Download as pdf
Download as pdf
You are on page 1of 621
Theoretical mate) A lee Volume |: Fauugiy leuk s623525 oe Ss = = T. Padmanabhan ‘Theoretical Astrophysics Volume I: Astrophysical Processes Graduate students and researchers in astrophysics and cosmology need a solid under- standing of a wide range of physical processes. Ths clear and authoritative textbook has been designed to help them to develop the necessary toolkit of theory. Assuming only an undergraduate background in physics and no detailed knowledge of astronomy, this ‘book guides the reader step by step through a comprehensive collection of fundamental theoretical topies. The book is modular in design, allowing the reader to pick and choose a selection of chapters, if necessary. It can be used alone or in conjunction with the forth- coming accompanying two volumes (covering stars and stellar systems and galaxies and cosmology, respectively), After the basics of dynamics, electromagnetic theory, and statistical physics art re viewed, asolid understanding of ll the key concepts such as radiative processes, spe fluid mechanics, plasma physies and magnetohydrodynamics, dynamics of gravi systems, general relativity, and nuclear physics is developed. Each topic is devel ‘methodically from undergraduate basic physics. Throughout, the reader's undUxstan is developed and tested with carefully structured problems and helpful hit D This welcome volume provides graduate students with an indispensable tion run row Psst ig tiventy one Syl Quantum Theory. He has published over hundred technical paps jese areas and has written four books: Structure Formation in the Universe, Cogomofogy and Astrophysics Through Problems, After the First Three Minutes, and, with LV. Narlikar, Gravity, Gauge Theories and Quantum Cosmology. He is a member of the Indian Academy of Scien« and International Astronomical Union. He has the Shanti Swarup Bhatnagar Prize in Physics (I! awarded by the Council of Scientific and Industfial’Research, India, Professor Padmanabhan has also written #iogfiin 100 popular science articles, a ‘comic strip serial, and several regular columns on astronomy, recreational mathematics, and history of science that have appeared in international journals and papers. He is ‘married, has one daughter, and lives in Pune, India. His hobbies include chess, origami, and recreational mathematies. ial Academy of Sciences, jumerous awards, including the Millenium Medal (2000) & THEORETICAL ASTROPHYSICS Volume I: Astrophysical Processes T. PADMANABHAN Inter-University Centre for Astonomy and Astophysis Pune, India & & Ss & ° CAMBRIDGE @ UNIVERSITY PRESS PUBLISHED BY THE PRESS SYNDICATE OF THE UNIVERSITY OF CAMBRIDGE ‘The Pitt Building, Trumpington Stet, Cambridge, United Kingdom ‘The Edinburgh Building, Cambridge CB2 2RU, UK _htp/vrwcup.cam ek +40 West 20h Stoet, New York, NY 10011-4211, USA hapu/lwww-up.org 10 Stamford Road, Oakleigh, Melbourne 3166, Australia. Ruiz de Alar 13,28014 Madrid, Spain (© Cambridge University Press 2000 “This book sin copyright Subject to statutory exception and othe provisions of relevant collective licensing agreements, ‘no reproduction of any part may take place without {he writen permission of Cambridge University Press. Firs published 2000, Printed inthe United States of America Tpeface Times Roman 11/13 pl. System TEX" [78] A catalog record for this book s avilable from the British Libra Library of Congress Cataloging in Publication Data Padmanabhsn,T (Than), 1957= ‘Theoretical astrophysics astrophysical processes / : Padmanabhan, © poem ISBN 0-521-56240-6 1. Astrophysics. 1. Title ©G QB461.833 2000 S301 ded 00-025837 ISBN 0521 56240 6 hardback ISBN 0 521 565320 papetback Dedicated tothe memory of LD. Landau, ‘who understood the importance of pedagogy & © a Q o THEORETICAL ASTROPHYSICS ~ in three volumes — VOLUME I: ASTROPHYSICAL PROCESSES 1: Order-of-magnitude astrophysics; 2: Dynamics; 3: Special Relativity, Electro- ‘dynamics, and Optics; 4: Basics of Electromagnetic Radiation; 5: Statistical Mechanics, 6: Radiative Processes; 7: Spectra; 8: Neutral Fluids; 9: Plasma Physics; 10: Gravitational Dynamics; 11: General Theory of Relativity; 12: Basics of Nuclear Physics. & VOLUME II: STARS AND STELLAR SYSTEMS 1: Overview: Stars and Stellar Systems; 2: Stellar Structure; 3: Stellar Gay 4: Supernova; 5: White Dwarfs, Neutron Stars, and Black Holes; 6: P ry Stars and Accretion; 8: Sun and Solar System; 9: Interstellar Mediu Globular jes and Interactions; Clusters 4: Friedmann Model; 5: Active Galactic Nuclei-Structural 6: Thermal History of the Universe; 7: Structure Formation; 8: Cosmic Migrowive Background Radiation; 9: Formation of Baryonie Structures; 10: Active Galacii¢Nueei-Cosmological Aspects; 11: Intergalactic Medium and Absorption Systems; ological Observations VOLUME III: GALAXIES AND COSMOL Contents Preface Chapter 1 Order-of-Magnitude Astrophysics 1 12 13 Introduction Energy Scales of Physical Phenomena 1.2.1 Rest-Mass Energy 1.2.2 Atomic Binding Energies 1.23 Molecular Binding Energy 1.2.4 Nuclear Energy Scales 1.2.5 Gravitational Binding Energy 1.2.6 Thermal and Degeneracy Energies of Particles Classical Radiative Processes 1.3.1 Thermal Bremsstrahlung 1.3.2 Synchrotron Radiation Radiative Processes in Quantum Theory 1.4.3 Thermal Radiation 1.5.1 toot * tyav: Existence of Galaxies 1.5.2 €yav © €q: Existence of Giant Plat 1.5.3 gy ~ nue: Existence of Stars 1.5.4 Existence of H-R Diagram fe 1.5.5 yay = €¢: Existence of Ste} Detecting the Photons 1.6.1 Role of Earth’s Atmosp] 1.6.2 Radio 1.63 Microwave and Submillimeter 1.64 Infrared ° nts vii 1.4.1 Fine Structure and Hyperfine Structure Ss" 1.4.2 Transition Rates and Cross Sections § xv viii Contents 1.6.5 Optical and Ultraviolet 37 1.6.6 X-ray and y ray 38 Chapter 2. Dynamics 2 2.1 Introduction 2 2.2. Time Evolution of Dynamical Systems 2 2.3 Examples of Dynamical Systems 47 2.3.1 Motion Under a Central Force 4 2.3.2 Motion in a Rotating Frame 48 2.3.3 The Reduced Three-Body Problem st 2.4 Canonical Transformations 54 2.5 Integrable Systems 60 2.6 Adiabatic Invariance 70 2.7. Perturbation Theory for Nonintegrable Systems B 2.8 Surface of Section n Chapter 3 Special Relativity, Electrodynamics, and Optics 83 3.1 Introduction 8 3.2. The Principles of Special Relativity 8 3.3. Transformation of Coordinates and Velocities 3.4 Four Vectors 3.5. Particle Dynamics 3.6 Distribution Functions and Moments 3.7 External Fields of Force 3.8 Motion of Charged Particles in External Fields : © 104 3.8.1 Motion in a Coulomb Field 3.8.2 Motion in a Constant, Uniform, Electric Field 3.8.3 Motion in a Constant, Uniform, Magnetic Fiel 107 3.8.4 Motion in a Slowly Varying Magnetic Field 3.8.5 Drifts in Magnetic Fields 3.9. Maxwell’s Equations 113 3.10 Energy and Momentum of the Electromagneti 116 3.11 Time-Independent Electromagnetic Fields 18. 3.11.1 Coulomb Field of a Charged Parti 119 3.11.2 Dipole and Multipole Moments 120 3.11.3 Magnetic Field of a Steady 121 3.11.4 Maxwell's Equations in a le Medium 122 3.12 Electromagnetic Waves 123 3.12.1 Monochromatic Plane. 124 3.12.2 Polarisation of Light 126 3.13 Diffraction 129 3.13.1 Fraunhofer Diffraction (r >> D®/2) 132 3.13.2 Fresnel Diffraction (r < D?/2) 134 3.14 Interference and Coherence 136 3.15 3.16 Contents Linear Optical Systems ‘Wave Propagation through a Random Medium Chapter 4 Basics of Electromagnetic Radiation 41 42 43 44 45 Introduction Radiation from an Accelerated Charge 4.2.1 Why Does an Accelerated Charge Radiate? 4.2.2 Relativistically Invariant Derivation General Properties of the Radiation Field 4.3.1 Radiation in the Nonrelativistic Case 4.3.2 Radiation in the Relativistic Case 4.3.3 Radiation During an Impulsive Motion 4.3.4 Relativistic Formula for Radiated Four Momentum. Radiation Reaction Quantum Theory of Radiation 4.5.1 Quantisation of an Electromagnetic Field 4.5.2 Interaction of Matter and Radiation 4.5.3 Quantum Dipole Radiation Chapter 5 Statistical Mechanics Sa 52 33 34 55 56 37 58 39 5.10 Sal 5.12 53 5.14 Introduction Operational Basis of Statistical Mechanics The Density of States and Microcanonical Distribution Mean Values in Canonical Distribution Derivation of Classical Thermodynamics Partition Function for Bosons and Fermions Fermions 59.1 Classical Limit: 2 <1 5.9.2 Quantum Limit: < >> 1 Bosons © Statistical Mechanics of the Electromagnetig Fie Ionisation and Pair-Creation Equilibria o 5.12.1 Ionisation Equilibrium for Hy. 5.12.2 Pair Creation 5.12.3. Nuclear Statistical Equilib Time Evolution of Distribution Ful Evolution under Scattering o Description of Macroscopic Thermodynamics we Quantum Statistical Mechanics Chapter 6 Radiative Processes 6.1 62 63 Introduction Macroscopic Quantities for Radiation Absorption and Emission in the Continuum Case 140 144 150 150 150 130 153 187 137 160 162 163 164 167 168 173, 17 196 200 207 212 217 218 220 225 228 233 233 239 240 241 245 251 251 251 263 64 65 66 67 68 69 6.10 6.11 6.12 6.13 Contents Scattering of Electromagnetic Radiation Radiation Drag on a Charged Particle Compton Scattering and Comptonisation Kompaneets Equation Equations of Radiative Transport Bremsstrahlung 6.9.1 Classical Bremsstrahlung 6.9.2 Quantum Bremsstrahlung, 6.9.3 Thermal Bremsstrahlung 6.9.4 Free-free Absorption ‘Synchrotron Radiation: Basics ‘Synchrotron Radiation: Rigorous Results 6.11.1 Angular Distribution of Radiation 6.11.2 Spectral Distribution of Radiation 6.11.3 Radiation from a Power-Law Spectrum of Electrons 6.11.4 Synchrotron Self-Absorption Photoionisation Collisional lonisation Chapter 7 Spectra 1A 72 13 14 15 16 1 Chapter 8 Neutral Fluids 8 82 83 84 Introduction Width of Spectral Lines 7.2.1 Natural Width of Spectral Lines 7.2.2 Doppler Width of Spectral Lines 7.23 Collisional Broadening of Spectral Lines we Curve of Growth S Atomic-Energy Levels 7.4.1 Energy Levels in the Nonrelativistic Theory 7.4.2 Fine Structure of Energy Levels, © 7.4.3. Hyperfine Structure of Energy Levels 3S 7.4.4 X-ray Emission From Atoms Selection Rules © Energy Levels of Diatomic Molecules Aspects of Diatomic Spectra 7.7.1 Rotational Transitions 7.7.2 Vibrational-Rotational Sp. 7.1.3 Electronie—Vibrational-Rot Introduction Molecular Collisions and Evolution of the Distribution, Function Stress Tensor for an Ideal Fluid Stress Tensor for a Viscous Fluid 265 am 22 ann 285 295 295 298 300 302 304 307 308 309 312 313 315 324 331 335 339) 339) 343 345 348 349 351 356 356 358 359) 361 361 361 367 368 Contents 8.5 Equations of Motion for the Viscous Fluid 8.6 Flow of Ideal Fluids 8.6.1 Barotropic Fluids 8.6.2 Steady Flows of Ideal Fluids 8.6.3 Irrotational, Isoentropic Flow of Ideal Fluids 8.6.4 Incompressible, Irrotational Flow of Ideal Fluids 8.7. Flow of Viscous Flui 8.7.1 Incompressible Flow of Viscous Fluids 8.7.2 Scaling Relations in Viscous Flows 8.8 Sound Waves. 8.9 Supersonic Flows 8.9.1 de Laval Nozzle 8.9.2 Spherical Accretion 8.10. Steepening of Sound Waves 8.11 Shock Waves 8.12 Sedov Solution for Strong Explosions 8.13 Fluid Instabilities 8.13.1 Rayleigh—Taylor Instability 8.13.2 Kelvin-Helmholtz Instability 8.13.3 Thermal Instability 8.14 Conduetion and Convection 8.15 Turbulence 372 378 380 380 382 383 383 383 384 385 387 388 391 397 401 407 409 410 Chapter 9 Plasma Physics Ga 9.3. Collisions in Plasmas 9.4 Collisiontess Plasmas 9.5 Waves in Magnetised Cold Plasmas 9.5.1 Propagation Along the Magnetic Field 9.5.2 Propagation Perpendicular to the Magnet 9.5.3 Propagation Along a General Direction 9.6 Magnetohydrodynamies 9.6.1 The assumptions and Equations of 9.6.2. Batteries and Generation of the 9.6.3 Ambipolar Diffusion 9.6.4 Magnetic Virial Theorem 9.7 Hydromagnetic Waves 9.7.1 Alfven waves 9.7.2 Fast and Slow Magnetohydrodynamic Waves 9.1 Introduction we 428 9.2 The Mean Field and Collisions in Plasma S 428 tohydrodynamics ic Field Chapter 10 Gravitational Dynamics 10.1 Introduction 10.2 Gravitational Interaction in Astrophysical Systems, 433, 440 452 454 436 456 458 458 462 463 466 467 468 469 474 474 474 xii Contents 10.3. Self-Gravitating Barotropic Fluids 476 10.3.1 Polytropic Equation of State 4 103.2 Degenerate Fermionic Systems 479 10.3.3 Isothermal Sphere 480 103.4 Time-Dependent Isothermal Sphere Solutions 485 10.3.5 Fluid Spheroids 486 10.4 Collisionless Gravitating Systems in Steady State 489 10.5 Moment Equations for Collisionless Systems 493 10.6 Approach of a Collisionless System to Steady State 494 10.7, Collisional Evolution 499 10.8 Dynamical Evolution of Gravitating Systems 504 Chapter 11 General Theory of Relativity 508 11.1 Introduction 508 11.2 Inescapable Connection between Gravity and Geometry 508 11.3. Metric Tensor and Gravity 513 11.4 Particle Trajectories in a Gravitational Field S17 11.5. Physics in Curved Space-Time 522 11.6 Curvature 52 11,7 Dynamics of Gravitational Field 5 11.7.1 Action for the Gravitational Field 1172 Pld Equations rhe Gavia el @) 11.7.3 Properties of the Energy-Momentum Tensor 3 11.74 General Properties of the Gravitational- Field Equations we 535 11.8 Schwarzschild Metric S 338 11.9 Orbits in the Schwarzschild Metric 543 11.9.1 Precession of the Perihelion 546 11.9.2 Deffection of an Ultrarelativistic Particle 548 11.93 Post-Newtonian Precession 550 11.10 Gravitational Collapse and Black Holes © 352 11.11 The Energy-Momentum Pseudotensor for Gravi 553 11.12 Gravitational Waves o 554 Chapter 12 Basics of Nuclear Physics 563 12.1 Introduction 563 12.2 Nuclear Structure 563 12.3, Thermonuclear Reactions 567 12.3.1 Nonresonant Reaction 570 12.3.2 Resonant Reaction Rates 572 12.4 Specific Thermonuclear Reactions 317 12.4.1 Hydrogen Burning: PP Chain 517 Contents 12.4.2 Hydrogen Burning: CNO Cycle 12.43 Helium Burning 12.4.4 Burning of Heavier Elements 12.4.5 Neutron Capture Reactions Notes and References Index xiii 580 581 583 584 987 593 Preface “..yoyum varo gudham anupravisto, naanyam thasman Nachiketa vrinithe.” (.Nachiketa does not choose any other boon but [learning about] that of which Knowledge is hidden.") Katho Upanishad, Verse During the past decade or so, theoretical astrophysics has emerged as one ‘most active research areas in physics. This advance has also been reflectg@finhé greater interdisciplinary nature of research that is being carried outin thiS eeu in the recent years. Asa result, those who are learning theoretical a the aim of making a research career in this subject need to assimil amount of concepts and techniques, in different areas of astrophi period of time. Every area of theoretical astrophysics, of textbooks that allow the reader to master that particular cs ‘well-defined ‘way. Most of these textbooks, however, are written ina tra fe, focussing on one area of astrophysics (say stellar evolution, gal namics, radiative processes, cosmology ete.) Because different authors, ee perspectives regarding their subject matter it is not very easy(for a,student to understand the key unifying principles behind several diffefent’ystrophysical phenomena by studying a plethora of separate textbooks, do not link up together as a series of core books in theoretical astrophysits covering everything which a student would need. A few books, whic! et the whole of astrophysics, deal with the subject at a rather elem “first course”) level. What we require is clearly some jalogous to the famous Landau~ Lifshitz course in theoretical physics, but Yocussed to the subject of theoretical astrophysics at a fairly advanced level. In such a course, all the key physical concepts (¢-g., radiative processes, fluid mechanics, plasma physics, ete.) can be xv xvi Preface presented from a unified perspective and then applied to different astrophysical situations. This book isthe first ofa set of three volumes that are intended to do exactly that. They form one single coherent unit of study through the use of which a student can acquire mastery over all the traditional astrophysical topics. What is ‘more, these volumes emphasise the unity of concepts and techniques in different branches of astrophysics. The interrelationship among different areas and com- mon features in the analysis of diferent theoretical problems will be stressed throughout, Because many of the basic techniques need to be developed only once, itis possible to achieve a significant economy of presentation and crispness of style in these volumes Needless o say, there are some basic “boundary conditions” one has to respect in such an attempt to cover the whole of Theoretical Astrophysics in approxi- rately 3 x 580 pages. Not much space is available to describe the nuances in greater length orto fill in the details of algebra. For example, I have made con- scious choices as to which parts of the algebra can be left to the reader and Which need to be worked out explicitly