241

Page 243: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

242 20 Formation Length of Radiation and Coherent Effects

Fig. 20.1 The orbit consistsof a circular arc limited byangles −ϕ2 < ϕ < ϕ1. Thehorizontal arrow shows thedirection in which theradiation is observed

The concept of the formation length has important practical applications. In ac-celerators particles emit synchrotron radiation when passing through dipole magnetsof finite length, where their trajectory is an arc of a circle. Outside of the magnet thetrajectories are represented by segments of straight lines for which the formal valueof the bending radius is infinity. The question then arises: under what conditions canone apply the results of Chap.18 derived for a circular orbit to the case of a magnetof finite length? The answer to this question is that the circular orbit approximationis valid if the magnet length L is several times longer than l f . Because the changeof angle of a charged particle passing through a magnet is roughly equal to L/ρ (weassume that the angle is small), this means that the angle must be several times largerthan 1/γ.

Radiation from a magnet that is shorter than l f has very different properties thanwhat we have calculated in Chap. 18. Even for longer magnets, the radiation comingfrom the region within a distance ∼l f from the edges of the magnet will be different.Some of the properties of radiation from short magnets can be explained using thetime profile of the radiation pulse shown in Fig. 18.2. Recall that each point on thisplot has a corresponding position on the orbit from which the field at this point wasemitted. This position is determined by the value of the variable ξ approximatelyequal to the angle with the vertical axis.

Let us now consider an orbit that consists of a circular arc with the angularextension−ϕ2 < ϕ < ϕ1 as shown in Fig. 20.1; outside of the arc the particle movesalong straight lines (tangential to the end points of the arc) with a constant velocity.Since there is no acceleration in the straight parts of the orbit, the radiation pulseshown in Fig. 18.2 will be truncated: the value of the radiation field B becomes zerofor ξ < −ϕ2 or ξ > ϕ1, while it remains the same for the points on the arc where−ϕ2 < ξ < ϕ1. Recalling the relation ζ = γξ in Eq. (18.11), we conclude that theradiation pulse for a short magnet is given by the same Eq. (18.11), in which ζ is nowconstrained by −γϕ2 < ζ < γϕ1. An example of the pulse shape for γϕ2 = 0.7 andγϕ1 = 0.5 is shown in Fig. 20.2. In reality, the discontinuities of the field at the frontand the tail will be somewhat smeared out due to the fact that the magnetic fields donot abruptly drop to zero at the entrance to and the exit from the magnet. Instead the

Page 244: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

20.1 Longitudinal Formation Length 243

Fig. 20.2 The radiationpulse of the electromagneticfield for a short magnet withγϕ2 = 0.7 and γϕ1 = 0.5.The variables B and t are thesame as in Fig. 18.2. Theshape of the central part ofthe pulse is the same as inFig. 18.2; the field drops tozero outside of this pulse

fields decrease over some finite extent, usually related to the magnetic aperture. Thisregion of non-uniform field is called the fringe field1.

The estimate for the formation length given in Eq. (20.1) is valid for the formationof the main part of the radiation pulse which carries the bulk of the electromagneticenergy. The formation length is different for the low frequency part of the radiation,with frequencies ω � ωc. Low frequencies are produced by the long tails of theelectromagnetic pulse in Fig. 18.2 and require a longer formation length l f (ω). Toestimate l f (ω) we will analyze Eq. (18.14) and find the length �ξ of the region thatmakes a dominant contribution to the integral for a given frequency ω. We first notethat for ω � ωc one can neglect the term with γ−2 in the exponent of Eq. (18.14),and the integral becomes

∫ ∞

−∞ξeiωρξ3/6cdξ , (20.2)

fromwhich it follows that�ξ ∼ (c/ωρ)1/3.When ξ is much larger than this quantity,the function eiωρξ3/6c begins to rapidly oscillate thus suppressing the contributionfrom the region |�ξ| � (c/ωρ)1/3. Recalling that ξ represents the angle along theorbit, we estimate the formation length as

l f (ω) ∼ ρ�ξ ∼ ρ2/3λ1/3 , (20.3)

where λ = c/ω is the reduced wavelength. For the critical frequency ω = ωc thisformula reduces to the previous expression (20.1).

Using the result of Problem 18.3 that the angular spread of the synchrotron radia-tion at frequency ω � ωc is of order of �ψ ∼ (λ/ρ)1/3, we can relate the formation

1The specific features of the radiation caused by the jumps in B at the entrance and the exit from themagnet are usually referred to as the edge radiation. It has a similarity with the transition radiationstudied in Sect. 17.1.

Page 245: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

244 20 Formation Length of Radiation and Coherent Effects

length to the angular spread of the radiation �ψ,

l f ∼ λ

�ψ2. (20.4)

20.2 Transverse Formation Length

In addition to the requirement of having a long enough path length for the radiationformation, the charge also needs some space in the direction perpendicular to theorbit. We can estimate the needed transverse extension if we note that the radiationwith frequencyω and angular spread�ψ has a transversewavenumber k⊥ ∼ ω�ψ/c.Any electromagnetic field with a characteristic value of k⊥ extends transversely atleast over the distance of one reduced transverse wavelength k−1

⊥ , hence the minimaltransverse size needed for the formation of radiation is

l⊥ ∼ λ

�ψ∼ ρ1/3λ2/3 , (20.5)

where on the last step we have used the relation �ψ ∼ (λ/ρ)1/3 from Problem 18.3.We will call l⊥ the transverse formation length.

An interesting connection of l⊥ to the properties of Gaussian beams can be tracedif we recall the results of Sect. 13.2. For a Gaussian beam with an angular spread θ,the minimal transverse size (at the focal point) is the beam waist w0 which, withina factor of two, is equal to λ/θ (see Eq. (13.20)). Hence l⊥ is analogous to the laserbeam waist w0. Moreover, comparing Eq. (20.4) with Eq. (13.20) for ZR we see thatthe longitudinal formation length is analogous to the Rayleigh length of a focusedlaser beam.

The practical importance of the transverse formation length is that synchrotronradiation can be suppressed by nearby metal walls of the vacuum chamber if theyare located closer than l⊥ — the so-called shielding effect. More specifically, thesynchrotron radiation of a relativistic beam that moves on a circular orbit in a metalpipe having a transverse size a is suppressed in the frequency range where a � l⊥.From Eq. (20.5) we find that the suppression occurs at long wavelength, λ �

√a3/ρ.

The suppression factor, defined as the ratio of the power radiated in the pipe to the freespace radiation, depends on the cross section shape of the pipe. In Fig. 20.3 we plotthe suppression factor for a model in which the circular orbit of the particle is locatedin the middle plane between two parallel perfectly conducting plates separated bya distance 2 h (so that the distance from the orbit to each plate is h). The resultshown in Fig. 20.3 is valid in the limit of small frequencies, ω � ωc, when the freespace radiation is given by Eq. (18.25a). Note that the horizontal axis in the plot isωh3/2ρ−1/2/c ∼ (h/ l⊥)3/2, and one can see that the suppression factor approacheszero when h becomes much smaller than l⊥. Interestingly, the radiation is actuallyamplified by about 30% in the region of frequencies where ωh3/2ρ−1/2/c ≈ 2.

Page 246: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

20.2 Transverse Formation Length 245

Fig. 20.3 Suppressionfactor for the intensity of thesynchrotron radiation as afunction of the frequency ωfor the case of parallelconducting plates

Fig. 20.4 Two particles, iand j , in a bunch emitseparate pulses ofelectromagnetic radiation

This shielding effect plays a crucial role in electron accelerators with intense shortelectron bunches, acting to limit the beam energy loss going into radiation at longwavelengths.

20.3 Coherent Radiation

Wenowconsider radiation of a bunch of particles. First, we neglect the transverse sizeof the bunch and consider the beam as infinitely thinwith the longitudinal distributionfunction λ(s). This function gives the probability for a particle to be located at s; itis normalized so that

∫λ(s)ds = 1.

When twoparticles radiate electromagnetic pulses as shown inFig. 20.4, the pulsespropagate to the observer with delays that are determined by the arrival times of theparticles to the point on the orbit from which the radiation is emitted. Let us denoteby B(t) the magnetic field of the pulse at the detector emitted by a reference particlein the beam (for synchrotron radiation, the function B(t) is calculated in Sect. 18.1).In what follows, we will also need the Fourier transform of this field,

B(ω) =∫ ∞

−∞dt B(t)eiωt . (20.6)

The total field B(t) radiated by the bunch is the sum of the pulses,

Page 247: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

246 20 Formation Length of Radiation and Coherent Effects

B(t) =N∑i=1

B(t − ti ) , (20.7)

where ti = si/c with si the position of the particle i , and N is the total number ofparticles in the bunch. The Fourier image of this field is

B(ω) =∫

dt Beiωt =N∑i=1

∫dt B(t − ti )e

iωt =N∑i=1

B(ω)eiωti . (20.8)

The spectral intensity of the radiation is proportional to |B(ω)|2 for which we have

|B(ω)|2 =∣∣∣∣∣

N∑i=1

B(ω)eiωti

∣∣∣∣∣2

= |B(ω)|2⎛⎝N +

∑i �= j

eiω(ti−t j )

⎞⎠

= N |B(ω)|2 + 2|B(ω)|2∑i< j

cos

(ωsi − s j

c

). (20.9)

The first term in this equation is the incoherent radiation — it is proportional to thenumber of particles in the beam N . The second one is the coherent radiation term.The number of summands in the last sum is N (N − 1)/2 ≈ N 2/2. Instead of doingsummation over i and j we can average cos(ω(si − s j )/c) and multiply the resultby N (N − 1)/2 ≈ N 2/2 assuming that si and s j are distributed with the probabilitygiven by λ(s)2:

2∑i< j

cos

(ωsi − s j

c

)≈ N 2

∫ds ′ds ′′λ(s ′)λ(s ′′) cos

(ωs ′ − s ′′

c

)

= N 2F(ω) , (20.10)

where the form factor F(ω) is

F(ω) =∫

ds ′ds ′′λ(s ′)λ(s ′′) cos(

ωs ′ − s ′′

c

). (20.11)

Because the spectral intensity of the radiation dW/dωd� is proportional to |B(ω)|2(see Eq. (17.10)) we obtain

dWdωd�

∣∣∣∣bunch

= dW1

dωd�

[N + N 2F(ω)

], (20.12)

where dW1/dωd� refers to the radiation of a single particle.

2Here we implicitly assume that the positions of particles i and j are not correlated.

Page 248: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

20.3 Coherent Radiation 247

Equation (20.11) can also be written as

F(ω) =∣∣∣∣∫ ∞

−∞ds λ(s)eiωs/c

∣∣∣∣2

, (20.13)

which is easily established by writing the square of the absolute value as a product ofthe integral

∫ ∞−∞ ds λ(s)eiωs/c with its complex conjugate. For aGaussian distribution

function,

λ(s) = 1√2πσz

e−s2/2σ2z , (20.14)

we have

F(ω) = e−(ωσz/c)2 . (20.15)

We see that for reduced wavelengths that are longer than the bunch length, λ � σz ,F(ω) approaches unity and the coherent radiation term in Eq. (20.12) dominatesbecause the power scales as the number of particles squared. However, this radiationcanonly occur at longwavelengths, and inmany cases it is suppressedby the shieldingeffect. In the opposite limit, λ � σz , the form factor F is exponentially small andthe coherent radiation is negligible.

20.4 Effect of the Transverse Size of the Beam

In the previous analysis we neglected the transverse size of the beam and tacitlyassumed that the radiation propagates in the forward direction. We now take intoaccount that the radiation can be emitted at an angle to the direction of motion ofthe beam and consider a 3D distribution of the beam illustrated in Fig. 20.5. The 3Ddistribution function is λ(r) normalized so that

∫d3rλ(r) = 1. From Fig. 20.5 it is

seen that the delay between pulses radiated by two particles located in the bunch atcoordinates r i and r j is equal to �t = (r i − r j ) · n/c, where n is the unit vector inthe direction of radiation. The radiation field of the bunch (20.9) can now be writtenas

|B(ω, n)|2 = N |B(ω, n)|2 + 2|B(ω, n)|2∑i< j

cos

(ωn · (r i − r j )

c

), (20.16)

where the second argument n in the magnetic field indicates the dependence versusthe direction of radiation.

Repeating the derivation from the previous section we arrive at the followingresult for the form factor F :

Page 249: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

248 20 Formation Length of Radiation and Coherent Effects

Fig. 20.5 Radiation pulsesthat propagate at an angle ψto the direction of motion ofthe beam

F(ω, n) =∫

d3r ′d3r ′′λ(r ′)λ(r ′′) cos(

ωn · (r ′ − r ′′)

c

), (20.17)

which can also be written as the square of the absolute value of the three dimensionalFourier transform of the distribution function,

F(ω, n) =∣∣∣∣∫

d3rλ(r)eiωn·r/c∣∣∣∣2

. (20.18)

The calculation of the form factor for a highly simplified beam distribution inProblem 20.1 shows that the coherent radiation is suppressed if

σr � λ/ψ , (20.19)

that is if the transverse size of the beam is larger than the transverse formation size ofthe radiation. As an order of magnitude estimate, this conclusion is valid for a broadclass of other distribution functions.

Worked Examples

Problem 20.1 Calculate the integral equation (20.18) for a “pancake” distribution

λ(r) = δ(z)1

2πσ2r

e−(x2+y2)/2σ2r .

Solution: We need to calculate the integral

∫d3rλ(r)eikn·r ,

with k = ω/c and the vector n directed at angle ψ to the z axis. By symmetry of theproblem, we can choose n to lie in the x-z plane, n = (sinψ, 0, cosψ). For the formfactor F we obtain:

Page 250: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

20.4 Effect of the Transverse Size of the Beam 249

F =∣∣∣∣∫

dx dy dzδ(z)

2πσ2r

e−(x2+y2)/2σ2r eik(x sinψ+z cosψ)

∣∣∣∣2

= 1

2πσ2r

∣∣∣∣∫

dx e−x2/2σ2r eikx sinψ

∣∣∣∣2

= e−(k2σ2r /2) sin

2 ψ .

For small angles, the coherent radiation is suppressed if λ � σrψ.

Page 251: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

Chapter 21Topics in Laser-Driven Acceleration

A focused laser beam can easily produce an extremely high electric field at the focalpoint. For example, a laser beam with an energy of 1 J and pulse duration of 100 fsfocused to a spot size of 10µm, has a maximum electric field of about 40GV/cmwhich ismany orders ofmagnitude larger thanwhat is used in traditional accelerators.In this chapter we will discuss several topics related to the possibility of using laserfields for accelerating charged particles.

21.1 The Lawson–Woodward Theorem

In the simplest setup one can try to accelerate charged particles by sending themthrough a focal point of a laser beam, as shown in Fig. 21.1. This is referred toas direct laser acceleration, or DLA. Because the electric field in a paraxial laserbeam is directed predominantly transversely to the direction of propagation, one canexpect that the particle trajectories should cross the focal region at an angle so thatthe component of the transverse laser field along the particle trajectory is not zero.Nevertheless, it is not immediately clear what would be the net acceleration effectafter the interaction ceases — the electromagnetic field in the laser beam oscillateswith time, and a particle crossing the focal point would experience both acceleratingand decelerating phases.