in the text, and I have omitted a detailed discussion of elementary concepts and derivations. However, Ido not expect the reader to know anything about astrophysics, Al astrophysical concepts are d veloped ab initio in these volumes. The approach used in these three volumes similar to that used by Gengis Khan, namely, (1) cover as much area as pos (2) capture the important points, and (3) be utterly ruthless! ws) ‘To cut out as much repetition as possible, the bulk of the physical rg are presented at one go in Vol. I and are applied in the other different situations. This implies that there will be a lot of litte of “concrete” astrophysics in Vol. I; that comes in Vol. I ( Systems) and Vol. Il (Extragalactic Astronomy and Cosmo for the election of material for Vol. have been the followis topics that are used in only a specific chapter in Vol. in situ rather than in Vol. I. (2) By and large, everythingtistussed in Vol. I will be utilized directly somewhere in Vols. II and IIf.Qp rare occasion, I do cover a topic in Vol. I even if it is not fully utilized iN ology are, of course, discussed theoretical aspects. Observation and phi ce the motivation clear. However, in Vols. IL and III to the extent necessar reliability, etc., ofthe observations or the astronomical techniques. (Maybe there should be a fourth volume describing observational astrophysics!) ‘The target audience for this three-volume work will be fairly large and is made up of (I) students in the first year of their Ph.D. Program in theoretical physics, Preface xvii astronomy, astrophysics, and cosmology; (2) research workers in various fields of theoretical astrophysics, cosmology, etc.; and (3) teachers of graduate courses in theoretical astrophysics, cosmology and related subjects. In fact, anyone working in or interested in some area of astronomy or astrophysics will find something useful in these volumes. They are also designed in such a way that parts of the ‘material can be used in modular form to suit the requirements of different people and different courses, Let me briefly highlight the features which are specific to Vol. I. The reader of Vol. Lis assumed to have done basic courses in classical mechanics, nonrelativis- tic quantum mechanics, and classical electromagnetic theory. Of the 12 chapters in Vol.I, the first one is a broad-brush overview of physical principles in an order- ‘of- magnitude manner and is intended to set the stage. I expect the reader to survey this chapter rapidly but to come back to it periodically at later stages. This chapter is probably the easiest or the most difficult, depending on one’s background and aptitude. It is easy in the sense that very little sophisticated mathematics is used; difficult because it takes a high level of maturity to appreciate some of the phys- ical arguments that are presented. Chapters 2 (Dynamics), 3 (Special Relativity, Electrodynamics, and Optics), and 5 (Statistical Mechanics) cover the EeC the reader may already be familiar with — but from an advanced perspective. Ths aim is to introduce powerful techniques in familiar contexts so that the reader c learn and appreciate them. For example, no apologies are made for introdt four-vector notation up front or dealing with distribution functions waht ees) the beginning, so as to get the main results as quickly as possible. The offipt throughout is on topics relevant in astrophysics, such as the reduced tk problem, action-angle variables, diffraction and interference, systems, propagation in random media, ionisation equilibria, ete. Chapter; ith the basics of radiation theory —both classical and quantum — that i iped from scratch and the reader is not assumed to be familiar with qui eld theory. Chapters 6-12 develop the toolkit for astrophysics in a sefCapiained manner, virtually ab initio. Chapters 6 (Radiative Processes) and shanics) are fairly exhaustive and detailed. The short chapter on. Kate (Chap. 7) covers general material that is of astronomical relevance; mokéspptific aspects will be dealt with in Vols. Il and III within the appropriate@ontexts. In Chap. 9 (Plasma Physics) I had to make choices as to which tofR@se of sufficiently general nature to appear in Vol. I; some specific topicsfe-BDinstability of axisymmetric systems with magnetic fields, alpha effect, amos) will appear in the rel- ‘evant chapters of Vols. Il and III. Chaptegt@YGiravitational Dynamics) covers the background needed for galactic dyes, globular cluster evolution, ete. ‘Chapter 11 is a compact introduction:ée\geveral relativity and no previous fa- miliarity with tensor analysis is assumed. Finally, Chap. 12 deals with aspects of nuclear physics that are needed in the study of stellar evolution. Any one of these topics is fairly vast and often requires a full textbook to do justice to it, whereas I have devoted approximately 60 pages to each of them! xviii Preface I would like to emphasise that such a crisp, condensed discussion is not only possible but also constitutes a basic matter of policy in these volumes. After all, the idea is to provide the student with the essence of several textbooks in one place. It should be clear to lecturers that these material can be easily regrouped to serve different graduate courses at different levels, especially when complemented by other textbooks. Because of the highly pedagogical nature of the material covered in Vol. 1,1 have not given detailed references to original literature except on rare occasions ‘when a particular derivation is not available in standard textbooks. The annotated list of references given at the end of the book cites several other textbooks that I found very useful. Some of these books, of course, contain extensive bibliographies and references to original literature. The selection of core books cited here clearly reflects the personal bias of the author and I apologise to anyone who feels that their work or contribution has been overlooked. Several people have contributed to the making of these volumes and especially to Vol. I. The idea for these volumes originated over a dinner with JP. Ostriker in late 1994, while I was visiting Princeton. I was lamenting to Jerry about the lack of a comprehensive set of books covering all of theoretical astrophysics and Jerry said, “Why don’t you write them?” He was very enthusiastic and supportive o the idea and gave extensive comments and suggestions on the original outlin produced in the next one week. I am grateful to him for the comments any the moral support that I needed to launch into such a project. I sincere ‘the volumes do not disappoint him. Ch ‘Adam Black of Cambridge University Press took up the proposal acteristic enthusiasm and initiative; this isthe third project on wl ith’ helped to mould the proper framework Many of my friends and colleagues carried out the job drafts and providing comments. Of these, M. Vivekanan of the book with meticulous care and has given extensi colleagues, especially Roger Blandford, George Djot , Peter Goldreich, John Huchra, Donald Lynden-Bell, JV. Narlikar, Ry Nityananda, Sterl Phinney, and Douglas Richstone looked at the whole dr provided comments and suggestions at different levels of detail. JS. ai Iyer, Nissim Kanekar, Ben Oppenheimer, K. Subramanian, S. S: arayanan, and K. Srinivasan gave detailed comments on selected chay é last two took pains to check most of the derivations and algebraic exptégsions. I thank all of them for their help. Thave been visiting the Astronomy Department of Caltech during the past several years and the work on this book has benefitted tremendously through my discussions and interactions with the students and staff of the Caltech Astronomy Department, I especially thank Roger Blandford, Peter Goldreich, Shri Kulkarni, Preface xix Sterl Phinney, and Tony Readhead for several useful discussions and for sharing with me their insights and experience in physics teaching, This project would not have been possible but for the dedicated support from ‘Vasanthi Padmanabhan, who not only did the entire TEXing and formatting but also produced most of the figures — often writing the necessary programs for the same. | thank her for the help and look forward to receiving the same for the next two volumes! I also thank Sunu Engineer who was resourceful in solving several computer-related problems that cropped up periodically. It is a pleasure to acknowledge the library and other research facilities available at the Inter- University Centre for Astronomy and Astrophysics, which were useful in this task, T, Padmanabhan & © a Q o 1 Order-of-Magnitude Astrophysics 1.1 Introduction ‘The subject of astrophysics involves the application of the laws of physics to large macroscopic systems in order to understand their behaviour and predict new phenomena. This approach is similar in spirit to the application of the laws of physics in the study of, say, condensed-matter phenomena, except for the following three significant differences: () tee ios om ofan on pm trophysics than in, say, condensed-matter physics. It is not possible to s systems under controlled conditions so that certain physical processes. inate the behaviour. Identifying the causes of various observed pheriGingya in astrophysics will require far greater reliance on statistical argumen in laboratory physics. ee (2) The astrophysical systems of interest span a wide range of pj space and require inputs from several different branches of phys ally, the densities can vary from 10-25 gm em? (interstellar mi to 10" gm ‘em’ (neutron stars); temperatures from 2.7 K (microwiqveé¢kground ra- diation) to 10° K (accreting x-ray sources) or even to I ty universe); radiation from wavelengths of meters (radio waves) lons of angstroms (hard gamma rays); typical speeds of particles can Btexs/t0 0.99 (relativis- tic jets). Clearly we require inputs from quantu regimes as well as from more familiar classi (3) The primary source of information abou the electromagnetic radiation detected complete picture about any source, it wave bands. Because of technologigal limitations, it is often quite diffi- cult to have uniform coverage acetss:tie entire electromagnetic spectrum Hence the information we have abou the sources is often distorted or incomplete. iechanical and relativistic ies, astrophysical sources is m. Therefore, to obtain a ssary to examine it in all the 2 1 Order-of- Magnitude Astrophysics These considerations suggest that two aspects will be most important in the study of different astrophysical systems. The first is the appreciation of the different states in which bulk matter can exist under different conditions and the dynamics of the matter governed by different equations of state. The sec- ‘ond is an understanding of different radiative processes that lead to the emis- sion of photons, which act as prime carriers of information about astronomical objects. ‘We shall be concerned with these and related topics in several chapters of this book. The purpose of this introductory chapter is twofold: It will first pro in Sections (1.2) ~ (1.4) — a rapid overview of several physical processes at an order-of-magnitude level and introduce the necessary concepts. Then we will ‘make an attempt to understand the existence of different astrophysical structures from first principles to the extent possible. Implementing such a plan, of course, has not been possible even in laboratory physics, and itis unlikely to succeed in the case of astrophysics. At present, astrophysics does require a fair amount of observational and phenomenological input, just like any other branch of applied physics. Nevertheless, we will make such an attempt as itis useful in providing the most basic and direct connection between physics and astrophysics. 1.2 Energy Seales of Physical Phenomena & Let us consider a system of N particles (V > 1), each of mass m, oc a spherical region of radius R. In dealing with the dynamics of suct\a léfge collection of particles, itis useful to introduce the concept of pi by the system of particles as the momentum transferred per sec toa (fictitious) surface of unit area. The contribution tothe rate of mor iransfer (per unit area) from particles of energy € is n(€)p(€) - v(e), ) denotes the number of particles per unit volume with momentum p elocity v(e) ‘We obtain the net pressure by averaging this expression 0 gles defined by p (or v) and summing over all values of the energy. € the momentum and the velocity are parallel to each other, the vector luct p - v averages to (1/3)pv (in three dimensions), giving “w cd p=if n(e)ple) (ay 0 sm is called ideal ifthe kinetic ‘the particles. In that case € is ith the relations, where the integration is over all energies. ‘energy dominates over the interaction ent essentially the kinetic energy of the p; ay -12 (7 —Dme?, v= (1-3) . (12) where ¢ is the kinetic energy of the particle, the pressure can be expressed in the paym, ¢€ 1.2 Energy Scales of Physical Phenomena 3 ire me? me)! te ne(4 F )(+25) de (13) In the nonrelativistic (NR) limit (with mc? > €), this gives Pyx (2/3) (ne) = (2/3)Unr, where Uy is the energy density (ie., energy per unit volume) of the particles. In the relativistic case (with € >> mc? or when the particles are mass- less), the corresponding expression is Pex * (1/3) (ne) =(1/3)Unr. Henee, in general, P~U up to a factor of order unity. This result can be converted into a more useful form of equation of state whenever the mean free path of the particles in the system is small compared with the length scales over which the physical parameters of the system change significantly. Then the pressure can be expressed in terms of density and tempe- rature if the energy density can be expressed in terms of these variables. This is possible in several contexts leading to different equations of state. To understand each of these cases it is useful to start by identifying the characteristic energy scales of bulk matter. We now tur to this task. ee & ‘We can associate the rest-mass energy mc? with each individual partic] «ai ‘mass m. In normal matter, made up of nucleons and electrons, the lowe; isnigstly ae, form for rest-mass is provided by electrons with mec? ~0.5 MeV. For nucle rest-mas energy is mpc? ~ 1 GeV. Because the total mass ofthe ys due to the nucleons, the total rest-mass energy will be Emass = NAD where Am, mis the mass of each nucleus and Nm = M isthe t system. Rest-mass energy is extensive that is, Ems N ~ phenomena in which masses of individual nuclei do not chayi 1.2.2 Atomic Binding Energi If the particles of the system have internal struct tc.) then we get further energy scales that are cl The simplest is the atomic binding energy of at from the electromagnetic coupling between The Hamiltonian describing an electro, nucleus of charge Zq, is given by Ho =. is described by a wave function ¥/( ere L denotes the characteristic scale over which y varies significantly, then the expectation value for the en- ergy of the electron in this state is E(L) = (| Holy”) © (i? /2m_L?) — (Zq?/L). The first term arises from the fact that (¥Ip?|W) = —h? (WIV?|p) = (h?/L?), which is equivalent to the uncertainty principle stated in the form p =h/L. This jolgeular, atomic, nuclear, istic of the interactions. ‘molecules, which arises ig in the Coulomb field of a 1e) — (Zq?/r). If this electron 4 1. Order-of- Magnitude Astrophysics expression for E(L) reaches a minimum value of Emin =—Z?€q when L is var- ied, with the minimum occurring at Lmin = (ao/Z), where he “052x108 em, & a?m.c? © 13.6eV, (4) with the definitions x. =(h/m.c) and « = (q?/hc). ao and €, correspond to the size and the ground-state energy of a hydrogen atom with Z=1. The wave- length 2 corresponding to €, is k= (hc/€q) = 2a~*k, ~ 103A and lies in the UV band. The fine-structure constant & * 7.3 x 107? plays an important role in the structure of matter and arises as the ratio between several interesting variables: a = (v/Zc) = (2u8/qao) = (ro/Re) = (Xe/aa), where v is the speed of an electron in the atom, y= (gh/2mec) is the Bohr ‘magneton representing the magnetic moment ofthe electron, and ro = (q2/mec*) is called the classical electron radius. When atoms of size ap are closely packed, the number density of atoms is ‘soa © (2dg)~3 ~ 10% em-5. The binding energy of such a solid arises essen- tially because of the residual electromagnetic force between the atoms, and & typical binding energy per particle is fe, with f ~(0.1-1). 