Unfortunately, as we will show below, no matter how we organize the interactionof the laser beam with the particles, in the linear approximation, there is no netacceleration in free space. This statement is often called the Lawson–Woodwardtheorem.

We will now explain the meaning of the linear approximation. In this approxima-tion we calculate the energy gain �W of a particle passing through an external fieldE(r, t) assuming that it moves along a straight line with a constant velocity v,

© Springer International Publishing AG, part of Springer Nature 2018G. Stupakov and G. Penn, Classical Mechanics and Electromagnetismin Accelerator Physics, Graduate Texts in Physics,https://doi.org/10.1007/978-3-319-90188-6_21

251

Page 252: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

252 21 Topics in Laser-Driven Acceleration

Fig. 21.1 A particletrajectory (straight line) istilted relative to the directionof the laser beam (outlinedby the two hyperbolas)

�W = q∫ ∞

−∞v · E(r0 + vt, t)dt . (21.1)

Hence, in this approximation, we neglect the influence of the accelerating field onthe particle velocity and its orbit. This is a reasonable approximation for relativisticparticles that move with almost constant velocity close to the speed of light and dueto the relativistically increased inertia are difficult to deflect from a straight path.This approximation becomes less accurate when the dimensionless electromagneticparameter a from Chap.16 is much larger than unity.

Let us now prove that in free space far from material boundaries, and in theabsence of static fields, the above integral is equal to zero. The proof is based onthe fact that an arbitrary electromagnetic field in vacuum can be represented as asuperposition of plane electromagnetic waves propagating in all possible directions(see Problem 13.1). Mathematically, this representation is written as

E(r, t) =∫

d3k E(k)eik·r−iωt , (21.2)

where k is the wavenumber of a plane electromagnetic wave, E(k) is the amplitudeof the electric field in the wave and ω = ck.1 The integration over k is carried overthe whole space. Substituting Eq. (21.2) into (21.1) we obtain

W = qv ·∫ ∞

−∞dt

∫d3k E(k)eik·(r0+vt)−iωt

= 2π∫

d3k qv · E(k)eik·r0δ(ω − k · v) . (21.3)

The argument in the delta function in the last integral is never equal to zero, because

ω − k · v = k(c − v cosα) > 0 , (21.4)

1Here we ignore the fact that the frequency can actually be both positive and negative, ω = ±ck.This does not change the proof.

Page 253: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

21.1 The Lawson–Woodward Theorem 253

where α is the angle between the directions of v and k. Hence the integral (21.1) isequal to zero. This result can be explained as follows.As electromagneticwaves prop-agate with the speed of light they overtake particles moving with a velocity smallerthan c. As a result, a particle is slipping through the accelerating and deceleratingphases in the wave, and the total acceleration averages to zero.

There are several ways to overcome the limitations of the Lawson–Woodwardtheorem and achieve acceleration in an electromagnetic field. First, one can limit theinteraction between the particle and the waves by introducing material boundariesand preventing the field from occupying the whole space. This would make therepresentation (21.2) invalid. We consider a model for such acceleration in the nextsection. Second, for particles of relatively small energy and intense laser fields onemay need to take into consideration the wiggling of the orbit caused by the laser.This means that one has to drop the assumption of constant velocity and a straightorbit in Eq. (21.3). Such effects play an important role in the acceleration of not veryrelativistic electrons by a strong laser beam with parameter a � 1. Finally, one canintroduce an external magnetic field and intentionally bend the orbit. One exampleof such acceleration — the so-called inverse FEL acceleration — will be consideredin the last section of this chapter.

Other types of laser acceleration use some form of indirect acceleration, whichincorporates materials that mediate the interaction between the laser fields and theparticles to be accelerated; examples employ dielectrics, thin foils, and plasmas.

21.2 Laser Acceleration in Space with Material Boundaries

Wewill now calculate the energy gain of a charged particle passing through a focusedlaser field reflected back by a flatmirror as shown in Fig. 21.2, seeRef. [1]. Themirroris located at z = 0 and has a small hole for the passage of the particle.We assume thatthe hole does not perturb the laser field except for a small vicinity near the hole. In thecalculations, we neglect interactionwith the reflected part of the field, which turns outto be small (see Problem 21.2). A similar setup has been tested experimentally [2, 3].

We assume that the particle moves along a straight line parallel to the z axis withvelocity v and an offset x0. The z coordinate of the particle at time t is equal toz0 + vt . The energy gain is then given by the following integral:

�W = q∫ 0

−∞dt vEz(x0, 0, z0 + vt, t) = q

∫ 0

−∞dz Ez(x0, 0, z, (z − z0)/v) .

(21.5)

Note that the integration extends from z = −∞ to z = 0 because there is no fieldbehind the metal mirror. The longitudinal component of the electric field in the laserfocus was calculated in Problem 13.3, and is given by

Page 254: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

254 21 Topics in Laser-Driven Acceleration

Fig. 21.2 The laser beam isreflected by a metal surface;the particle passes through ahole in the metal

Ez(x, y, z, t) = − 1

ik

∂Ex

∂x= −2x

ikA(z)Q(z)eQ(z)ρ2e−iωt+ikz , (21.6)

where A and Q are given by Eqs. (13.17) and (13.18). We need to take the real partof the field to calculate �W :

�W = −Re

[2x0q

ikeikz0β

−1∫ 0

−∞dz A(z)Q(z)eQ(z)x20 e−ikz(β−1−1)

]

= Re

[2E0x0q

ikw20

eikz0β−1

∫ 0

−∞dz

(1 + 2i z/kw20)

2e−x20/w

20(1+2i z/kw2

0)e−ikz(β−1−1)

].

(21.7)

Finally, switching to the integration variable ξ = 2z/kw20 = z/ZR we obtain

�W = −Re

[i E0x0qe

ikz0β−1∫ 0

−∞dξ

(1 + iξ)2e−x20/w

20(1+iξ)e−iξk2w2

0(β−1−1)/2

]. (21.8)

We first note that if there is no mirror and the particle interacts with the laser beamfrom z = −∞ to z = ∞ the upper limit in this integral should be set to infinity. Itcan be proven that the resulting integral is equal to zero for β ≤ 1. This is, of course,in agreement with the Lawson–Woodward theorem.

Returning to the integral (21.8) let us now consider an ultrarelativistic particleand set β = 1 (here the ultrarelativistic limit corresponds to γ � kw0). Then theintegration in (21.8) is easy to carry out, and the result is

Page 255: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

21.2 Laser Acceleration in Space with Material Boundaries 255

Fig. 21.3 A particletrajectory in an undulatorand the laser pulseco-propagating with theparticle. The laser pulse isrepresented by a planeelectromagnetic wave (bluevertical lines)

�W = −Re

[i E0x0qe

ikz0

∫ 0

−∞dξ

(1 + iξ)2e−x20/w

20(1+iξ)

]

= E0qw20

x0(1 − e−x20/w

20 ) cos(kz0) . (21.9)

The factor cos(kz0) in this equation indicates that the sign of the energy change de-pends on the position of the charge relative to the phase of the laser field; for bunchesof particles longer than the wavelength of the laser radiation such an interactionmodulates the energy of the beam with the period equal to the laser wavelength.

Many more details of laser acceleration can be found in Ref. [4].

21.3 Inverse FEL Acceleration

Wewill now consider a laser-particle interaction when the particle is moving throughan undulator as shown in Fig. 21.3. This kind of laser-beam interaction is called theinverse FEL acceleration, where FEL stands for Free-Electron Laser. To simplifycalculations, we will assume that the undulator parameter K is small, K � 1.

We will represent the laser field by a plane electromagnetic wave propagating inthe z direction, with electric field Ex given by

Ex (z, t) = E0 cos(ωt − ωz/c) , (21.10)

where E0 is the wave amplitude and ω is the laser frequency. The energy change isgiven by the following formula:

�W = q∫

dt E · v = q∫

dt Exvx . (21.11)

The particle orbit in a small-K undulator has been calculated in Chap.19, seeEq. (19.3), and the velocity vx can be obtained by differentiating x with respectto time:

vx = cK

γsin(kuz) . (21.12)

Page 256: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

256 21 Topics in Laser-Driven Acceleration

Using the approximation z ≈ z0 + ct (1 − 1/2γ2) gives

�W = q∫

dt E · v

= q∫

dtcK

γsin[ku(z0 + ct)]E0 cos

[ωt − kct

(1 − 1

2γ2

)− kz0

]

≈ cqK E0

∫dt sin

[kuct − ωt + ωt

(1 − 1

2γ2

)+ (k + ku)z0

]

≈ cqK E0

∫dt sin

[(kuc − ω

1

2γ2

)t + (k + ku)z0

]. (21.13)

In this equation we discarded the term with the sum of the arguments in the sinefunction because it rapidly oscillates and adds only a small contribution to the result.The most effective acceleration occurs if

ω = 2γ2kuc , (21.14)

when the time dependent part of the sine argument cancels. This means that the laserfrequency should be equal to the frequency of the undulator radiation in the forwarddirection. In this case we find

�W = qK E0Lu

2γsin [(k + ku)z0] . (21.15)

Depending on the position of the particle z0 relative to the phase of the laser, theenergy gain can have both positive and negative signs. This will result in an energymodulation of the beam.

As a final note, we mention that the plane wave approximation is valid if theRaleigh length for the laser beam is much larger than the undulator length.

Worked Examples

Problem 21.1 Prove that Eq. (21.8) with the upper limit of integration changed toinfinity evaluates to zero for β = 1.

Solution: For β = 1, the integral in the modified form of Eq. (21.8) becomes

∫ ∞

−∞dξ

(1 + iξ)2e−x20/ω

20 (1+iξ) = − iω2

0

x20e−x20/ω

20 (1+iξ)

∣∣∣∣∞

ξ=−∞= 0 .

Problem 21.2 Estimate the contribution to�W in Eq. (21.8) from the reflected partof the laser field.

Solution: To adapt Eq. (21.8) for the reflected wave, we have to change the signof k to be −k which implies that we also replace ξ with −ξ. There is one exceptionhowever, because the ξ/β term actually came from the frequency ω instead of k, so

Page 257: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

21.3 Inverse FEL Acceleration 257

that sign does not change. Thus, instead of the quantity β−1 − 1 in the last exponentwe have β−1 + 1 ≥ 2. In the previous case, this term could become small as β → 1.Because in the paraxial approximation a laser waist will have to satisfy w0 � 1/k,therewill bemany oscillations formoderate ranges of the ξ parameter.While the finalintegral is difficult to simplify numerically, we can see that there will be a strongsuppression of the contribution from the reflected wave to the total acceleration,even when stopping the integration at ξ = 0. At its largest, this contribution shouldbe a factor of 1/(k2w2

0) = 1/(2kZR) smaller than the acceleration from the forwardswave.

Problem 21.3 Assume that you are given a laser with a given energy EL , frequencyω and duration τ of the laser pulse. Optimize the parameters of the laser accelerationexperiment shown in Fig. 21.2 to achieve the maximum energy gain for relativisticparticles. Express the energy gain in terms of EL , ω and τ .

Solution: The energy gain is given by Eq. (21.9),

�W = qE0w20

x0(1 − e−x20/w

20 ) cos(kz0) .

Depending on the sign of the charge, kz0 should be either 0 or π. For eitherx0 � w0 or x0 � w0, �W → 0. The optimum x0 corresponds to the maximum ofthe function x−1

0 (1 − e−x20/w20 ), which is found by equating the derivative to zero,

2x20w2

0

= ex20/w

20 − 1 ,

so that x0 ≈ w0 (actually x0/w0 1.12). We then see |�W | ∝ E0w0. The laserpower, PL , is proportional to E2

0w20, so �W ∝ √

PL . In terms of pulse energy, PL ≈EL/τ so shorter pulses are more effective.

Problem 21.4 Take the following parameters of the inverse FEL acceleration ex-periment from Ref. [3]: beam energy 60MeV, laser pulse length 0.55 ps, laser energy0.65mJ, laser focused spot size 200 µm, undulator period 1.8cm, number of periods3, K = 0.65; and estimate the amplitude of the energy modulation of the beam.

Solution: The energy modulation for an inverse FEL is given by Eq. (21.15),

�W = qE0K Lu

2γsin[(k + ku)z0] .

For the given parameters, γ = 118, the total undulator length Lu = 0.054 m, andthe laser peak power is PL = WL/τ = 0.65 mJ/0.55 ps = 1.18 GW. The E field isgiven by

E0 = √PL4Z0/π/w0 = 6.9 GV/m .

Page 258: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

258 21 Topics in Laser-Driven Acceleration

Plugging in above we find for the peak energy gain and loss

�W = 1.0 MeV .

References

1. A. Chao, Lecture Notes on Topics in Accelerator Physics. SLAC-PUB-9574 (2002)2. T. Plettner, R.L. Byer, E. Colby, B. Cowan, C.M.S. Sears, J.E. Spencer, R.H. Siemann, Visible-

laser acceleration of relativistic electrons in a semi-infinite vacuum. Phys. Rev. Lett. 95, 134801(2005)

3. C.M.S. Sears, E.R. Colby, R. Ischebeck, C. McGuinness, J. Nelson, R. Noble, R.H. Siemann, J.Spencer, D. Walz, T. Plettner, R.L. Byer, Production and characterization of attosecond electronbunch trains. Phys. Rev. ST Accel. Beams 11, 061301 (2008)

4. Eric Esarey, Phillip Sprangle, Jonathan Krall, Laser acceleration of electrons in vacuum. Phys.Rev. E 52, 5443 (1995)

Page 259: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

Chapter 22Radiation Damping Effects

In this chapter we will show how the radiation damping in electron and positron ringscan be added to the Hamiltonian and Vlasov formalism, and calculate how radiationdamping affects the energy, transverse actions and distribution function.

22.1 Radiation Damping in Equations of Motion

Relativistic electrons and positrons in accelerators radiate intensely when movingin a circular orbit of an accelerator. As we discussed in Chap.15, for a transversemagnetic field the radiation is emitted in the forward directionwithin an angle∼ 1/γ.The radiation reaction force acts in the opposite direction, and, similar to a frictionforce, tends to slow down the motion of the particle. It maintains the energy balancein the process: the energy of the radiation is taken from the kinetic energy of theparticle. It also causes damping of the betatron oscillations and affects the energyspread of the beam. In this section we will derive the equations that describe theeffect of the radiation on the beam dynamics.