1.2.3 Molecular Binding Energy © The simplest molecular structure consists of two atoms bound to eacththe})in the form of a diatomic molecule. The effective potential energy U(r) between the atoms in such a molecule arises from a resid coupling and has a minimum at a separation r ~ ao, approximé(ely the atom. The depth of the potential well at the minimum jgz0hpdrable with the electronic-energy level ¢, of the atom. In addition to the‘hter\h binding energies of the atoms comprising the molecule, thed= me)" VRE RIC — me? { where the two forms are applicable in the non-relativistic (NI relativistic (ER) limits. The behaviour of the system depends o} momentum distribution of the particles. ‘The familiar situation is the one in which short-range called ‘collisions’) between the particles effectively cae to randomize the momentum distribution. This will happen the effective mean free path of the system / is small compared with the scale L at which physical parameters change, (The explicit form takéy byhe condition 1 < L can be very different in different cases; this conditioR @Wiscussed in detail towards the end of this section.) When such a system is. iy state, we can assume that the local thermodynamic equilibrium, chard@tetized by a local temperature T, exists in the system. Then the probabilit ‘éupying a state with energy E will scale as P(E) exp[—(E/kaT)] mc?; ER) (19) me? me? 1.2 Energy Seales of Physical Phenomena 7 In this case, the momentum and the kinetic energy of the particles vanish when T>0. ‘The situation is actually more complicated for material particles like electrons. ‘The mean energy ofa system of electrons will not vanish even at zero temperature because electrons obey the Pauli exclusion principle, which requires that the ‘maximum number of electrons that can occupy any quantum state be two, one with spin up and another with spin down. Because the uncertainty principle requires that Ax Ap, 2h, we can associate (d°xd° p)/(2zh)> microstates with a phase-space volume d?x d? p. Therefore the number of quantum states with momentum less than p is [V(4s-p?/3)/(2h)°], where V is the spatial volume available for the system. The lowest energy state will be the one in which the NV electrons fill all levels up to some momentum pr, called Fermi momentum, This requires that *-(7) giving pr =A(327n)"/5. Itis obvious that if py 2 mc the system must be treated as relativistic, even in the zero-temperature limit. The energy corresponding & (1.10) (nob) (ney. Qany 3 pr will be Ph _(®) ga,ea ep = y phe? + meet = me? 5 = (FE) anny aR) pre= (he)3r?n) — (ER) erate), and the classical theory will be valid for er (h/me)-> (ER), respectively. In the first case [7 10°! cm™3), the system is nonrelativistic; it will also be degenerat (€r/kaT)> 1 and classical if Ra <1. The transition occurs at Rg ~* 1, which corresponds to nT~*/2 [(mks)*/?/h?] =3.6 x 10! in cgs units, In the second case [n > (h/ mec)? ~ 10°! cm™; p=mpn 10’ gmem™}, electrons have pr >> mc and are 8 1 Order-of- Magnitude Astrophysics relativistic irrespective of temperature, The quantum effects will dominate ther mal effects if kT <(he)n'/?, and we will have a relativistic, degenerate gas. In general, the kinetic energy of the particle will have contributions from the temperature as well as from Fermi energy. If we are interested in only the asymptotic limits, we can take the total kinetic energy per particle to be ~er(n) +kgT. Note that such a system has a minimum energy Nep(n) even atT =0. By using our general result P ~ne [see Eq, (1.3)], we can obtain the equation of state for the different cases discussed above. First, for a quantum-mechanical gas of fermionic particles with kg7 < er and € ~ er, it follows from Eq, (1.11) that P= nep varies as the (5/3)rd power of density in the nonrelativistic case and as the (4/3)rd power of density in the relativistic case. Whether the system is relativistic or not is decided by the ratio (pr/mc) or ~ equivalently —the ratio (¢/me?), The transition occurs atx = npg ~ (h/me)~>. Second, ifthe system is classical with ky T > €r so that ¢=kgT, then P ~nkgT in both nonrelativistic and extreme relativistic limits. The energy scale of the individual particles also characterizes the energy involved in the collisions between the particles. If this quantity is larger than the binding energy of the atomic system, the atoms will be ionised and the og will be separated from the atoms. The familiar situation in which this happensis high temperatures with kT’ € when the system will be made of free ele and positively charged ions, whereas, if ky 7 <¢q, the system will be sea) The transition temperature at which nearly half the number of atoms ar¢foniged oceurs around kgT ~ (€q/10), which is ~10* K for hydrogen. Fog T aK, the kinetic energy of the free electrons in the hydrogen plasma The electrons can be stripped off the atoms in another differ occurs if the matter density is so high that the atoms are pac other, with the electrons forming a common pool with €¢ electrons will be quantum mechanical and the relevant et will be €r. The temperature does not enter into the pictu may call this a zero-temperature plasma. Conventional degenerate. For normal metals in the laboratory the Fe with the binding energy within an order of magnitud ergy is comparable Ifthe temperature is below -mned by Fermi energy. energy. To treat a plasma as ideal, itis energy of ions and electrons be negligib between the ions and the electrons in-he: plasma is given by cou Zq?n'*. If the classical high-temperature plasma is to-be treated as an ideal gas, this energy should be small compared with the energy scale of the particle «kg T, which requires the condition nT? < (kg /Zq?)’ ~2.2 x 1082 in egs units. On the other hand, to treat the high-density quantum gas as ideal, we should require that 1.2. Energy Scales of Physical Phenomena 9 the Coulomb energy €cou * Zq?n' be small compared with the Fermi energy €r © (f?/2m)n2/>, The condition now becomes n > 8Z3ay*~ Z? x 10° em™?. Note that such a system becomes more ideal at higher densities; this is because the Fermi energy rises faster than the Coulomb energy. Let us now go back to the tacit assumption we made in the above analysis, viz,, that physical interactions between the particles of the system are capable of maintaining the thermal equilibrium. Determining the precise condition that will ensure this is not a simple task; but — naively — we would require that (1) the ‘mean free path for particles, 1 = (no)~' based on a relevant scattering process ‘governed by a cross section cy, be small compared with the scale L over which various parameters change significantly, and (2) that the mean time between col- lisions t = (nvc)~! be small compared with the time scale over which physical parameters change. To apply this condition we need to know the relevant mean free path for the system. For a neutral gas of molecules, this is essentially determined by molecu- lar collisions with a) ~:2ra? 8.5 x 10°!” em? and! =(nao)~!. The time scale for the establishment of a Maxwellian distribution of velocities will be approxi- matelly Tyey ~1/v «cn! TY’, For an ionized classical gas, the cross section for scattering is decided by Coulomb interaction between charged particles. Becaus an ionized plasma is made of electrons and ions with vastly different inertia, t interparticle collisions can take different time scales to produce thermal e brium between electrons, between ions, and between electrons and ions. these needs to be discussed separately. The typical impact parameter between two electrons is b~ (QZq?)miz#*), where v is the typical velocity of an electron. The correspondin; Setting cross section is + (113) and the mean free path varies as / ‘mean free time between the electron-electron scattering will be tee wo)", where n is the number density of electrons and o © rb?, This gives te © (m2v3/2nZ2q'n), which is the leading dependence. (A more precise alysis changes the numerical coefficient and introduces an extra logarithmi ; see Chap. 9.) Note that r ocm?v3 oc T3/2m'/? at a giver ature T oc (1/2)mv?, There- fore the ion-ion collision time scale tp», Wi Fer by the factor (mp /m,)!/2~ 43, giving typ =(mp/me)' tee ~ 43: The time scale for significant tran *hergy between electrons and ions is still larger because of the following Fatt. When two particles (of unequal ‘mass) scatter off each other, there is no energy exchange in the centre-of-mass frame. In the case of ions and electrons, the centre-of-mass frame differs from the lab frame only by a velocity voy ~ (me/mp)!/?vy Kup. Because there is 10 1 Order-of- Magnitude Astrophysics no energy exchange in the centre-of-mass frame, the maximum energy transfer in the lab frame (which occurs for a head-on collision) is approximately AE = (1/2) p(2vem)? = 2m pq = 2m_v3, giving [AE /(1/2)mpv;] = (m./mp) <1, ‘Therefore it takes (m p/m.) times more collisions to produce equilibrium be- tween electrons and ions, that is, the time scale for electron-ion collision is Tpe = (mp/ Me Tee ~ 1836%e. The plasma will relax to a Maxwellian distribu- tion in this time scale. Finally, it must be noted that in a high-temperature tenuous plasma, this mean free path can become larger than the size of the system. If that happens, it is necessary to check whether there are any other physical processes that can provide an effective mean free path that is lower. Most astrophysical plasmas host magnetic fields that make the charged particles spiral around the magnetic- field lines. We can estimate the typical radius ofa spiraling charged particle in a ‘magnetic field by equating the centrifugal force (mv?/r) to the magnetic force (qvB/c). This leads to a radius called the Larmor radius, given by (mev/qgB) = 13 om (T'/10° K)(B/1 Gy! in a thermal plasma. When the Larmor radius is small, it can act as the eff tive mean free path for the scattering of charged particles. The ratio betwee the mean free path from Coulomb collisions [1 « (T?/n)] and the Larmor dius (rz (T"?/B)] varies as (BT?/?/n) and can be large in tenuous ‘temperature plasmas with strong magnetic fields. This ratio is unity for: magnetic field - TY) in Se 4 10 *o(arx) (5): ‘tay The magnetic field in most astrophysical plasmas will be lj Be, and. hence this effect will be important. 1.3 Classical Radiative Proc v from the radiation received from them. Torelatetedforation received through the electromagnetic waves tothe properties ofthe amulting system, itis necessary to understand the process of electromagnetic tadiation from different systems and the nature of the spectrum emitted by 2aSfro? them, Inclassical electromagnetic theory, radi ‘emitted by any charged particle that is in accelerated motion. A detaifel@iment given in Chap. 3 shows that the total amount of energy radiated: FN Sond in all directions by a particle with charge q moving with acceleration a is given by dé 2g, dt 1 (1.15) 1.3 Classical Radiative Processes ul provided the acceleration is measured in the frame in which the particle is instan- teneously at rest. The rate of energy emission, of course, is independent of the frame used to define it. This result is called Larmor’s formula and can be used to understand a host of classical electromagnetic phenomena. Because d= (qx) is the dipole moment related to an isolated charge located at position x, this formula shows that the total power radiated is proportional to the square of d. In bounded motion, if d varies at frequency « (so that d= —«*d), then the energy radiated is given by (1.16) Different physical phenomena are essentially characterised by different sources of acceleration in Eq, (1.15) for the charged particle. Let us consider two specific examples. 1.3.1 Thermal Bremsstrahlung As a first example consider a scattering between an electron of mass me and a proton, with an impact parameter b and relative velocity v in a hydrogen plasm; The acceleration of the electron is a~(q?/m¢b2) and lasts for a time (b, Such an encounter will result in the radiation of energy € ~(q?a/c>)(b/| (q®/chm2b*v) = (Sn; /c3m2v), as b~n; \” on the average. The total/erie radiated per unit volume will be me. Because each collision ae ie (b/»), there will be very little radiation at frequencies greater [le For @< (v/b), we may take the energy emitted per unit frequency inten nearly constant. Further, in the case of plasma in thermal equilibrium, » Pym.) Putting all these together, we get q co kT). = (sciraw)= (a (ite (in This process is called thermal bremsstrahlung. Thetbrengsstrahlung spectrum is flat for 0 < w S(kaT /a) and will fall rapidly for @2V4 57 /h), where the upper limit comes from the fact that an electron withaBausal energy of (kg T) cannot emit photons with energy higher than kT / ‘otal energy radiated, over all frequencies, from such a plasma can be fi ‘integration of this expression over @ in the range (0, kg '/h). This give dé tain) dé g \(mekaT \'? (ar) f 40( aoarav) = (ts) ( rg ) ae 12 1 Order-of Magnitude Astrophysics 1.3.2 Synchrotron Radiation Another major source of acceleration for charged particles is the magnetic fields hosted by plasmas. This process, called synchrotron radiation, can be estimated as follows: Consider an electron moving with velocity v in a magnetic field B. To use Eq. (1.15) we need to estimate the acceleration in the instantaneous rest frame of the particle. In that frame, the magnetic force (q/c)[v B] is zero, but the magnetic field in the lab frame will lead to an electric field of magnitude E’~yB (see Chap. 3) in the instantaneous rest frame of the charge, inducing an acceleration a’ =(gE’/me). Accelerated by this field, the charged particle will radiate energy at the rate (which is the same in the rest frame and in the lab frame) dé _ (df) _2@ 4 202 (@? aio (ie)= 3507-55 (Gre) a ‘The power radiated by an electron of energy € = ym,c? is (dE /dt) ce? B*, Fur- ther, the energy density in the magnetic field is Up = (B?/87-);hence we can write dé) _ 16 p\? (F) 7 (23) y?cUp = (orcUs)y”, we (8/3)(q2/m Tin) ‘The total radiation emitted from a bunch of partigles wiltbe j, o<(B7«2)[n(€)] (de /de), where the first factor, Be”, is the ener the second factor is the number of particles wi the Jacobian (de /dw) cce~! cw"? from ‘a power law n(e) = . de \? Pee Smid where we have reintroduced all the constants. (A more precise calculation mul- tiplies the expression by a p-dependent factor, which is ~0.1.) This leads to a power-law spectrum, jy x v-* with a =(p ~ 1)/2. ited by a single particle, ye, and the last factor is 1 spectrum of particles is rum will be BPH 2y-(0-D/2, (122 1.4. Radiative Processes in Quantum Theory B 14 Radiative Processes in Quantum Theory The radiation field in quantum theory is described in terms of photons, and the emission or absorption of radiation arises when a physical system makes the transition from one energy level to another. We have already determined the main ‘energy levels of atoms and molecules in Subsection 1.2.2. The transition between these energy levels in an atom will correspond to photon energies upwards of few electron volts, and the corresponding wavelength will be in optical and UV bands in most of the cases, This estimate of atomic-energy levels was, however, based on the simple form of the Hamiltonian for the electron in an atom. The actual Hamiltonian is a lot more complicated than Ho used in Subsection 1.2.2. The corrections to Ho lead to splitting of the original energy levels and allow emission of photons of widely different frequencies by the atomic system. We now consider these corrections. 1.4.1 Fine Structure and Hyperfine Structure The Hamiltonian for an electron in a hydrogen atom can be expressed as a sum, H © Hy + Hpi + Hoyo + Hop~sp, Where Hy = (p?/2m,) — (Zq?/r) is the orig- inal (zeroth-order) Hamiltonian and the rest are lowest-order corrections. Thi first correction Hai = —(p*/8m3c2) is the relativistic correction to the kine energy p?/2m, that arises as the second term in the Taylor series expansic €(p) in Eq. (1.8); the correction H.p-or arises from the coupling betwes spin magnetic moment of the electron 1, = (gh/2m,c) and the magne{é BX (v/c)E =(v/c)(Zq/r?) in the instantaneous rest frame of theyelect tained by transformation of the Coulomb field. This should hax por = WB 2? mee where I and s are the orbital and the spin angular momé)tacaéthe electron. The actual result is half of this value where the extra, “aries due to a phenomenon called Thomas precession, to be discuss: ip. 3, exercise 3.4. The next correction, by y Hay-sp = Me (4 ): (1.24) is the coupling between the nuclear magneti snt and the magnetic moment (1.25) 14 1 Order-of Magnitude Astrophysics where j2 =(gh/2m,c) is called the Bohr magneton, jy is the corresponding quantity for the proton, and s and S are the spin vectors of electron and proton, respectively. It is now easy to evaluate the order of magnitude of these corrections. The zeroth-order term Ho is of the order of (q?