We consider relativistic particles moving with γ � 1 in a circular accelerator. LetP be the power of radiation (the energy emitted per unit time) for a given particle at aspecified location in the ring. Since for a relativistic particle we have approximatelyp = h/c, where h is the energy, the quantity P/c is equal to the decrease of themomentum of the particle per unit time. As mentioned above, the force is acting inthe direction opposite to themomentum, sowe canwrite the change in themomentumcomponents per infinitesimally small time dt as

© Springer International Publishing AG, part of Springer Nature 2018G. Stupakov and G. Penn, Classical Mechanics and Electromagnetismin Accelerator Physics, Graduate Texts in Physics,https://doi.org/10.1007/978-3-319-90188-6_22

259

Page 260: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

260 22 Radiation Damping Effects

dpx = − pxp

Pcdt = −vx

v

Pcdt ≈ −P

c2vxdt = −P

c2dx = −P

c2dx

dsds ,

dpy = − pyp

Pcdt ≈ −P

c2dy

dsds ,

dh = −Pdt = −P dt

dsds . (22.1)

As we learned in Chap.5, without the radiation effects the particle dynamics in acircular accelerator is governed by the Hamiltonian (5.19). We now need to addthe additional changes of the momentum and energy defined by Eq. (22.1) to theHamiltonian equations. Because Eq. (22.1) only modifies the canonical momenta,the differential equations for the evolution of the coordinates remain Hamiltonian:

dx

ds= ∂K

∂ px,

dy

ds= ∂K

∂ py,

dt

ds= −∂K

∂h, (22.2)

while the original Hamiltonian equations for themomentum and energy are amendedby the terms from Eq. (22.1):

dpxds

= −∂K

∂x− P

c2∂K

∂ px,

dpyds

= −∂K

∂y− P

c2∂K

∂ py,

dh

ds= ∂K

∂t+ P ∂K

∂h. (22.3)

In these equations, we replaced the derivatives dt/ds, dy/ds and dt/ds on the right-hand sides of Eq. (22.1) by their Hamiltonian expressions Eq. (22.2). We emphasizehere that Eqs. (22.2) and (22.3) are not Hamiltonian any more, but it is convenientto keep writing them using the previously introduced canonical variables and theHamiltonian function K .

An expression for the radiation power was derived in Chap. 18, Eq. (18.27),

P(h, s) = 2

3

rcγ4mc3

ρ2(s)= 2

3

q2rch2

m3c3B2(s) , (22.4)

where the bending radius for a relativistic particle is expressed through the magneticfield, |ρ| = h/c|qB(s)|, B(s) is the magnetic field taken at the particle’s position,and rc is the classical radius. In Eq. (22.4) we indicated the explicit dependence ofP versus the coordinate s and the energy h. For the nominal energy h0, we will usethe notation

P0(s) = 2

3

q2rch20m3c3

B2(s) . (22.5)

Page 261: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

22.1 Radiation Damping in Equations of Motion 261

Averaging P0 over the ring and dividing it by the nominal energy p0c defines acharacteristic damping time τs for the ring,

1

τs= 〈P0(s)〉

p0c= 1

γmc21

cT

∮ds P0(s) = 2

3

rcγ3

T

∮ds

ρ2(s), (22.6)

where T is the revolution period and the angular brackets denote the averaging overthe ring circumference. The time τs has the meaning of the time scale needed for aparticle to lose all its initial energy to radiation, if this energy is not replenished. In atypical accelerator ring the damping time is much longer than the revolution periodT and the period of betatron oscillations. This observation will be used in the nextsection.

22.2 Synchrotron Damping of Betatron Oscillations

In this section we will consider the effect of the synchrotron damping on betatronoscillations using the machinery developed in Sect. 22.1.

First we need to transform Eqs. (22.2) and (22.3) to the variables of the linearizedHamiltonian H from Eq. (6.10) used in our analysis of the betatron oscillations1,

H ≈ −1 − η − ηx

ρ(s)+ 1

2P2x + 1

2P2y + x2

2ρ2(s)− e

p0

1

2G(s)

(y2 − x2

). (22.7)

Recall thatH was obtained from K by division by p0 with a simultaneous transitionfrom px , py to Px = px/p0, Py = py/p0, and from h to h/p0. As a result the twofirst pairs of Eqs. (22.2) and (22.3) become

dx

ds= ∂H

∂Px,

dPxds

= −∂H∂x

− Pp0c2

∂H∂Px

,

dy

ds= ∂H

∂Py,

dPy

ds= −∂H

∂y− P

p0c2∂H∂Py

. (22.8)

For relativisticmotion,we can identify h/p0 with pc/p0 = c(1 + η) (see Eq. (5.22)),which means that in the third equation in (22.3) we can use cη:

cdη

ds= ∂H

∂t+ P

p0c

∂H∂η

= − Pp0c

(1 + x

ρ

), (22.9)

where in the last equation we assumed ∂H/∂t = 0 and used Eq. (6.10) for thederivative ∂H/∂η. One immediately sees from this equation that η monotonically

1In our analysis we will assume that ρ does not depend on the transverse coordinates x and yignoring the case of the so-called combined function magnets where such dependence exists.

Page 262: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

262 22 Radiation Damping Effects

decreases with time due to the continuous energy loss to the radiation. If this loss isnot compensated the initial energy of the particle will be radiated away within a timeof the order of τs . In reality the particle energy is replenished by RF cavities in thering, with the corresponding term in the Hamiltonian given by Eq. (5.27). However,to avoid complications associated with a treatment of a time-dependent Hamiltonianwewill adopt a simpler model for the energy source that compensates for the radiatedenergy. We will assume that this source is uniformly distributed in the ring and isequal to the average over the ring circumference energy loss as defined by Eq. (22.5):

cdη

ds= − P

p0c

(1 + x

ρ

)+ 〈P0(s)〉

p0c. (22.10)

As it turns out, our results obtained with the model of a continuous energy source arevalid, to a good approximation, for real machines with localized RF cavities in thering. Since we assume that the relative energy deviation η is small we can expandthe difference P − P0 keeping only the linear term in η,

P − P0(s) ≈ p0cη∂P∂h

∣∣∣∣h=h0

= 2ηP0(s) , (22.11)

where we took into account the quadratic dependence of P versus h in Eq. (22.4).As a result, we obtain

ds= −P0(s)

p0c2

(2η + x

ρ

), (22.12)

where we have neglected second order terms such as xη and η2 as well as the quantityP0(s) − 〈P0(s)〉, which by definition is an oscillation that will average to zero duringeach turn around the ring. There are two terms on the right-hand side of this equation.The first one describes damping of the energy deviation. The mechanism of thisdamping is due to the monotonic increase of the energy loss with energy, P ∝ h2;the energy loss is higher for particles with energy larger than the nominal energy,h > h0, and smaller for energy h < h0. The second term is a driving force due to thechanges in the path length of the orbit with a nonzero offset x in comparison with thenominal one. In contrast to Eq. (22.9) this equation exhibits an equilibrium solutionwith x = η = 0.

Let us now assume that a particle has an initial energy deviation η. How does thisenergy deviation evolve with time? We know that when the tune is not an integer,an off-momentum particle oscillates around the modified closed orbit defined byEq. (8.14): x0(s) = ηD(s). Substituting this relation into Eq. (22.12) we obtain

ds= −P0(s)

p0c2

(2 + D(s)

ρ

)+ x − x0(s)

ρ

]. (22.13)

Page 263: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

22.2 Synchrotron Damping of Betatron Oscillations 263

As was mentioned at the end of the previous section, the synchrotron damping isa slow process that lasts for many revolution periods. Therefore, it makes senseto consider the cumulative effect of the damping at a given s over multiple turns.The term x − x0(s) will oscillate and average nearly to zero, so this term will beremoved. It also makes sense to average Eq. (22.13) over the circumference of thering. Recalling that P0 ∝ ρ−2 and using Eq. (22.6) it is easy to obtain

〈dη

ds〉 = − η

cτs(2 + D) , (22.14)

where D is the damping partition number,

D =(∫

ds

ρ2

)−1 ∫ds

ρ3D(s) . (22.15)

In many accelerators the parameter D is small and to the lowest approximation canbe neglected.We then have energy perturbations, on average, exponentially decayingwith a time constant equal to half of τs .

We will now consider the damping of vertical betatron oscillations in a ring dueto the synchrotron radiation. We know that without the damping, when the system isHamiltonian, the action Jy given by Eq. (7.8) is conserved. Due to the synchrotronradiation it will be slowly (over many revolution periods) decreasing with time. Tofind its damping time, we need to calculate the derivative d Jy/ds using Eq. (22.8) andaverage it over the ring circumference (as we did above for dη/ds). The calculationis simplified if we note that when P = 0, d Jy/ds = 0, and that the damping term inEq. (22.8) involves only Py . Hence

d Jyds

= ∂ Jy∂Py

×[−P0(s)

p0c2∂H∂Py

]= −P0(s)

p0c2(βPy + αy)Py (22.16)

(to simplify the notation we dropped the index y in β and α). We then use Eq. (7.10)to obtain

d Jyds

= −P0(s)

p0c2[2Jy (sin φ + α cosφ)2 − 2αJy cosφ (sin φ + α cosφ)

]

= −2P0(s)

p0c2Jy

[sin2 φ + α sin φ cosφ

]. (22.17)

We now average this expression over the ring circumference. The averaging consistsof two parts. First we note that at subsequent revolutions in the ring the phase φat a given location s increases by the betatron phase advance 2πν (see Sect. 7.3).Assuming that the tune is not close to an integer or half-integer, the angular part ofthe expression (22.17) effectively averages over the angle φ in a small number ofturns, such that the change in Jy can be ignored over this time scale. Therefore we cantake sin2 φ → 1

2 and sin φ cosφ → 0. After that we need to do one more averaging

Page 264: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

264 22 Radiation Damping Effects

of P0(s) over s, again neglecting the change in Jy during one revolution around thering:

⟨d Jyds

⟩= −〈P0〉

p0c2Jy = − 1

cτsJy . (22.18)

We see that the vertical betatron oscillations decay exponentially with a time constantequal to τs , whichwe assume to bemuch greater than the revolution period T .We alsoignore the role of the energy offset, which would add an overall factor of (1 + 2η)that is a higher order correction that will be ignored.

Calculation of the damping of the betatron oscillations in the horizontal plane ismore complicated. This complication comes from the fact that the betatron oscilla-tions are coupled to the energy through the term −ηx/ρ in the Hamiltonian (6.10)and, in addition, the evolution of η is coupled to x through the term x/ρ in Eq. (22.12).The closed orbit for off-energy particles is also distorted. As before, we take intoaccount that d Jx/ds = 0 whenP = 0, and the damping comes through the variablesPx and η. However, to take into account the fact that there is an orbit distortion asso-ciated with η, we have to use the action Jx defined by Eq. (8.16) instead of Eq. (7.8)to find that

d Jxds

= ∂ Jx∂Px

×(

−P0(s)

p0c2∂H∂Px

)+ ∂ Jx

∂η

ds(22.19)

= − P0(s)

p0c2Px

[β(Px − ηD′) + α(x − ηD)

]

− P0(s)

p0c2

(2η + x

ρ

)1

β

{− D(x − ηD)

− (βD′ + αD

) [β(Px − ηD′) + α(x − ηD)

] }.

We also must generalize the transformation to action-angle coordinates, replacingEqs. (7.9) and (7.10) with:

x − ηD = √2β Jx cosφ ,

Px − ηD′ = −√2Jxβ

(sin φ + α cosφ) . (22.20)

We can use these expressions to recast the evolution of Jx in terms of action-anglevariables:

Page 265: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

22.2 Synchrotron Damping of Betatron Oscillations 265

d Jxds

= − P0(s)

p0c2

{2Jx

[sin φ(sin φ + α cosφ) − D

ρcos2 φ (22.21)

+1

ρ(βD′ + αD) sin φ cosφ

]

+ η√2β Jx

[−D′ sin φ + 1

β

(2 + D

ρ

) (−D cosφ + (βD′ + αD) sin φ)] }

.

We now perform the two-step averaging using the same approach as for the verticalaction Jy above. At this point, we ignore any variations in η along the ring as beinga higher order correction. Averaging over the phase φ leaves only two terms on theright-hand side,

d Jxds

= −P0(s)

p0c2Jx

(1 − D

ρ

), (22.22)

and then averaging over the circumference of the ring gives

⟨d Jxds

⟩= − 1

cτsJx (1 − D) . (22.23)

We see that the horizontal betatron oscillations decay with the time constantτs/(1 − D).

From the results of this section it follows that the synchrotron radiation causesenergy deviations and betatron oscillations to damp down. This however does notmean that in a real machine they completely disappear after several damping times.There are various driving sources for these oscillations, and the most prominentone is due to the so-called quantum diffusion effect. The effect is explained bythe quantum nature of the synchrotron radiation and the recoil momentum that aparticle receives when it emits a single synchrotron photon. The competition betweenthe synchrotron damping and excitation of betatron and synchrotron oscillationsestablishes a dynamic equilibrium and in many cases determines the energy spreadand horizontal emittance of the beam in circular electron accelerators. In the absenceof vertical bends, the equilibrium vertical emittance can be much smaller and is oftendetermined by magnetic field errors.

22.3 Vlasov Equation and Robinson’s Theorem

In Chap.10 we introduced the kinetic equation to describe the evolution of an ensem-ble of beam particles. While our initial formulation of the continuity equation (10.7)was general and valid for arbitrary equations of motion, the subsequent assumptionof Hamiltonian motion leads to the Vlasov equation (10.11) and the even more ele-gant Eq. (10.14). Our goal now is to include the effect of the synchrotron damping,as described by Eqs. (22.8) and (22.9), into the formalism of the Vlasov equation.

Page 266: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

266 22 Radiation Damping Effects

The distribution function f now depends on 7 variables: f (x, Px , y, Py, t, η, s).With an evident generalization of equation (10.11) for 3 degrees of freedom, thekinetic equation is

∂ f

∂s+ ∂

∂x

(dx

dsf

)+ ∂

∂Px

(dPxds

f

)+ ∂

∂y

(dy

dsf

)+ ∂

∂Py

(dPy

dsf

)

+ ∂

∂t

(dt

dsf

)+ ∂

∂η

(dη

dsf

)= 0 . (22.24)

Let us calculate the full (or convective) derivative d f/ds:

d f

ds= ∂ f

∂s+ dx

ds

∂ f

∂x+ dPx

ds

∂ f

∂Px+ dy

ds

∂ f

∂y+ dPy

ds

∂ f

∂Py+ dt

ds

∂ f

∂t+ dη

ds

∂ f

∂η.

(22.25)

Substituting the partial derivative ∂ f/∂s from Eqs. (22.24) into (22.25) we obtain

d f

ds= − f

(∂

∂x

dx

ds+ ∂

∂Px

dPxds

+ ∂

∂y

dy

ds+ ∂

∂Py

dPy

ds+ ∂

∂t

dt

ds+ ∂

∂η

ds

).