/ao) * 10 eV, corresponding to 4 ~ 1200 A. The first correction gives rc, fa~F(£) ~(:) © watey~ lev, — (1.26) me mee. c) ay with the corresponding photon wavelength of 4 1 mm, The second correction, with |~s~h, is ag 2 o~t(4) wa?Eo, (2 imac2ay ~ ap \ mean which is of the same order as that of the first correction. These two together are called fine-structure corrections. The third correction is of the order of 22 me ith the MBN 9, MoH 2e( a ) © 1050? By 10% eV, (1.28 a my a mp \meeaa) ap (corresponding to 4~ 107 cm), which is smaller and is called the hy, correction. A more precise calculation gives the wavelength of radiation ¢1 in the hyperfine transition of hydrogen to be ~21 cm. This radiation, (hich is in the radio band, is used extensively in astronomy as a diagnoytig of wtathic (neutral) hydrogen. o 1.4.2 Transition Rates and Cross Sectionsy To complete the quantum-mechanical analysis, we also estimate the rate of transition between the various levels. Consider fal State with an atom in ground state |G) and m photons present. Abs photon, the atom makes the transition to the excited state |E), leaving an (n ~ 1) photon, state. Let the probability for this process be P[|G; The fact that this absorption probability P is tively acceptable. Consider now the probabil cess [|E)|n—1) + |G)In)]. By principl expect P’ =P, giving P’ «x, Qn A{E)In = I] xn = On tional to n seems intui- for the time-reversed pro- ieroscopic reversibility, we =m, we get P'{|E)|m) — |G) vero even for m = 0; P'[|E)|0) > for a process conventionally called spontaneous emission. The term Qm gives the corresponding probability for stimulated emission. Thus the fact that absorption probabilities are proportional ton whereas emission probabilities are proportional to + | originates from the principle of microscopic reversibility. 1.4. Radiative Processes in Quantum Theory 15 The basic rate, governed by Q, can be estimated by use of the correspondence with classical theory. If the rate of transition between two energy levels is Q and the energy of the photon emitted during the transition is ho, then the rate of energy emission is Ohco. Classically, the same system will emit energy at the rate (da /c3), where d is the electric dipole moment of the atom. (This assumes that the radiation is predominantly due to direct coupling between the radiation and the dipole moment of the atom; if not, we have to use the relevant moment like the electric quadrupole moment or magnetic dipole moment, instead of d — in this equation.) Writing Qha ~ (d«/c), we find that 2a a8 (2\"! aS mec? ox vt $() F(7E) =o. (1.29) ne ~ 8 \e aa where we have used d = gap ~ (qke/a) and hw ~ (1/2)a? mec®. The rate Q can also be expressed in different forms as, om where rp =(q?/mec”) is the classical electron radius. A more precise quantur ‘mechanical calculation corrects this by contributing an extra numerical factor: (called oscillator strength). The transition rate for other processes, such as the 21-cm radiation, can ‘mated in the same manner by using Eq. (1.16). In the relation Oho ~( , ‘we now have to use the magnetic dipole moment of electron d (aft gives Q©a(%¢/c)*e,. Estimating hon: from expression (1.28 for the hyperfine-structure energy level, we can evaluate t const) ea ow 9 op The ratio between this rate and the transition rate primary energy levels of the hydrogen atom computed in relation (1 (130) 4 (1.32) ses. te has a probability of decaying spontaneously to the ground state wit rate Q. Hence the lifetime At of the excited state is approximately ols (2) = 107s. (1.33) 16 1 Order-of-Magnitude Astrophysics for optical radiation. (It is correspondingly larger and is approximately Ar = nt cm ~ 10'S § for 21-cm radiation.) From the uncertainty principle between the energy and time, it follows that the energy level of the excited state will be uncertain by an amount AE of the order of (h/t) ~ hQ. In the absence of such ‘an uncertainty the transition between the two energy levels could lead to infinitely sharp spectral lines ata specific frequency «. The width ofthe excited states leads toa corresponding width to the spectral line called the natural width Ac, where (Ae/w)=(Q/w)* 107 for the main hydrogen lines. In terms of wavelength, the natural width A2 = 2mc(Aw/o*) is of the order of the classical electron radius: M2 27r. Because of the finite linewidth, it is convenient to introduce a (frequency- dependent) bound-bound cross section o44(«) for the absorption of radiation by the system, which is sharply peaked at «= ow with a width of Aw. If the phase~ space density of photons is dN =n[d°xd* p/(2mh)*], then the number density ‘of photons per unit volume participating in the absorption is M’= nd? p/(2xh) (naj /c>Aw/2x). The corresponding flux is \Vc, and the rate of absorp- tion is do re Nowe = aya = On = 03 Sn, 1.34) we = Iw ag — OR = OO as ving ops(co/2)= nrc. This result can be stated more formally as [Positeannn teas) (Se) OP o mee 0 mee which suggests expressing the cross section as 2 ce ono) = = 6,, S (136) mec where ¢ is called the fine-profile function; the integral of Q (dw /2x) is unity, When the linewidths are ignored, ,, is proportit thé Dirac delta function; when the finite width of the energy levels into account, ¢ will become a function that is peaked at oy with a’ width Aw, such that (0) =(Ae)”'(2z1). The absorption cross sé@tjontat the centre of the line is given by o ~2x(rc/ Aw) ~ (23/2), is the wavelength of the photon. A more rigorous quantum-mechanie will provide the explicit form for #(w) (which turns out to be a Lo in this case) and an overall ‘multiplicative factor f called the oscillat ath, 1.4.3 Therm The result regarding the emission and the absorption rates also allows us to determine the the energy distribution of photons in equilibrium with matter at temperature 7. In such a situation, photons will be continuously absorbed and ‘adiation 1.4 Radiative Processes in Quantum Theory "7 emitted by matter. Consider the rate of absorption (or emission) of photons between any two levels, say, a ground state |G) and an excited state |). We saw above that the absorption rate per atom is given by Qn and the emission rate is Q(n + 1), where n is the number of photons present and Q is determined from the quantum theory of radiation. In steady state, the number of upward and downward transitions must match, which requires that product (number of atoms in G) x (rate of upward transitions per atom) equal (number of atoms in E) x (rate of downward transitions per atom), that is, N¢ Qn = Ne Q(n + 1). Because matter is in thermal equilibrium at temperature T, we must also have (Ne/Nc)= exp(-AE/ksT), where AE is the energy difference between the ‘wo levels. This relation should hold forall forms of matter with arbitrary energy levels; hence we can take the ground-state energy to be zero (ie, arbitrarily small) and the excited-state energy to be E, leading to 1 1 (No/Ne)— 1 exp(E/kaT) — (1.37) This equation gives the number of photons with energy E [or ~equivalently —with momentum p =(E/c)] in thermal equilibrium with matter. To be more precise, the number of photons with momentum in the interval [p, p + d°p] is given by, IN =2n{Vd* p/(2h*)}, where the factor in the square brackets gives the m ber of quantum states in the phase volume Vd? p and the factor 2 takes into} count the two spin states for each photon. The corresponding energy dE. through d3x = dA(cdt) will be dE = hv dN = 2n(v)hv dAlcdt\{a° p/h? ing d p = p*dp dQ = (h/c)’v? dv dQ, we can determine the intensity{ whigh is, the energy per unit area per unit time per solid angle per frequené : smal radiation as dE 2m? yy 2 dAdtdQdv an eT The quantity vB, reaches a maximum value around hv ae ‘hich translates (1.38) to the fact that a blackbody at 6000 K will have the maxibkumfor vB, at 6000 A. The maximum intensity is (v By)max (7/100 K)* 1 sr. Such thermal radiation can arise in many different contexts in whi imary source of energy is thermalised because of some physical proceg=theymost important example being stellar radiation ‘At low frequencies, the intensity of th will be By ~(2kgT/X*). Because of this a brightness temperature for any sour which is (in general) a function of frequency” It is clear from Eq, (1.37) that there are very few photons with momentum ‘greater than p~ (Kg /c), so that (V/V)*~ (4z/3)(p/2nh) ~ (kaT /hc)? and ‘the mean energy is Urr © kpT(N/V)*(kgT)4/(he)>, lation given by Eq. (1.38) is conventional to define 221, /2kp), 18, 1 Order-of- Magnitude Astrophysies 1.4.4 Photon Opacities in Matter Let us next determine the conditions under which our original assumption — that the radiation is in thermal equilibrium with matter — holds. If the cross section, for the relevant process that scatters or absorbs radiation is given by o and the number density of scatterers is n, then the mean free path of a photon is given, by l= (no). In the case of radiation, it is conventional to define a quantity x (called opacity) such that (1.39) where p is the mass density of the scatterers. The optical depth of a system of size R is defined to be t =cR=(R/1). From the standard theory of random walk, we know that a photon will traverse a distance R with N. collisions, where R = N21; that is, the number of collisions is given by Ne=(R/I)? = r?, provided that >> 1. The opacity for the photons is provided mainly by three different processes: (A) scattering by free electrons, (2) the free-free absorption of photons, and through it. (3) the bound-free transitions induced in matter by the photons that are “< (1) The simplest case is the one in which the charged particle is accelerate by an electromagnetic wave that is incident upon it. Consider a chad 9) placed on an electromagnetic wave of amplitude E. The wave will i acceleration a~(qE /m), causing the charge to radiate. The ie : will be P =(2q2a*/3c?) =(2q*/3m?c?)E*. Because the ing pager in the electromagnetic wave is $= (cE?/4m), the scattering es On (for electrons with m =m.) is ay? oat #(5) wor GQ) (1.40) which is the Thomson scattering cross section. TER 's section governs the basic scattering phenomena between charged‘pasticles and radiation. The corresponding mean free path for photgf\through a plasma is Jy = (ne or)-!, and the Thomson scattering opaci ied to bexr =(neor/p), : wr=(t)(2) goo! a for ionised hydrogen with ne == The opacity for the process in (2) and (3) can be determined by the principle of detailed balance which allows us to relate the rate for certain processes to the rate for the corresponding “inverse” process. (2) The time-reversed process corresponding to bremsstrahlung is the one in which a photon is absorbed by an electron while in the proximity of an ion, 1.4 Radiative Processes in Quantum Theory 19 In equilibrium, the rate for this free-free absorption should match that of thermal bremsstrahlung. We write the free-free absorption rate as nop B(.), where ogis the free-free absorption cross section and B(v) ox v3(e"/ks! — 1)! oc v7 (when hv < kT) is the intensity of the thermal radiation. Equat- ing this to the bremsstrahlung emissivity found in Eq, (1.17), jy ocn?7-", we get oy o(n/v*T/?), Taking the typical frequency of the photon to be proportional to 7, we find that « 7 ocnT~>5. The corresponding opacity can bbe written in the form ky x pT. (1.42) (3) To obtain the bound-free opacity, which arises when a photon ionizes an atom, we begin by relating the photoionization rate to the recombination rate in which an ion and an electron get bound together, releasing the excess energy as radiation. The recombination rate per unit volume of a plasma will be proportional to (1) the number density of electrons ne, (2) the number density of ions n;, (3) the relative velocity of encounter v, and (4) the cross section for the process, which will be ~22, where A gives the effective range of interaction between electron and proton. For Coulomb interaction, this range is 21 = (2Zq?/mev?); on the other hand, an electron of speed v has a de Broglie wavelength of 22 = (h/m,v). We should choose 2 to be t larger of Ay or 22, depending on the context. Let us first consider 5) with 2.~ Az, which corresponds to (v/c) 2 (q”/he). Each recombination release an energy of approximately (1/2)mev*. Hence, S (5) (82nd) xen where we have assumed that mv? “kpT, nex p, and 1 rium, the photoionisation rate (which removes energy field) should match the recombination rate. The amour by photoionisation is proportional to don oc igs ie opp is the photoionisation cross section. Equating dog to dj ‘using Eng oT, we get NaomOnyT* 0 nen Introducing the bound-free opacity kyy by’ and taking n, x p and n, oc p, we find th Koy which scales just like relation (1. In a radiation bath with temperature T, the typical energy of photons is hy ~kpT. The result of relation (1.45) suggests that the frequency dependence of o4p(v) oc kog XT will be of the form o(v) =onr(v/vy)* for v > vy (and zero otherwise). Here s © 3-3.5 and vy = (€q/h) is the frequency corresponding 20 1 Order-of- Magnitude Astrophysics to the ionisation energy ¢g of the atom. Because the cross section for photoi sation o(v) satisfies constraint (1.35), [Pewrar=anet = (Yi (1.46) f mec. where the term within the parentheses comes from classical theory and the oscillator strength f is supplied by the quantum theory, we get ong = (8 = I) frody ~ 2frodt, (147) where 2 =(c/v1) and s ~3. This shows that the photoionisation cross section is essentially the product of the classical electron radius and the wavelength at ionisation threshold. Using 4; =912 A, we get op * 107" cm? for hydrogen. For heavier elements ap will scale as sy (2) « (4) «Ztv (148) » v if we take s =3. The cross section for recombination, ojec, is related to op by srg Bossa! py where 4=2 x2 is due to electron and proton spins, 8=2 x 2 is due to electron, proton, and photon spins, and p, and py refer to th menta of the electron and the photon, respectively. Because dp, ~ gives Gree = 200/(py/ Pe)? if the plasma is at some temperature T, tn) 2hw)? & &a -( kaE XO Sy fee = ony = 2( 25) (Vong ~ 10 ons = meet aa) iar) 10 7 (1.49) atthe threshold with fo ~ €, ~ 10eV. The rate of recombi rrunit volume per second is n?a,-cv =n7ap, where og = 2 x 107 (ky 12 em} 5-1. In the analysis so far we have assumed that the regoribibation proceeds di- rectly to the ground state. Such a process, however, Will#lease a photon with energy he & €q that will immediately ionise anotker atom. The net recombina- 6 jing an excited state that hu <€q and cannot ionise the neutral atoms can write a relation similar to relation (J@33b Then the energy loss that is due to recarbination is dE pe ae 7 =) oc npn (<2) « (Jna?) [2 cena oe, This can be one source of cooling for a plasma at T 2 10* K in addition to the bremsstrahlung cooling (4Eiye/dV dt) on?" obtained in Eq, (1.18). The net 15 Varieties of Astrophysical Structures 21 rate of cooling for a plasma can be expressed in the form (dE /dVdt)=n? A(T), where A(T)=aT"/? + bT~"?; the first term which arises due to bremsstrahlung dominates for T= 10° K. ‘When both photoionisation and recombinations occur in a region, the equi- librium is described by the relation nemiare ~ onyx F, where F is the flux of ionising photons with v > v). Taking ne=n;=.xno, ny =(1 —x)mo, we can write this equation in the form aaa? (2) a) =* (eR) Ga) 0 which determines the ionisation fraction x in many astrophysical contexts. If'a source of photons emitting ', ionising photons per second (with v > vy) ionises a region of volume V around it, then the same argument gives n2apV ~ N, Taking ne =.xno =o, we get V =(N, forand). It is also possible to imagine a situation in which the mean free path for collisions between atoms is small (50 that the matter is in thermal equilibrium) ‘but the mean free path for collisions between photons and matter is lange (so that radiation is not in equilibrium with matter). In that case, the photons emitted by the system will escape from it with negligible scattering and there could aris radiation from atomic or molecular transitions, characteristic of compositio ‘Then the amount of energy emitted per second by unit volume of gas, per frequency range, and solid angle can be written as, = (—#F__) _ [8 exp (=u) ] (ew += (qeema) =[2e0( KT, JC) anne 1) Z (the partition function) providing the normalisation. Becay are known in this expression, it can be used to relate the souy observed intensity. An important example of this is the, neutral hydrogen atoms. The intensity of this line an various effects, make it a useful probe of several syste! 