(22.26)

The Hamiltonian part of the equations of motion (22.8) does not contribute to theright-hand side of Eq. (22.26). Following the notation from Chap.10 of writing thenon-Hamiltonian part of the kinematic equations in terms of damping functions Fi ,as in (10.32), this can be written as

d f

ds= − f

(∂Fx

∂Px+ ∂Fy

∂Py+ ∂Fη

∂η

)= 4 f

P0(s)

p0c2, (22.27)

where

Fx = − Pp0c2

∂H∂Px

−P0(s)

p0c2Px ,

Fy = − Pp0c2

∂H∂Py

−P0(s)

p0c2Py ,

Fη = Pp0c2

∂H∂η

−P0(s)

p0c2

(2η + x

ρ

), (22.28)

where we have used Eqs. (22.8), (22.12), and the linearized Hamiltonian.According to Eq. (22.27) the distribution function f at a phase space point mov-

ing with a particle grows exponentially with time. This happens because, due to thesynchrotron radiation, the phase space volume occupied by a given ensemble of par-ticles decreases. Since f is the particle density in the phase space, it grows inverselyproportionally to the phase space volume. This effect is associated with the name

Page 267: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

22.3 Vlasov Equation and Robinson’s Theorem 267

of K. Robinson who pointed it out in [1]. As was already mentioned in the previ-ous section, in reality, due to the various diffusion sources that we neglected in ouranalysis, this exponential growth eventually stops and a steady state is established.

Reference

1. Kenneth W. Robinson, Radiation effects in circular electron accelerators. Phys. Rev. 111(2),373–380 (1958)

Page 268: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

Appendix AMaxwell’s Equations, Equations of Motion,and Energy Balance in an ElectromagneticField

A.1 Maxwell’s Equations

Classical electrodynamics in vacuum is governed by the Maxwell equations. In theSI system of units, the Maxwell equations are

∇ · D = ρ , (A.1a)

∇ · B = 0 , (A.1b)

∇ × E = −∂B∂t

, (A.1c)

∇ × H = j + ∂D∂t

, (A.1d)

where ρ is the charge density, j is the current density, D = ε0E and H = B/μ0.Traditionally B is called the magnetic induction, and H is called the magnetic field,but in this book we refer to B as the magnetic field. The Maxwell equations arelinear: a sum of two solutions, E1, B1 and E2, B2, is also a solution correspondingto the sum of densities ρ1 + ρ2, j1 + j2.

For a point charge q moving along a trajectory r = r0(t) the charge density andthe current density are

ρ(r, t) = qδ(r − r0(t)) , j(r, t) = qv(t)δ(r − r0(t)) , (A.2)

with v(t) = d r0(t)/dt .To find a particular solution of the Maxwell equations in a volume, proper bound-

ary conditions should be specified at the volume boundary. On a surface of a goodconducting metal the boundary condition requires the tangential component of theelectric field to be equal to zero, Et |S = 0.

© Springer International Publishing AG, part of Springer Nature 2018G. Stupakov and G. Penn, Classical Mechanics and Electromagnetismin Accelerator Physics, Graduate Texts in Physics,https://doi.org/10.1007/978-3-319-90188-6

269

Page 269: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

270 AppendixA: Maxwell’s Equations, Equations of Motion, and Energy Balance …

A.2 Wave Equations

In free space with no local charges and currents the electric field satisfies the waveequation,

1

c2∂2E∂t2

− ∂2E∂x2

− ∂2E∂y2

− ∂2E∂z2

= 0 . (A.3)

The same equation is valid for themagnetic field B. A particular solution of Eq. (A.3)is a sinusoidal wave characterized by the frequency ω and the wave vector k,

E = E0 sin(ωt − k · r) , (A.4)

where E0 is a constant vector perpendicular to k, E0 · k = 0, and ω = ck.

A.3 Vector and Scalar Potentials

It is often convenient to express the fields in terms of the vector potential A(r, t)and the scalar potential φ(r, t):

E = −∇φ − ∂A∂t

,

B = ∇ × A . (A.5)

Substituting these equations into Maxwell’s equations, we find that the second andthird equations are satisfied identically. We only need to take care of the first andfourth equations.

A.4 Energy Balance and the Poynting Theorem

The electromagnetic field has an energy and momentum associated with it. Theenergy density of the field (energy per unit volume) is

u = 1

2(E · D + H · B) = ε0

2(E2 + c2B2) . (A.6)

The Poynting vector,

S = E × H , (A.7)

Page 270: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

AppendixA: Maxwell’s Equations, Equations of Motion, and Energy Balance … 271

Fig. A.1 Point chargesshown by red dots moveinside volume V and interactthrough the electromagneticfield

gives the energy flow (energy per unit area per unit time) in the electromagnetic field.Consider charges that move inside a volume V enclosed by a surface A, see Fig. A.1.The Poynting theorem states

∂t

∫Vu dV = −

∫Vj · E dV −

∫An · S d A , (A.8)

where n is the unit vector normal to the surface and directed outward. The left-handside of this equation is the rate of change of the electromagnetic energy due to theinteraction with moving charges. The first term on the right-hand side is the workdone by the electric field on the moving charges. The second term describes theelectromagnetic energy flow from the volume through the enclosing surface.

A.5 Photons

The quantum view on electromagnetic radiation is that the electromagnetic field isrepresented by photons. Each photon carries the energy �ω and the momentum �k,where the vector k is the wavenumber which points to the direction of propagationof the radiation, � = 1.05 × 10−34 J· sec is the Planck constant divided by 2π, andk = ω/c.

Page 271: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

Appendix BLorentz Transformations and the RelativisticDoppler Effect

B.1 Lorentz Transformation and Matrices

Consider two coordinate systems, K and K ′. The system K ′ is moving with velocityv in the z direction relative to the system K (see Fig. B.1). The coordinates of anevent in both systems are related by the Lorentz transformation

x = x ′ ,y = y′ ,z = γ(z′ + βct ′) ,

t = γ(t ′ + βz′/c) , (B.1)

where β = v/c, and γ = 1/√1 − β2.

The vector (ct, r) = (ct, x, y, z) is called a 4-vector, and the above transformationis valid for any 4-vector quantity. The transformation from K to K ′ is also a Lorentztransformation, but the original frame K has a velocity −v relative to the system K ′,so the inverse transformation is obtained from Eq. (B.1) by changing the sign of β:

Fig. B.1 Laboratory frameK and a moving frame K ′

© Springer International Publishing AG, part of Springer Nature 2018G. Stupakov and G. Penn, Classical Mechanics and Electromagnetismin Accelerator Physics, Graduate Texts in Physics,https://doi.org/10.1007/978-3-319-90188-6

273

Page 272: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

274 Appendix B: Lorentz Transformations and the Relativistic Doppler Effect

x ′ = x ,

y′ = y ,

z′ = γ(z − βct) ,

t ′ = γ(t − βz/c) . (B.2)

The Lorentz transformation (B.1) can also be written in the matrix notation

⎛⎜⎜⎝xyzt

⎞⎟⎟⎠ =

⎛⎜⎜⎝1 0 0 00 1 0 00 0 γ cβγ

0 0 βγc γ

⎞⎟⎟⎠

⎛⎜⎜⎝x ′y′z′t ′

⎞⎟⎟⎠ = L

⎛⎜⎜⎝x ′y′z′t ′

⎞⎟⎟⎠ , (B.3)

where L denotes the 4 × 4 matrix in the middle of the equation. The advantage ofusing matrices is that consecutive transformations reduce to matrix multiplication.

B.2 Lorentz Contraction and Time Dilation

Two events occurring in the moving frame at the same point and separated by thetime interval �t ′ will be measured by the lab observers as separated by �t ,

�t = γ�t ′ . (B.4)

This is the effect of relativistic time dilation.An object of length l ′ aligned in the moving frame with the z′ axis will have the

length l in the lab frame:

l = l ′

γ. (B.5)

This is the effect of relativistic contraction. The length in the direction transverse tothe velocity is not changed.

B.3 Relativistic Doppler Effect

Consider awave propagating in amoving frame K ′. It has the time-space dependence:

∝ cos(ω′t ′ − k′ · r ′) , (B.6)

where ω′ is the frequency and k′ is the wavenumber of the wave in K ′. An observerthat measures this wave in the reference frame K will see the time-space dependence

Page 273: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

Appendix B: Lorentz Transformations and the Relativistic Doppler Effect 275

that is obtained from the Lorentz transformation of coordinates and time in Eq. (B.6):

cos(ω′t ′ − k′ · r ′) = cos(ω′γ(t − βz/c) − kx′x − ky

′y − kz′γ(z − βct))

= cos(γ(ω′ + kz′βc)t − kx

′x − ky′y − γ(kz

′ + ω′β/c)z) . (B.7)

We see that in the K frame this process is also a wave

∝ cos(ωt − k · r) , (B.8)

with the frequency and wavenumber

kx = kx′ ,

ky = ky′ ,

kz = γ(kz′ + βω′/c) ,

ω = γ(ω′ + βckz′) . (B.9)

Hence, the combination (ω, ck) is a 4-vector.The above transformation is valid for any typeofwaves (electromagnetic, acoustic,

plasma waves, etc.) Let us now apply it to electromagnetic waves in vacuum. Forthese waves we know that

ω = ck . (B.10)

Assume that an electromagnetic wave propagates at angle θ′ in the frame K ′,

cos θ′ = k ′z

k ′ , (B.11)

and has a frequency ω′ in that frame. What is the angle θ and the frequency ω ofthis wave in the lab frame? We can always choose the coordinate system such thatk = (0, ky, kz), then

tan θ = kykz

= k ′y

γ(kz ′ + βω′/c)= sin θ′

γ(cos θ′ + β). (B.12)

In the limit γ � 1 almost all angles θ′ (except for the angles very close to π) aretransformed to angles θ ∼ 1/γ. This explains why radiation of an ultrarelativisticbeam goes mostly in the forward direction, within an angle of the order of 1/γ.

A general expression for how the Lorentz transformation changes the angle θ ofan electromagnetic wave with respect to the direction of the velocity is:

cos θ′ = cos θ − β

1 − β cos θ, sin θ′ = sin θ

γ(1 − β cos θ). (B.13)

Page 274: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

276 Appendix B: Lorentz Transformations and the Relativistic Doppler Effect

For the frequency, a convenient formula relates ω with ω′ and θ (not θ′). To deriveit, we use the inverse Lorentz transformation

ω′ = γ(ω − βckz) = γ(ω − βck cos θ) , (B.14)

which gives

ω = ω′

γ(1 − β cos θ). (B.15)

Assuming a largeγ and a small angle θ and usingβ ≈ 1−1/2γ2 and cos θ = 1−θ2/2,we obtain

ω = 2γω′

1 + γ2θ2. (B.16)

The radiation in the forward direction (θ = 0) gets a large factor 2γ in the frequencytransformation.

B.4 Lorentz Transformation of Fields

The electromagnetic field (E, B) is transformed from K ′ to K according to thefollowing equations:

Ez = E ′z , E⊥ = γ

(E′

⊥ − v × B′) ,

Bz = B ′z , B⊥ = γ

(B′

⊥ + 1

c2v × E′

), (B.17)

where E′⊥ and B′

⊥ are the components of the electric and magnetic fields perpendic-ular to the velocity v: E′

⊥ = (Ex , Ey), B′⊥ = (Bx , By).

The electromagnetic potentials (φ/c, A) are transformed exactly as the 4-vector(ct, r):

Ax = A′x ,

Ay = A′y ,

Az = γ(A′z + v

c2φ′

),

φ = γ(φ′ + vA′z) . (B.18)

Page 275: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

Appendix B: Lorentz Transformations and the Relativistic Doppler Effect 277

B.5 Lorentz Transformation and Photons

It is often convenient, even in classical electrodynamics, to consider electromagneticradiation as a collection of photons. How do we transform the parameters of a photonfrom K ′ to K ? The answer is rather evident: the combination (k,ω) constitutes a4-vector and is transformed according to Eq. (B.9). This is of course in agreementwith the fact that the pair (�k, �ω) is the momentum-energy 4-vector for the photon.The number of photons in K ′ to K is the same — it is a relativistic invariant.

Page 276: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

Index

AAccelerator tune, 80Action, 4Action-angle variables, 33Adiabatic invariant, 53Adiabatic process, 52Anharmonicity, 55

BBeam emittance, 128Beta function, 80Betatron oscillations, 78Betatron phase, 79

CCanonical transformation, 21, 23Canonically conjugate variables, 9Circular polarization, 164Classical radius, 204Closed orbit, 63Closed orbit distortions, 96Coherent radiation, 246Compton scattering, 209Configuration space, 4Courant-Snyder invariant, 90Curvilinear coordinates, 65Cut-off frequency, 174Cyclotron frequency, 8Cylindrical resonator, 176

DDamping partition number, 263Diffraction radiation, 218

Dipole magnet, 75Dipole radiation, 202Dispersion function, 98Dynamic aperture, 117

EElectromagnetic field pressure, 179Elliptic polarization, 164Energy deviation, 98Extended phase space, 125

GGaussian beams, 165Generalized coordinates, 4Generalized momentum, 9Generating functions, 25, 27

HHamiltonian, 9Hamiltonian flow, 36Harmonic oscillator, 3, 47Hill’s equation, 79

IIncoherent radiation, 246Incompressible flow, 123Integral of motion, 12Inverse Compton scattering, 209Inverse FEL acceleration, 255

KKinetic equation, 123

© Springer International Publishing AG, part of Springer Nature 2018G. Stupakov and G. Penn, Classical Mechanics and Electromagnetismin Accelerator Physics, Graduate Texts in Physics,https://doi.org/10.1007/978-3-319-90188-6

279

Page 277: Gennady Stupakov Gregory Penn Classical Mechanics and ...webéducation.com/wp-content/uploads/2021/02/Graduate...Preface This book is a graduate-level textbook covering topics in classical

280 Index

LLagrangian, 4Lawson–Woodward theorem, 251Leontovich boundary condition, 153Liénard–Wiechert potentials, 195Light pressure, 209Linear polarization, 164Liouville’s theorem, 38Long-thin approximation, 142Longitudinal formation length, 241Lorentz detuning, 181Lorentz factor, 7Loss factor, 184

MMathieu equation, 51Mathieu functions, 51

NNon-conservative forces, 39Nonlinear oscillator, 53Nonlinear resonance, 53

PParametric resonance, 51Paraxial approximation, 166Pendulum equation, 4, 53Perfect conductivity, 154Phase mixing, 129Plane electromagnetic waves, 163Poisson bracket, 12, 23

QQuadrupole errors, 99Quadrupole magnet, 76Quality factor, 48, 178

RRadiation damping time, 263Radiation field, 191Radiation reaction force, 204Random force, 49Rayleigh dissipation function, 39Rayleigh length, 167Resonance, 49

Resonance overlapping, 112Resonant width, 49Retarded potentials, 198Retarded time, 192Robinson’s theorem, 267

SSeparatrix, 54Sextupole magnet, 76Shielding effect, 244Skew quadrupole magnet, 76Skin depth, 153Skin effect, 151Slater’s formula, 182Small amplitude approximation, 68Space charge effect, 144Spectrum of radiation, 216Spectrum of synchrotron radiation, 226Spherical electromagnetic wave, 202Spherical wave, 169Standard map, 113Surface impedance, 153Symplectic map, 36, 37Synchrotron radiation, 221

TThird-order resonance, 107Thomson cross section, 206Transition radiation, 213Transverse electric (TE) modes, 175Transverse formation length, 244Transverse magnetic (TM) modes, 173

UUndulator, 231Undulator parameter, 232

VVlasov equation, 124Vlasov-Fokker-Planck equation, 131

WWeak focusing, 78Wiggler, 232