15 Varieties of Astrophysical features of astronomical ples given above. It is clear, \s that are massive enough so ‘We now turn to the question of trying to dete systems from the description of the physi right at the outset, that we are interested i that gravity plays a significant role in nics. In the most extreme limit, we can think of the entire universe as &plysical system and try to determine its structure. At sufficiently large scales, we can ignore the graininess in the distribution of matter and think of the universe as reasonably homogeneous and isotropic. Further, because no location in such a universe can be considered as 2 1 Order-of Magnitude Astrophysics special, any possible motion of matter on a large scale should also maintain the same characteristics with respect to any observer. It immediately follows that the most general motion consistent with these requirements must have the form. v(t) = f(r, where r and v denote the position and the velocity, respectively, of any material body in the universe and (¢) is an arbitrary function of time. This is, the only kind of motion that is consistent with the requirement that from any point in the universe an observer will see matter moving in an identical manner. Using the fact that v= f, this equation can be integrated to give r=a(r)x, where a(t) is another arbitrary function related to f(t) by f(t) =(4/a) and x is a constant for any given material body in the universe. It is conventional to call x and r the comoving and proper coordinates of the body and a(t) the expansion factor (even though, if a <0, it acts as a contraction factor), ‘The dynamics of the universe is entirely determined by the funetion a(t). The simplest choice will be a(t)= constant, in which case there will be no motion in the universe and all matter will be distributed uniformly in a static configuration It is, however, clear that such a configuration will be violently unstable when the mutual gravitational forces of the bodies are taken into account, Any such instability will eventually lead to random motion of particles in localised regions, thereby destroying the initial homogeneity. Observations, however, indicate aS this isnot true and that the relation v = (@/a)r does hold in the observed univer In that case, the dynamics of a(t) can be qualitatively understood alor following lines. Consider a particle of unit mass at the location r, wit to some coordinate system. Equating the sum of its kinetic energy {&/2 find gravitational potential energy that is due to the attraction of métier within a sphere of radius r to a constant, we find that a(t) should satisfy fion Via ArGolt) 2 _ 3 5 constant, (1.52) where pis the mean density of the universe, that is, +550 .es from the proper appli- cation of Einstein's theory of relativity to a hofhiggeiieous and isotropic distribu- tion of matter. Observations suggest that our’ se today (at f= 1p) is governed by this equation with p(t) * 10-2" gm ef nd (a/a)o 03x 10-7 s~!, where h*0,5-1. This is equivale eed Ho = 100h km s~! Mpe~! where 1 Mpc ~3 x 10? cm is a convenientumie{or cosmological distances. (We will also use the units 1 kpe = 10-3 Mpc and 1 pe = 10° Mpe in our discussion.) From Hp wecan form the time scale funiy = Hy | 10'h-" yrand the length scale cH, | ~3000h—! Mpc; tyniv characterises the evolutionary time scale of the uni- verse and H,,' gives the largest length scales currently accessible in the universe. 1.5. Varieties of Astrophysical Structures 23 The light emitted at an earlier epoch by an object will reach us today with the wavelengths stretched because of the expansion. If the light was emitted at a=a, and received today (when a =o), the wavelength will change by the factor (1 +z-)= (ao/de), where ze is called the redshift of the emitting object. ‘The observed luminosity L of a source will decrease as (1 +2)-*, as L is pro- portional to (pyc) d° py x vdv (1 -+2)~4, where py =(€/c)=(hv/c) is the photon momentum If neither particles nor photons are created or destroyed during the expansion, then the number density of particles or photons will decrease as a~* as a in- creases. In the case of photons, the wavelength will also get stretched during expansion with 4 «a; because the energy density of material particles is nmc? whereas that of photons of frequency v will be nhv = (whc/2), it follows that the energy densities of matter and radiation vary as Prag a~* and Pmatter Xa~?, ‘Combining this result Prag a7 with the result Prag & T* for thermal radiation, it follows that any thermal spectrum of photons in the universe will have its temperature varying as T (157) ‘The condition for efficient cooling fase < pled with kT ~GMmy/R, leads tothe constraint R < Ry, where R, = ag'r.( 74 kpe, (1.58) (Gm2 nc) ~ 6 x 10 is the gravitational (equivalent) ofthe fine- structure constant. In the above analysis we assume that kyT > a°m,c?; for 1.5 Varieties of Astrophysical Structures 28 R~ Ry this constraint is equivalent to the condition M > M,, where ‘m,\'2 M, ~ acia'm,( 2) =3x 104 gm. (1.59) This result suggests that systems having a mass of ~3 x 10 gm anda radius of ~T0 kpe could rapidly cool, fragment, and form gravitationally bound structures. Most galaxies have masses around this region. This is one possible scenario for forming galaxies. Note that the mass and the length scales in relations (1.59) and (1.58) arise entirely from the fundamental physical constants. We now consider some more properties of these structures. The original (maximum) radius of the cooling plasma estimated above is ~10 kpc. After the matter has cooled and contracted, the final radius is more like 10-20 kpe, which is the typical radii of large galaxies. For M, ~3 x 10 gm and R, ~ 20 kpe, the density is pga ~ 10-75 gm cm™3, which is ~10° times larger than the current mean density, py = 10>" gm cm, ofthe universe. Ifwe assume that high-density regions with )% 100jypiv collapsed to form the galaxies, then the galaxy formation must have taken place when the density of the universe was ~1000 times larger; the value of a(t) would have been 10 times smaller and the redshift of the galaxy formation should have been zat <9. If the protogalac plasma condensations were almost touching each other at the time of fo tion, these centres (which would have been at a separation of ~150 kpc) have now moved apart to a distance of 150(1 + zyal) kpc ~ 1500 kpe= | This is indeed the mean separation between the large galaxies to est galaxy with a radius of ~10 kpc, at a distance of 1 Mpc, an angle of 8a 10"? rad~30'. A galaxy at a distance of ~: subtend 0.5". Observations have indicated that many different kinds of férastufes exist at redshifts of 2 $5. Because the process of gravitational ins which leads to the condensation of galaxy like objects, cannot be 1 nt, it would leave some amount of matter uniformly distributed ir ithe galaxies, The light from distant galaxies will have to pass through thk{wuahter and will contain signature of the state of such an intergalactic mediugy (IGM). The photons (with > v1) produced in the first-generation objects caldg\dqusce a significant amount of ionization of the IGM, especially the low: photons (with v > vz) impinge on a gas of ney ‘ny, it will have an ionisation optical deptl a critical column density for ionisation 1471 = 10"7 cm, Regions with a hydrogen column dens sig 2:nR greater than 10'7 cm? will appear as patches of neutral regions in tft ionized plasma of the IGM. Such regions can be studied by absorption of light from more distant sources, (espe- cially through Lyman alpha absorption corresponding to the transition between n=1 and n=2 levels) and are called Lyman alpha clouds. 100g R. Setting t = 1 gives 26 1 Order-of- Magnitude Astrophysics Still larger structures than galaxies, called galaxy clusters, with masses of ~10" gm, a radius of ~3 Mpe, and a mean density of 10-77 gm em=?, exist ir the universe as gravitationally bound systems. Our argument given above shows that the gas in these structures could not have yet cooled and will have a virial temperature of T= (GMm p/Rkp)~4 x 107 K. Let us next investigate the nature of smaller-scale structures that can form inside a galaxy. Here, the existence of the different energy scales and equations of state allow for the possible existence of several - widely different — astrophysical systems, and most of these systems can be understood by systematic comparison of the energy scales (€, €a, Ent ***) with 1.5.2 €gray © €q: Existence of Giant Planets The laboratory systems have negligible gravitational potential energy; in a plot of (cr/kaT) against (GMm_/RkgT) they exist (almost) along the y axis [see Fig. 1.1 (top)]. We now see how gravity affects the structures as they get big- ger. The atomic binding energy (per particle) of a system is approximately qa mec? © (g? /ag) © qn"! if the atoms are closely packed (with naj ~ 1), and the gravitational energy per particle is €y = (47/3)'/*Gm},N7n'/9, The ratio is given by f (#)\(1 Ne\?? _ (10#)?8 S fx(*)(.)=(*2) (2). 0 é*()(wa)=(%) ~Gr) Clearly, the number Ng =a*/?a5*? ~ 10%, arising out of fu tants, sets the smallest scale in astrophysics, in which the gravi energy becomes as important as the electromagnetic binding-enyey of matter. The corresponding mass and length scales are Myina ~ Rouwer ~ Neg ~ 10" cm and correspond to those of a Jakes ‘masses, gravitational interaction changes the structure sii antly whereas for smaller masses gravity is ignorable and matter is roKcous with constant density so that M oc R?. Because we used Newtonian gravity to arrive a to verify thatthe parameter Ry for M~ Mplage R~ Ryranet 8 Rem = a7(mg/r Of Rym follows purely from the values of Most of the astrophysically interestin Rea el cdiiclusion, itis necessary sal for this scale; this ratio "1, Note thatthe smallness mental constants shave larger mass and require ‘oyees other than normal solid-state forces. In general, such systems can 6 €hissified into two categories. The frst set has the gravitational force balanced by the kinetic energy of classical mo- tion, whereas the second one has the gravitational free balanced by degeneracy pressure. For a system with naj ~ 1, the non-relativistic Fermi energy of elec- trons is comparable with the atomic binding energy and we can compare €, or €e with €., This is meaningful as long as the temperature of the system is low 1.5 Varieties of Astrophysical Structures 27 ie Teaco S naw ite de tGeoen] — = Pane S == S =, *) [eescaveam S a ems = Boh S a A om Fig. 1.1. Aspects governing the dynamic jous cosmic structures are summarised. See text fora detailed description. 28 1 Order-of Magnitude Astrophysics and kpT ep ~€q~ 10 eV. Ifthe temperature is significantly higher, then itis, the thermal energy kT that should be compared with ¢,; that is, the gravita- tional pressure will be balanced by the Fermi pressure at low temperatures and by thermal pressure at high temperatures. We now examine such structures. 1.5.3 €grav~ €nat: Existence of Stars ‘When the mass of the system is increased further, the gravitational pressure in- creases and — to balance it both the Fermi pressure and the thermal pressure will increase. The dynamics of the system will then depend on the relative significance of these two quantities. To take into account both thermal and quantum degener- acy contributions, we take the matter pressure to be P “nkgT +nep, which is a simple interpolation between the two limits. This pressure can balance the gravitational pressure if (kgT +r) Gm?,.N?n'/3, Using expression (1.11) for er for the non-relativistic electrons, we get an2)2/3 2 Gait A ka Gm2n2%q12 — me (1.61) For a classical system, the first term on the right-hand side dominates, and we s that the gravitational potential energy and kinetic energy corresponding to, temperature T are comparable; this is merely a restatement of the virial th As the radius R of the system is reduced, the second term on the right-h (cen?) grows faster than the first (acn'/9) and the temperature of thé will increase, reach a maximum, and decrease again; equilibrium [g-pos any of these values with gravity balanced by thermal and degetist ay The maximum temperature Tmax is reached when n =n, with ays 2 nsx —7o_(N kn To ~ te 2 (1.62 = inl Re ): Toes ~ 532M ». (1.62) where 2. = (h/mec) and ag = Gm? jhe. ‘An interesting phenomenon arises if the maximum {eenp)rature Tras is sufi- ciently high to trigger nuclear fusion in the systemz ally bound, self-sustained nuclear reactor. The reaction has to come from detailed study of t ically at energy scales higher than éyut ~ corresponding to the maximum temperatus en,ave obtain a gravitation- jon for triggering nuclear ie nucleus and occurs typ- , with n~0.1. The energy ie Will be larger than éyyci When @ \3? N= anten?'2( Jo x40 sy ag for n= 0.1. The corresponding condition on mass is M > M,, where 3g \32 = any'4x?)!2(‘"2) () mp ~4x 10 gm, (1.64) re) \aa 1.5. Varieties of Astrophysical Structures 29 which is comparable with the mass of the smallest stars observed in our universe. The mass of the Sun, for example, is Mo =2 x 10°? gm. Comparison of relations (1.59) and (1.64) shows that the number of stars N, =(M,/M,) in atypical galaxy will be given by the combination of fundamen- tal constants N, =a7?a5/?(m,/m,)!/4~ 10". Typical galaxies indeed have approximately 10''—10"? stars, although there is a fair amount of spread in this number. The mean distance between stars in a galaxy will be dar ~ (Reat/ Ne!) I pc. A starlike the Sun, with a radius of 10! em, located at a distance of 10 pe, will subtend an angle of about ~1 milliaresecond; itis clear that most stars will look like point objects. 1.5.4 Existence of H-R Diagram for Stars Once the nuclear reactions occur at the hot central region of the gas cloud, its structure changes significantly. Ifthe transport ofthis energy to the outer regions is through photon diffusion, then the opacity of matter will play a vital role in determining the stellar structure. In particular, the opacities determine the relation between the luminosity and the mass of the star. ‘Aphoton with mean free path = (no)~!, random walking through the plasm: will have Noon = (R/1)? collisions in traversing the radius R. This will take t 'Neoi/¢) ~(R/e)(R/I) for the photon to escape. The luminos} a star L is the ratio between the radiant energy content of the star Ey Because Ey ~(aT*)R?, where a = (17k% /15h°c?) = (mr? /15)~ | (in ut ith ks ), we find that L=(aR*T41/R?)~ RT‘1. For a of stars, we may assume that the central temperature T ~ (GM p/ ably constant because nuclear reactions — which depend strongly Hot as a thermostat, If Thomson scattering dominates, then ¢ = or, a Rr* rapt 3 Lx (1.65) orn a7N If, on the other hand, the plasma is only partially iokised)) we should use the opacities in expression (1.45), with I oc T7/?n~? oc R777 RM, and we have Lo RT* & ROTM? Taking R 0c M gives L cx M5. Itis conveni T, of the star by the relation L cx R°T, ‘Combining this result with the relation tially ionised, we get T, x L8L-'" On the other hand, if Thomson scattering dominates with L x M3, we get T; oc L'/'?, When the stars are plotted ina log T,-log L plane, (called the HR diagram) we expect them to lie within the lines with slopes (3/20)=0.15 and (1/12) ~ 0.08. The observed slope is ~0.13. lefine the surface temperature Foc LUAR-V? og LAMY, 175, when the interior is only par- 30 1 Order-of Magnitude Astrophysics The temperature inside a star varies from approximately 10° K at the core to approximately a few thousand degrees Kelvin at the surface. The physical conditions, equation of state, and the opacity of matter vary significantly inside a star even when it is in a steady state. Further, when the stars evolve because of different nuclear processes, the response of matter to changing physical con- ditions can be dramatically different. We shall encounter many of these features in Volume II of this series. The condensation of stars from galactic matter cannot also be a totally effi cient phenomena, and we do expect a fair amount of matter to be distributed in the galaxy in different forms. This constitutes the interstellar medium (ISM) in which structures of very different densities and temperatures exist in pres- sure equilibrium. In our galaxy, the ISM contributes a mass of ~10°Mo and has a pressure of approximately P =nkgT ~ 10-2 dyn em~®, There exist a hot diffuse component (T ~ 10° K, n= 10-3 em~), a warm ionized compo- nent (7 8000 K, n ~ 10! em™3), a warm neutral component (7 ~ 5000 K, n~ 107! cm™?), a cold neutral component (7 ~ 80 K, n = 10-100 cm~3), and giant molecular clouds (T ~ 10 K, n = 10?-108 cm~) in pressure equilibrium in the ISM. There are also processes that strongly couple the stars and the ne ‘Consider, for example, the region around a hot star with L = 3.5 x 10° erg s and T, ~3 x 10* K. The number of ionizing photons Ny (with v > v;) emit by such a star can be estimated from the Planck spectrum and will be a mately 3 x 10% s~!. In Subsection 1.4.4 we saw [see the discussion fo relation (1.50)] that the volume of the region ionised by such a flux is Ny fan) Using ny 10 em->, we find that matter wll be fulljonise for a region of radius R =(3V/4x)! he star, relation (1.50) gives (1 —x)~ 10-4, indicating nearly total ionffatiyp. Such a 1.5.5. gry =p: Existence of Stellar Keds sh which self-sustaining ich systems, the kinetic NkgT ©(GM?/R). The uuclei releases ~0.03m pc? of ‘4m yc?. Assuming that a frac- We saw above that stars are gravitationally bound s nuclear reactions are taking place in the centy energy and the potential energy are compara process of combining four protons into a energy, which is ~0.7% of the original tion €~0.01 of the rest-mass energy ‘made available for nuclear reac- tions, we find that the lifetime of the burning phase of the star will be tor = €M/L *3 x 10° yr (€/0.01)(M/M.)~ if the opacity is due to Thomson scattering. This defines the characteristic time scale in stellar evolution. When the nuclear fuel in the star is exhausted, the gravitational force will start contracting the matter again and the density will increase. Eventually, the LS Varieties of Astrophysical Structures 31 density will be sufficiently high so that the quantum degeneracy pressure will dominate over thermal pressure. The equilibrium condition for such a system will require the degeneracy pressure of matter to be large enough to balance gravitational pressure. Equivalently, the Fermi energy ¢r(n) must be larger than the gravitational potential energy €, = Gm2N2/n'/, When the particles are nonrelativistic, €-(n) = (h?/2m,)(3x?)°7n2/? and the condition €r > €, can be satisfied (at equality) if 2 Gmim. va ome) 2/3 nll - am ra we (1.66) With n =(3NV/4x R3) and N =(M/mp), this reduces to the following mass— radius relation: RM? ~ aG!hem\/3 ~ 8.7 x 1079 RoMg (1.67) Such structures are called white dwarfs. A white dwarf with M~ Mz will have R~10-*Ro and density p ~ 10°pq. As the density increases, electrons combine with protons through inverse beta decay to form neutrons, which can provide the degeneracy pressure. Equa- tion (1.66) is still applicable with me replaced with my; correspondingly t right-hand side of relation (1.67) is reduced by (Xp /%¢) = (me/m,) ~ 108 objects — called neutron stars — will have a radius of R ~ 10° Ro and adens| p=10!5po if M ~ Mo. For such values Rpm ~ | and general relativisti¢et are beginning to be important. ‘When the density is stil higher, the Fermi energy hasto be supp! tic particles, and ¢¢ now becomes € ~hen'/3, which scales as, 3, just like €y. Therefore the condition ¢¢ > €, can be satisfied only if fc! N23 or N kT) and the lower me dominated by thermal effects (er (which exist in the atmosphere) fall within the IR bands; this ¢ the IR radiation also to be absorbed by the atmosphere, although to si lesser degree than the higher energy radiation. Because of these effests:the ‘ground-based observations are essentially limited to visible (2 v= 10!85 x 10! Hz) and radio (> em, v <3 x 10"° Hz) There is, however, another limitation arising from the fact wavelength radiation ( 2 100 m) cannot propagate throu; ionosphere and is reflected back. Consider an slecvomaant e that moves the electrons in a plasma (relative to ions) by a small: 83x along the x axis. This deposits a charge @~e(nAdx) on a surface of area A perpendicular to the x axis. This charge density\in_tim, will lead to an electric field Ey ~47(Q/A)~4zendx that acts dn the,electrons in this small volume, pulling them back. Such a restoring sroportional to displace- ment, gives electrons a characteristic frequ: oscillation (called plasma frequency) 4nen\ o=( we ) Waves with frequencies lower than the plasma frequency cannot propagate through a plasma as the electrons can redistribute themselves sufficiently quickly to cancel the field of such an electromagnetic wave. We can estimate the num- ber density 1 of electrons in the ionosphere by equating the ionisation rate that (1.68) 1.6 Detecting the Photons 35 is due to solar radiation with the recombination rate, as done above for HIT regions. This gives an electron density of approximately 4 x 10° cm™>; the cor- responding plasma frequency is approximately vp = («p/2s) = 6 MHz. Thus we cannot observe radio waves with frequencies lower than ~6 MHz, correspond- ing to wavelengths larger than about ~S0 m. To obtain information about all other wavelength regimes, it is necessary to make observations at high altitudes: from balloons, aircrafts, spacecrafts, satellites, ete. We now consider each band separately, 1.6.2 Radio (A=3cm-l0m, v~3x107-10'°Hz, T= 10-3-0.5K). Several discrete sources (supernova remnants, radio galaxies, quasars, etc.) emit radio waves essentially because of synchrotron process. As a specific ‘example, consider the diffuse radio background in our galaxy that is due to synchrotron radiation from electrons in the ISM. Observations indicate that the spectrum of the electrons is given by (dN /dE)~3 x 10-"' E-*° particles em~ if E is measured in giga-electron-volts. If the magnetic field is approximately 6 x 10° G, the relation (1.22) will predict a volume emissivity of an jv=3x w( ) ergs“! em Hz! 10 MHz, This is close to the observed background emission. Over a line of s this will give a flux of 10-?? W m=? rad~? Hz~!. Similar power-I are seen in the case of radio galaxies, quasars, etc., and are tho. synchrotron radiation. ‘Along the spiral arms of galaxies, there exist clouds of HI thermal bremsstrahlung radiation with a relatively lat spectry ‘Hil regions in Orion have T ~ 10* K,n; ~2 x 10° em~ cifeetive line-of- sight thickness of 6 x 10~ pe. From Eq, (1.17) we can the flux to be ap- proximately 3 x 10-2! ergm™? s~! Hz~! rad~? = 300! }, where | Jansky = 1 Jy= 10-26 W m™ sr“! Hz. This is typical ofemisgion from HII regions. ‘The total diffuse radio emission from the galaxy Radio galaxies are another main class of ra cated pattern of emission. The radiation usu side of the central galaxy with a separati power from these radio sources is quite needed to explain this emission. The f due to such a source ata distance ‘of 3000 Mpc will be approximately 10™' m™? sro! Radio observations can also detect the presence of neutral hydrogen in the universe through the 21-cm radiation discussed in Subsections 1.4.1 and 1.4.2. Because hv KkgT in most cases involving 21-cm radiation, g/Z in Eq. (1.51) that emit Example, the 36 1 Order-of- Magnitude Astrophysics is essentially the ratio of spin states gi /(gi + go), Which is 3/4 for the hydrogen hyperfine transition. The total intensity from an optically thin column of length L will be (for hv (Gea) aa)"(2) = (en) 0 This force will exceed the gravitational force attracting the proton, fy = (GM p/ P),if L > Le, where 4xGmyec “( _M 7 Le cr M ~ 1.3 x 10" io"me) eS (1.73) is called the Eddington luminosity. The temperature of a system of size R ating at Le will be determined by (4 R)oT4 = Lr, that is, raraxwt(Z)"(LY Go Mo. Tkm, Ge > faa = (nr) & For a solar-mass compact (R~ 1 km) star this radiation will pe ray band. ‘The main extragalactic sources of x-rays are quasars Tnised gas in clusters of galaxies. Taking the cluster gas to be full hydrogen with a mass of approximately Myas= 10" gm spread. here of ra- dius R~3 Mp, we can estimate the number density, and electrons as ne * (Myas/mpV)* 10-4 cm™, where V is the volunig.o})the cluster. Such a {gas will be a source of thermal bremsstrahlung radia 10-27 n27'? erg s~! cm”? [see Eq. (1.18)] where aay Using the estimated values of n, and T-, we get tha cluster tobe £ = LV ~5 x 10 erg s!. Thi corresponding to the temperature of 108 4 x-ray-emitting clusters have been obser Inaddition, there exists a well-defin fe x-ray background in the range of 1 keV to 100 MeV. Part of this background €ould be due to unresolved pointlike sources and another part may be due to hot (T ~ 10° K) diffuse, intergalactic plasma. In the range 3-50 keV it can be fitted by an optically thin thermal bremsstrahlung at the temperature 40 + 5 keV. tign afte rate of L = 1.42 x hquantities are in cgs units. 40 1 Order-of- Magnitude Astrophysics ‘No object in the universe is hot enough to produce high-energy y rays by ‘thermal radiation. The y rays are produced by accretion of matter on compact ‘objects and by the collision of high-energy particles inthe cosmic rays with the nuclei of atoms in our galaxy. Figure 1.2 summarizes the electromagnetic spectra of the universe. In the top part of the figure the y axis is the flux per logarithmic band, F = vF, in units ‘of watts per square meter. The corresponding bolometric magnitude m, related to flux by log F = is « vf we leh in oa __” tiv stey sav | a7 Tete wapence | e ee ase Ab “baw eee onder | 0 BON Serene | © ah “f ot cantsrata, 1 ee 1GHe 10H soit 1 ze ado 28) gv e & - Pl 0p, Ho | # [ore Gas ] SSRaSmRA{ Srctwen ae Fig. 1.2. This figure summarises the v ‘The top part of the figure gives the spectra of a class of objects and the effective sensitivities reached in the surveys in different wave bands. The bottom part gives (1) the atmospheric absorption, (2) the key processes contributing to line radiation, (3) major thermal sources, and (4) main sources of nonthermal emission in different spectral bands. Fadiative phenomena in astrophysics, 1.6 Detecting the Photons 41 main x axis is marked in both frequency and wavelength at the bottom and in equivalent temperature, obtained by k»T = hv, along the top. The bottom part of the figure gives several general features discussed above. The first panel below the x axis divides the frequency range into different wave bands and indicates the major components of the atmosphere that absorb the radiation in each band, ‘The next panels indicates the processes which can lead to line emission in each of the wave bands. Finally, the main sources of thermal and nonthermal radiation in different wavebands are summarized in the last two panels. In the main figure, the typical spectra from different sources as well as the detection limit of different instruments in various bands are shown. A bright star, ‘lose tom = | and peaked in optical, has the maximum flux. Other sources are (1) atypical supernova remnant marked SNR (Crab), (2) a radio galaxy Cygnus-A, (3)a quasar (3C273) with emission in several wave bands, (4) an 18th-magnitude elliptical galaxy, and (5) a bright-x-ray source Sco X-1. In addition to these sources, also marked are (1) the flux from the cosmic microwave background radiation from one square arc second of the sky, (2) x-ray background per square degree that is due to unresolved sources, and (3) the sky background per square are second in the optical band. The filled rectangular boxes are the detection limits of different probes operating in various bands. Many of these, like mee HEAO, ROSAT-Einstein, GRO-EGRET, COS-B, and of course, the Hubble space, telescope, are satellite-based instruments. Some of the points marked in optical band like the Hubble deep field (HDF), come from integrated-pen beam kind of surveys, whereas many other limits are applicable to widey surveys. This figure summarizes our current technological “ee I contexts, as the expected flux in different astrophys Exercise 1.1 eo Why the rest ofthe Book? Arguments similar othe ones given in O-- appeal: ing as they seem to capture the essence of the physics behind eacih oF phenomena. Investigate carefully whether this is indeed true. In particular, AGB the scaling relations obtained by such arguments trustworthy? (2) Are iy estimates trust- worthy ~ especially because, for example, (2x) = 10°? yee se arguments have predictive power? (Answers: (1) The sealing relations are egually Fustworthy 2) No. In several places, the numerical estimates ean be wrong an an order of magni tude. (3) No. None ofthese effects were ever derived rguments. However, once the correct result is known from painstaking, rigo ematical analysis ~ which ‘we shall encounter in rest of the book — such arg ‘be provided to “explain” the “physical origin”, The tue value ofthese arguméh#89 in acting as useful mnemonics.] 2.1 Introduction This chapter develops several basic ideas of dynamics, emphasizing general principles that are useful in classifying the behaviour of dynamical systems. The reader is assumed to be familiar with elementary concepts of classical mechan- ies. Concepts developed here will be needed in the study of special relativity (Chap. 3), statistical mechanics (Chap. 5), general relativity (Chap. 11), Sun and solar system (Vol. II), binary stars (Vol. II), and galactic dynamics (Vol. III). © 2.2 Time Evolution of Dynamical Systems & of tereal Many systems encountered in nature can be described by a finite get variables [q1(t), a(t). «--. gi(t)s-»-.gn(#)] that evolve in time Ceo the study of two stars, moving under the influence of their mut force, we are interested in the positions of the stars as functio position of each star can be described by three coordinates fimensional space) so that the full system can be described by a se functions of time, The quantities qi(*) (with i = 1, 2, ..., N) are call obviously, we are free to choose any other set of N it it, single-valued functions of q: as dynamical variables to describe the syStes), with the particular choice often dictated by mathematical convenience\The,central problem of dy- namics is related to determining the time deper f qi(t) and studying the general characteristics of motion It turns out that the evolution of any pl terms of differential equations that are ( suitably chosen variables. This implies for all 1 > fo if we know the 2” quantit®S 7 (1), q2(to), --- av(to) 41(to)> 42 (to), --Gw(to)] att = fo. In general, the initial conditions could also be specified in terms of any other set of 2N functions that are independent and invertible (in unique manner) to provide the above set. Because fg is arbitrary, it is clear that system can be described in ‘second order in time in some could determine the functions qi(t) a 2.2 Time Evolution of Dynamical Systems 43 the state ofthe system at any given moment can be specified uniquely by giving 2N independent quantities; that is, the state of the system may be described by point in a 2/V-dimensional space, called phase space, at any given instant. The evolution of the system is described by a curve in this 2V-dimensional space. Although this curve is unique, the coordinates used to describe it do not have any unique significance and are essentially decided by mathematical convenience. It is conventional to introduce in this space a general set of 2N coordinates de- noted by [qi(t), ga(t), «++ Gilt), +--+ 4w(03 PUD): P2()s «++ Pi(Os «++ PMO] that provide the same amount of information as the dynamical variables and their first-time derivatives. Because the N-dynamical variables satisfy second- order differential equations, there must exist a set of 2N first-order differential equations governing the evolution of the 2NV variables [q), pj]. The form of these equations depends on the nature of the system and on the choice of coordinates in the phase space. We are concerned with systems called Hamiltonian systems that satisfy the following criterion: For these systems, itis possible to choose (in ‘many different ways) the coordinates in phase space such that the differential equations can be cast in the form where H(p. q) is a given fun indices i, etc., on qi and merely write q for qi, etc., when no confusion to arise. Also note that we use the symbols qi and q’, etc., inter this chapter.) As we shall see in Section 2.4, this does not uniq coordinates in phase space and a wide class of them are still ‘quantities g; are the dynamical variables and the second set of, are called canonical momenta. ‘The Hamiltonian is assumed to be an additive quantity in ng sense: ‘The total Hamiltonian # fora system made of two noninte subeomponents Aand B, with variables (4, p) and [q'®, p(®), wil intobe the sum of HAL pA, gM] 4 the individual Hamiltonians, i.e., H[p, p,q), Hp), gy The time evolution of any other function, f(p, (and possibly depending exp! ‘We have ° fined on the phase space df _af a or where the second equality is based on the summation convention, which states that any index that is repeated in a term should be summed over. By using 44 2 Dynamics Eqs. (2.1) we can write this as df _ af pny th 23) where the quantity [H, f] called the Poisson bracket between H and defined as, is aH a aH a H, f= 2.4) ue (iran aq Bq ant ce) This is a linear first-order differential operator acting on f. In general, we can define the Poisson bracket between any two functions g and f of phase-space coordinates by the relation ag 9 ag 8 ) = (2% _ 38 2 2.5) ed (=r Baqi Bq dpi) * a It can be easily verified that the Poisson brackets have the following properties: Uf 3] =-Ie. fl. (2.6) (i+ f. 8) =[hi. 8] +0 8). Q [s. fifal=[e. fla + le. Alf, Uy [e, Al] + [e. (h, #1 + Uh, LF 81] = 0. eae te Pole enschede dynamical equations governing the system can be derived fro Itis obvious from Eq, (2.3) that (dH /dt) = (3H /at) = 0 ifthe Hg not explicitly depend on time; thus the Hamiltonian is a con: a closed system. Exercise 2.1 Filing the deals, Prove Bas. @.9-(29) A ore rolls) notation fora Poisson bracket is Dy f =[g. f], which shows that [g, f] is the actidnrof a differential operator (dependent ong on j Rewrite Eqs. (2.6)-(29) in this tvatfen and interpret the results. fained from a variational princi- this variational principle, let us The dynamical equations (2.1) can also ple, which turns out to be very useful. To consider the value of the quantity (calle ) As [anna — 0-9 ; ata f [pidq' — H(p.q)dt]_ (2.10) la for all possible functions p; (1), q' (1) that satisfy the condition q' (n) =a), 4! (t2) =q}. The evolution of the system is postulated to be such that it makes 2.2 Time Evolution of Dynamical Systems 45 the value of A a local extremum, Setting the variation of A to zero, we get agi 4p: ph aH). nth lesa) ae)ae] em where P; =(q1, 1), ete. The first term vanishes because 5q vanishes at the end points, The second term can vanish for arbitrary Sp! and éq' if and only if Eqs. (2.1) are satisfied, thereby proving the equivalence of the two descrip- tions. Note that the variational principle only demands that the end values of the dynamical variables q; be fixed; the momenta p are not constrained in any ‘manner. For a wide class of Hamiltonians, it is possible to invert the relation 4; = (8H /9p,) and express p; in terms of gj. In that case the quantity L = pq — H, called the Lagrangian, can be expressed asa function ofq and4;i.e., L = L(q. 4). ‘Then the action principle involves extremizing the quantity Ps A=| La@.aat Qu >, with respect to all functions g(t) that satisfy the conditions g(r) =41, g(t) The variation now gives GS aL bA= [ae (se! +59 mwekw) — &S “fale $C) (eh 0» qi 84: 84: 0=5A= * ar( sna + pig’ i oH - ay) Ip, = pibq’ The second term vanishes because 5g =0 at the end point aL _d (aL © eo 4(24) os a4; agi is Thus, in the case of systems for which p; can ressed in terms of gj, we can study the dynamics either by using the tonian H (p,q) or by us- ing the Lagrangian L(q, 4). Given the Lagi L(q, 4), we can obtain the Hamiltonian by first computing the quan = (@L/2q) and then comput- ing H = pq —L and finally expressing theYs in terms of p and q. It follows from the definition that the Lagrangi ‘additive in the same sense as the Hamiltonian, Equation (2.13) also allows us to draw another important conclusion. Sup- pose we evaluate the value of the action A[qi, 12; q/, 1] along the actual tra- jectory of the particle and ask how this function varies when the end point q} 46 2 Dynamics is varied. In this case, the first term in Eq. (2.13) vanishes as the integrand is identically zero along the actual trajectory; from the second term, we find that (8A/8q}) =(@L /84") = p}. Denoting the upper limit of integration by just (q, 1) and treating Afq, f] as a function of the end point, we have the relation aA _ ab aq: aq” ‘Thus we can determine both the equations of motion and the canonical momenta from the action principle based on the Lagrangian. This result will be of use in later chapters. ‘The Lagrangian is not unique in the sense that itis possible to come up with very different Lagrangians that lead to the same set of differential equations (see exercises 2.3 and 2.4). One simple example of this nonuniqueness is provided by two Lagrangians that differ by a total time derivative [dF (q. 1)/dt] of a function F(q. t) of the dynamical variables and (possibly) time. Such a total time derivative contributes to the action a quantity that is independent of the path and hence does not affect the equations of motion. But this does change the Exercise 2.2 definition of canonical momenta. ‘Symmetries and conservation laws: (1) Show that the quantity E(q. 4) corsahe)) is conserved if L does not explicitly depend on time. This quantity E, called of the system, is numerically the same as the Hamiltonian; note, howeveb iba) the Hamiltonian is function of ¢ and p whereas E is funetion of ¢ ang’? (2) A Lagrangian for a free particle cannot depend on the position S {(ihe}gquations of 2.15) the direction of the velocity; so L = L(0v?), Further, let us assume th motion are invariant under @ Galelian transformation, defined ES co + Vt. This requires L(v?) to change (at most) by a total time derivative unde where m is called the mass of the particle. Also show th: particles, we can maintain Galelian invariance if the Lagra ‘where the function V depends on only the pairwis (This resutt shows how symmetry consideratio Lagrangian. We shal see several more exam G)A Lagrangian L, describing the inter ales X),X2,--. es» w and the corgesgling time derivatives ky, a=1,2,-....N as in part (2) above. Ifthe origin of the eabdinate system is shifted by an amount I 0 that x4 —> x +1, then the Lagrangian should remain invariant for a closed system. ‘Show that this implies the conservation of the quantity P= S>, ma¥o, called the total ‘momentum of the system. (4) Consider next an infinitesimal rotation of the coordinate system by an angle 3 about some axis with unit vector n, This changes the coordinate by bx, = 482 m x x, ion of the particle coordinates. lus to determine the form of the jis feature in Chaps. 3 and 11.) 'N particles, depends on the coordi 2.3. Examples of Dynamical Systems 47 50 x x,. For a closed system, this change should not have any effect on the dynamics. Show that this implies the conservation of the quantity Ju J) m,x, x ve, called the total angular momentum of the system. , P, and J are all additive quantities in the same sense as the Lagrangian. Exercise 2.3, Equivalent Lagrangians: Consider a Lagrangian L(q, 4) given by da 1 =a f Gro « [e+vo}, ey) where F(¢) is an arbitrary nonconstant function of its argument. Show that this Lagrangian leads othe equation of motion g = — (dV /dq), independently of the choice for F. This result illustrates that there can be several different Lagrangians that lead to the same equation of motion. Exercise 2.4 More on equivalent Lagrangians: Given a Lagrangian L = L (q, @), we can construct a modified Lagrangian L’ defined by the relation (2:18) Show that varying the path q (r) in L’, keeping the momentum (3 /34) fixed at the eng points, will lead to the same equations of motion as varying the path q (r) in L keepin the coordinates fixed at end points. What is the crucial difference between L and L relevant part of the action that should be varied, is the first term in Eq. (2.AOf2 Shagy that this variation can lead to the differential equation for the spatial trajecto ae article Exercise 2.5, Cx -Maupertuis’ principle: For a particle moving with constant energy E ibe ‘moving with energy E. 2.3 Examples of Dynamical Syster ‘As examples of simple dynamical systems, we consid Se important cases that find application in different areas of astrophysics Mt# these Lagrangians have the form of Eq. (2.16). 2.3.1 Motion under a ‘orce As the first example, let us study a particl under the action of a central force derivatis'thom a potential V(r). The dynam- ical variable describing the system c fen to be the two (polar) coordi- nates 4; =(r, 8). The phase space is four dimensional and has the coordinates (7,95 Pr, po) and the Hamiltonian is given by the function ig in a two-dimensional plane h(a. H(p.a) = x Prt Jt VO (2.19) 48 2 Dynamics in suitable units. The equations of motion, corresponding to Eqs. (2.1), are (2.20) Given the initial position in phase space, these equations can be integrated to determine the future time evolution of the system. For this particular case, we can invert the relations 4 the Lagrangian: (0H /dp;) and obtain im @+PH)- von. (2.21) Lg.g The second-order differential equations, corresponding to Eq. (2.14), are d a (mi) = mr6? — V(r), (mr) = (2.22) a (rs glen’) (2.22) ‘These equations can be integrated if we know the coordinates and their first derivatives at any given instant of time. 2.3.2 Motion in a Rotating Frame & As a second example, let us consider the motion of a particle acted oF potential V(x) but viewed from a frame of reference that is rotating eye constant angular velocity ©. To obtain the Lagrangian for such a sy begin with the Lagrangian in the inertial frame given by 1 we L(x, 8) = jm? ~ VOX) S (2.23) and note that the velocities transform under rotation by the Vine = Wot + 9X, (2.24) where the subscripts inertial and rot denote the veloc the inertial and the rotating frames, respectively. Substituting Eq. (2.24) |. (2.23), we find the Lagrangian in the rotating frame to be oS 1 L= my +mv-(Qxx)+ x)? — Vx), (2.25) 2 where all quantities are measured in the 1 this Lagrangian, we get the equations dv_ av —S— + amv ax rame and V = Vey. By varying "a The second term on the right-hand side is called Coriolis force, and the third term is the centrifugal force. (This analysis also shows that the Lagrangian or Hamiltonian description is not tied to inertial frames of reference and can be + m2 x (xx Q). (2.26) 2.3: Examples of Dynamical Systems 49 easily generalised for any coordinate system.) The momentum and energy are now given by SE mvt mO xx, B= dm? mx +V. 227 Note that rotation changes the form of the potential energy by adding to it aterm [-(1/2)m( x x) that is quadratic in the coordinates. Forasystem of particles, constituting a rigid body, the Lagrangian in Eq. (2.25) has to be summed over all the particles in the system (with the same 2 for all particles). Then Vier = V gives the velocity of the centre of mass of the rigid body and © denotes the angular velocity of rotation of the body. In that case, the kinetic energy will be 1 1 1 THY) zm V+ Mx? =P 5mvV? + YI mV- Dx xt Ym (Oxx), (2.28) where the summation is over all particles in the rigid body. The first term is the kinetic energy of the centre of mass motion, and the second term, which can also be expressed as (V x $2). mx, vanishes when x is measured from the centre Xe ‘mass. Therefore the rotation of the system essentially contributes only the thir term to the kinetic energy. This term can be written in the form 1 2, Trot = pF VQ, Tie = Limes — 3x0). (2.39) The symmetric tensor Jj, is called moment of inertia tensor and le in rotation that is similar to the one played by the mass in the case tional motion. In the expression for Tyo, summation over i, ed as per summation convention. The sum in the expression for [ix Il particles in the system. In the continuum limit, Jiz becomes a= f Px p05 -a) 2.30) The angular momentum of the system, meas rypect to the centre of mass of the body, is given by (0) = [mx x (x 0)], Dm x(x 9); Any symmetric second-rank tensor cangkeSteduced to the diagonal form by a suitable choice of axis. These directi ~ given rigid body are called the principal axes of inertia and the corresponding values /1, /2, and Js are called the principal moments of inertia. The moment /,, for example, will be = hi = Dom? x3) = Dm} +3) = Domed, fx Qp. (2.31) 50 2 Dynamics where x1 is the perpendicular distance to mass m from the x, axis. This is the usual formula for computing moments of inertia of bodies with significant degree of symmetry. Clearly Troe becomes the sum of three diagonal terms in this case, and we also have My = 111, ete, in this limit, From the variational principle applied to the basic Lagrangian, it is trivial to verify that the angular momentum of the body changes in accordance with the law ((M/dt)=K, where K= — 3° x x VV is called the torque, corresponding tothe force F = —VV derived from the potential V(x). Since the time derivative of any vector A in the inertial and the rotating frames differs by the term $2 x A, the equation of motion for the angular momentum in the rotating frame is given by OM FaxMeK, (2.32) dt where the prime on (d’ /dt) is a reminder of the fact that the derivative is evaluated in the moving system of coordinates. Taking the components of this equation along the principle axis of the system, we get dQ HSE + Us — by 205 = Ki, & dQ; BSP +h ~ I) 23% = Kr, & 425 BSE + (h- W)@iM= Ky These are called Euler's equations and govern the rotational body. As an example of the use of Eqs. (2.33), et us consider (1) To begin with, if the body is totaly symmetric with /1 4 motion (corresponding to K=0) is trivial and is givep-b\-a;yiniform rotation = constant about some axis. (2) A more interesting, i free motion of a system with axial symmetry, for which [1 fey = constant and 21 and &z satisfy the equation 2) = —w2, ®; 1, Whtefe @ = 2313 — 1h) /T, ‘aiid adding to the second PA c08 wt; =A sin ot. esting with an angular velocity ining constant in magnitude. f inertia are unequal, the motion is a constant. Multiplying one of the equations equation, we can easily solve the set and obtaj ‘The motion therefore consists of the vector around the symmetry axis of the body In the case in which all the principal moi can be extremely complex. Exercise 2.6 Effects of Coriolis force: (1) Find the deflection of a freely falling body from the vertical, produced because of the rotation of the Earth, to the lowest order of approximation @) Determine the effect of earth’ rotation on the plane of oscillation ofa pendulum to 2.3: Examples of Dynamical Systems st the lowest order of approximation, [Answers (1) Taking the gravitational potential to be U = —mg-r, we need to solve the equation of motion ¥=2v x 9+ g, where we have ‘ignored the quadratic terms in 2. Ignoring the term that is due to rotation, solving for v, substituting into this equation, and solving again, we get the solution as re nbvatt der + ex MAW xO (234) If the z axis points vertically upwards and the x axis towards the pole and } isa northern latitude, the net deflection is given by y=—(1/3)2h/g)?gQcos2, with the nega- tive sign indicating eastward deflection, (2) Ignoring the vertical displacement of the pendulum and terms quadratic in ©, the equations of motions are ¢+ 0x = 205 5 +0%y = —20 i, where Qs the original frequency of the pendulum and 2, = sin 2. ‘The solution to this equation is concisely given by x(¢)+ iy(¢)=[x0(t) + dyo(O]exp (-i.1), where x(t), yo(t) is the trajectory of the pendulum in the absence of Earth’s rotation, Its clear that the net effect of rotation isto cause a shift inthe plane of rotation atthe rate 2, = Qin] Exercise 2.7 Feynman's plate: A thin plate is thrown into the air spinning with a slight wobble. Show ‘that it will wobble at arate that is approximately twice as fast as it rotates. (In Feynman's, ‘memoirs, he says that “the medallion rotates twice as fast as the wobble rate”; he has, ‘mistakenly interchanged the rotation rate and the wobble rate.) 2.3.3 The Reduced Three-Body Problem three-point particles under their mutual gravitational attraction SobjeRt,t0 the following approximations: (1) We assume that, ofthe three bodieg with masses m and mo, are heavy and the third one, with mas, is a test Particle, ice, m3 «my, (2) We take m; and m, to desefe ocular orbits around their common centre of mass; this orbit is ass unaffected by m5. (3) The motion of ms takes place in the ro o of m; and m2. This problem (called the reduced planar iets pl astrophysics because it is tractable and approximates mg For example, the motion of an asteroid (ms) in t the Sun (m1) and Jupiter (m2) can be treated al To simplify the notation, we make a choic (mj +m2) as the unit of mass; the two m; (1 — ) and jz. Next, we take the unit of (diameter of the circular orbit) betweey 1d mz. It follows that the radii of the orbits are yx and (1 — jz) for m; sspectively. Finally, we choose the unit of time such that G = 1. From these choices, itis easy to see that the angular velocity of my and m2 is also unity. It is convenient to use a system of axes (x, y) that corotates with m, and m2 with the x axis pointing towards m, say. In this frame m and m2 are stationary cr) ) is of interest in ng realistic situations. gravitational field of lines. as follows. First, we take and mz can now be called to be the constant distance 32 2 Dynamics with fixed coordinates (11, 0) and (1 ~ 12,0) respectively. If the coordinates of m3 are (x, y) then the equations of motion obtained from Eq. (2.26) can be easily shown to be 25 s S42 (235) where (2.36) with Ua B eR gate tup ty male +uP sy)? (237) By multiplying the two equations of Eqs. (2.35) by + and j, respectively, and adding them, we see that 20 — 3? — y? = C = constant. (238) This constant C, called the Jacobi constant, is essentially the energy defi the second equation of Eqs. (2.27). No other integral for this system is and hence the solutions have to be found by numerical integration discuss some characteristics of the numerical solution in Section 2.8. Kight ‘we concentrate on the equilibrium solutions that can be obtainegédirect the properties of the Lagrangian. Given a value for the Jacobi constant C, we must satis! 4? +5? =26 —C> 0. This shows that the motion is re bounded by the curve © = C/2, called the Hill curve. If. for ®, expressing x and y in terms of py and po, we get, 3 11 \ ° Sea-m(b+ Jor +n (or — 1. (2.39) & Clearly, ® is always positive and has a minim p2 = 1. These points (at which py = py triangle with the positions of masses is kept at either of these two points wi the configuration is stationary with icle located at the minimum of a potential, We thus find that there exi juilibrium solution to the restricted planar three-body problem with the threé’particles located at the vertices of an equilateral triangle. The configuration rotates rigidly with a constant angular velocity when viewed from the inertial frame. This case corresponds to C=3 for which the Hill curve degenerates into two isolated points that are usually denoted by the symbols Lg and Ls (see Fig. 2.1). jow rom inequality a region te Eq. (2.36) ie vertices of an equilateral Jt follows that if the third body ‘velocity in the rotating frame, 2.3: Examples of Dynamical Systems 33 | Fe, eLegmp pi erent pn ang GS) More generally, the equilibrium solutions will correspond to points KS the condition +0, § =0, (8/8x)=0, and (3@/dy)=0 are satisfied simul- correspond to equilibrium situations and are called Lagr: The Hill curve can be thought of as an equipotential li depend critically on the value of C. Figure 2.1 shows the. mofHilllcurvesas the value of C is increased from C = 3. The point marked, is the centre of mass of the system with masses my and ma where (m Close to the masses, the gravitational potential is dominated by the i jl masses and has circular contours. The same result holds far away frog two masses. At intermediate scales, the contours for potentials have the, thown in the figure. Exercise 2.8 Filling in the details. Prove Eqs. (2.35)-( 9). Itis obvious that the Lagrange points L, Lo, and Ls correspond to unstable equilibria and are local maxima of the potentials. For future reference, we give 54 2 Dynamics a fitting function for the distance to the inner Lagrangian point L from the two masses: [p 500 — 0.227 logio ()]. h= afo 500 + 0.227 logiy (2) (2.40) he where a is the distance between the two masses my and m2. The stability of the equilibrium solution corresponding to the Lagrange points L or Ls is more dif- ficult to determine. Linear stability analysis can be performed by examination of the eigenvalues of the matrix of second derivatives (3°¢/@.x;3x,). This analysis is straightforward, although tedious, and leads to the conclusion that the solutions are not linearly unstable for 4(1—y1) < (1/27). This corresponds to the condition 1_ [3 He ( ~V jog) © 0.0385. an that is, the upper limit on the ratio of masses is ~1 /25. A more complex analysis shows that the motion in the equilateral solution is actualy Snag is satisfied except for two specific values of the ratio jz. Thus, for all pr purposes, we can take the above criteria as the one for stability of the equilat solution. © Exercise 2.9 Siabilising influence of rotation: A system of masses produces a gravthfignal pgfential (6, 1), Prove that such a gravitational potential cannot have a local ‘minima ‘at any x unoccupied by the mass points. Explain how the rotation, com! th gravity, can Iead to a Lagrange point that is stable against small =e 2.4 Canonical Transformatios v Itis clear from Eqs. (2. eee eee solving 2/V-coupled first-order ordinary differenti these equations will depend very much on the cl and conjugate momenta. As it may be easie particular choice rather than with another general formalism can be developed to, now discuss some of these general tec| It was mentioned in Section (2.2) orm of the Eqs. (2.1) presupposes a judicious choice of coordinates in phas®Space. An arbitrary transformation of coordinates from the set (q, p) to some other set (q’, p’) will not maintain the form of Eqs. (2.1). We begin by asking the following question: What is the most general class of transformation from a set (q, p, H) to another set (Q, P, H’), that will preserve the form of equations of motion (2.1)? equations. The structure of f the dynamical variables fe these equations with one important to ask whether any solving these equations. We 24 Canonical Transformations 55 Because the same dynamical equations are obtained by varying the two dif- ferent actions, made from (pq — H) and (PO — H’), the two Lagrangians can differ by a total time derivative (4F /dt) of some function F, so that a pa H-(PO- HW) =F (2.42) or dF = pdq~ PdQ +(H'— H)dt. (2.43) If F = F(q, Q,1), it follows that ar OF yy OE fat Rage MEAtG, (2.44) ‘The function F is called the generating function for the canonical transforma- tion from (q, p) to (Q, P). Given any function F(q. Q,), we can generate a transformation from one set of coordinates in phase space to another such that the structure of Eqs. (2.1) remains the same. In particular, if F has no explicit dependence on time, then the Hamiltonian does not change in its numerical value. In the above case, we have taken the generating function to depend on old and new dynamical variables. It is also possible to have a generating function that depends on the old coordinates q and new canonical momenta P. To obtain t formulas for this case, we only have to note that Eq. (2.43) can be rewritten, d(F + P,Q:) = pidgi + QidP; + (H' — H)dt. The argument of the differential on the left-hand side, expressed i variables q and P, is a new generating function (q, P, 1), say. ao ao ~ OR" at We can also obtain similar formulas with other combination oFzatiables. The existence of canonical transformations, under whigtrtte Torin of the dy- or H=H+ namical equations remains invariant, denies any speci to the dynamical variables and conjugate momenta. For example, generating fune- tion F = qQ, we can produce the transformation q, thereby inter- changing the roles of coordinates and momenta, P ;nonical transformation, ities f and g, in which the ind q, and let [f, g]p,g be the ect to P and Q. Then Let [f, g]p.q be the Poisson bracket for t differentiation is with respect to the variat bracket in which the differentiation is wit UF Blea = Urele.o- (2.47) To prove this, we may argue as follows: Because time appears only as a para- meter in the canonical transformation, it is sufficient to prove the invariance for quantities that do not depend explicitly on time. Let us now regard g as a 56 2 Dynamics Hamiltonian for some fictitious system. Then we know that df/dt = ~Lf, &lp.q- But the derivative d//dr can depend on only the properties of motion of the system and not on the choice of coordinates in phase space. Therefore the Poisson bracket [f, g] is invariant under canonical transformations. Invariant volumes in phase space: The canonical transformation also leaves volumes of certain regions in phase space invariant Let dV = dg) ---dqwdp1 --dpw denote an clement of volume in phase space. The integral of dV aver some region of phase space represents the volume of that region. If we now replace the variables p, q with P, Q through a canonical transformation, show that the volumes of corresponding regions of space remain equal: foo fe davdnr dry = fd Q,--dQydP-dPy. 48) Next prove the invariance of a larger class of phase-space integrals of the form ff Xeadn [fff Cae anda don... 249) in which the integration is over manifolds of two, four, etc., dimensions in phase sp: (These are called Poincaré invariants. Because the dynamical variables and cor ‘momenta get mapped tothe corresponding quantities during the time evolution, i that the evolution of the system itself can be thought of as a canonical transf(tmefion from the initial set of variables to the final set of variables. Because th ph ime remains invariant under any canonical transformation, it follows that gh Volum ofthe phase space remains invariant during the evolution.) Having determined the most general set of transforma leaves the dynamical equations invariant, we now turn to the questi Ing this infor- ‘mation to solve for the dynamical evolution of the “ct Hamiltonian we can think of is one that vanishes i in the new set of variables, that is, H’=0. In this case P and Q wilkbe &)nstant and the evo- lution will be frozen in the new coordinate systery. The vanishing of the new Hamiltonian can be achieved by use of a gen function S that satisfies +H(F We obtained this by setting H’=0 third equation of Eqs. (2.46) and substituting for p in H(p, q) by p ); it is conventional to use the sym- bol 5 rather than © for the generating fuitetion that makes H'=0. This equa- tion (called the Hamilton-Jacobi equation) determines the generating function; by using this function we can go to a new set of coordinates and vanishing Hamiltonian. Suppose that this equation can be solved to obtain a particu- lar solution = S(t, q',a%), where a (with /=1,2,...,.) are constants of (2.50) ar 2.4 Canonical Transformations 37 integration. This generating function, of course, makes H’=0 by Eqs. (2.44) and (2.50). We now treat a as the new momenta P; (which are constant in time because H’ =0). From Eqs. (2.46) we see that the new coordinates are Q; =(8S/8ai). But because H’ =0, these coordinates should also be constant; O' = constant = pi, say. We thus get the equations A; =[8S(t, q:,«)/da}, which implicitly determine the trajectories qj(t) in terms of the 2V constants (qi, fi). Thus we have reduced the problem of dynamics to one of solving the partial differential equation (2.50). The generating function S has a simple physical meaning; it actually corre- sponds to the action A treated as a function of the coordinates and time at the end point, that is, the functional dependence of $ on q and t will be the same as that of the quantity a Aao= f Lat, si) where the integral is taken along the extremal path that is a solution to the equation of motion. To prove this, we only have to show that (B) 1G) emg The first relation was proved earlier; see Eq. (2.15) in Section 2.2. To roel) second relation, we note that d.A/dt = L by definition, and we can also eS) dA aA aA, aA Tent aga a th Se a5) Comparison gives (8.4/8t)=L — pid or aA Go a-u. Q (2.54) 5c This shows that the solution to Eq, (2.50) is actually n treated as function of the end points. ‘As an example of the use of Eq (2.50), let us consider the motion of a particle in one dimension under the action of a potential V4 @ Hamiltonian for the system is Hn = (2.55) and the Hamilton-Jacobi equation is 2 as, 1 (as 8 1) veya . - u(#) +V@)=0. 2.56) This partial differential equation can be solved by the method of separation of variables with the ansatz $= —Et + A(x), where E is constant. By substituting 38 2 Dynamics this ansatz into Eq. (2.56) and rearranging terms, we can express A(x) as an integral, giving S=-Et+ f dx JIm{E— VO) This gives the solution to the Hamilton-Jacobi equation in terms of the coordi x, time f, and a constant £. We think of this function as a generating function, with the new canonical momentum identified with E. We also know that, by construction, the new Hamiltonian vanishes. From our canonical transformations in Eqs. (2.46), the new coordinates are given by (3/2). However, because the new Hamiltonian vanishes, the new coordinates are also constant in time. Equating (9/3 E) to some constant fo, we get wa tro faorlae] (2.58) This equation provides the complete solution to our problem. Given the form of V(x), we can now express x as a function of r and the two constants fo and E. The constants can be determined if x and # are given at some initial instant of time. We thus note that the problem of a particle, moving in one dimension under the action of an arbitrary potential, can be reduced to a quadrature. Exercise 2.11 € Potentials for which period is independent of amplitude: (1 Consider pari itasilates between x= a and x= a. How does the period of oscilla ‘amplitude a? (2) In particular, note that the period is independent of amy power-law potential only itm =. There exists, however, very many di V (2) (which are not pure power avs) such that patclcan oscil of the potential with a period independent of amplitude. Give a g SHaracterisation of sucha potential (3) One particular example of sucha potentials Sx (2+). Argue why the period of oscillation shouldbe independent of ampliiigein such a potential without doing any calculation. (Hint: Think of a tone frmonic oscillator.) SS ‘As a second example, consider the motion aiticle in a plane under the action of a central force with potential V‘ of the Hamiltonian in Eq. (2.19), the Hamilton-Jacobi equation b as, 1f/as\? 1 4 * aml lar) FB ‘We now substitute the ansatz = ‘equation and obtain > \qir A far[an(e-v-5)| . 2.60) +V(r) (2.59) (6 + A(r) into the Hamilton-Jacobi 2.4 Canonical Transformations 59) so that 2 \ qe -er4 0+ f drfan(e-v-5)] (2.61) Once again S(t;r,@; E, J) can be thought of as a generating function for a ‘canonical transformation depending on the old coordinates r, # andnew momenta E, J. Because the new Hamiltonian vanishes, the new coordinates, given by the derivatives of S with respect to E and J, must be constants. With these constants denoted by to and 6, the equations fo m rene far (alt=) : 2.62) _ We ont f dre Ppamryy ‘These equations provide the complete solution to the problem. Equation (2.62) relates r to r, and Eq. (2.63) relates 6 and r. There are also four arbitrary con- stants, J. E, to, and 6p, that need to be determined in terms of r, 6,7, at an initial instant. The problem has again been reduced to quadrature. © ‘The two examples above show how the method of the Hamilton—Iacobi equa’ tion works in practice. It must, however, be remembered that the success was tothe fact that we could separate the variables in the partial differential equa It is this feature that allowed us to reduce the partial differential equat the second example, the choice of coordinates also played a key used the Cartesian coordinates x, y rather than the polar coordi problem would have been less tractable. There has been exteng separability of the Hamilton-Jacobi equation in different It can be shown that there are 11 real coordinate systems th sfy the cri- teria for separability, but only a few of them are useful ani ying interesting physical phenomena. If the equation is not snp must resort to ‘numerical techniques, and it is better to start with Eqs. (2:¥ather than with the ‘Hamilton-Jacobi equation. o Exercise 2.12 Separabiltyof Hamilton-Jacobi equation in anggxidailiey coordinate system: Consider the motion of a particle under the action of a\petetial U =(a/r) — Fz, which is a combination of a Coulomb field and a uni ld. (1) To study the motion of such 4 particle, it is convenient to use parabolic éddndinates (E, ,n) that are related to the standard cylindrical coordinates (p, @, z) by the equations z=(§ —n)/2 and p = YEH. Show that the Hamiltonian fora particle can be written in these coordinates as 2bpit nme , Pb m §4n * Iméq +UE.n 9%). (2.64) 60 2 Dynamics (2) Show that the Hamilton-Jacobi equation is separable in these coordinates for any potential ofthe form alg) + b(n) _ alr +2)+ 0 ~ 2) u etn or 2.65) ‘The potential U=(a/r)—Fz has this form, with a(é)=a—(FE?/2),b(n)= a+ (Fn? /2). Hence obtain the complete solution to the problem. ‘The fact that there are only a limited number of cases in which a complete solution of the problem is possible gives rise to two important questions that are central to the investigations in modern mechanics. Under what circumstances can we reduce a given dynamical problem to quadrature? Can we say anything about the general characteristics of motion for systems that can be reduced to quadrature? We now turn to the discussion of these questions. 2.5 Integrable Systems To understand the nature of systems for which the problem of motion can reduced to quadrature, we need to introduce the concept of an integral of tion, Note that, if the equations of motion can be integrated, then we can the functions gy = qi(tsq°. p®) and py = pi(tsq?, p2) in terms of thei ditions (g°, p*). In principle we can invert these relations and expreé& thé)2V quantities q? and p! in terms of qj, p and t. By their very congtryctiop-these 2N functions of qj, pj and: are constants, thats, their values reiyaarthe ame as the system evolves. Out ofthese 21V constants, ne constants t origin of time is immaterial for a closed system. We can always changer (+ fo) in a parameterised curve C(t), keeping the physics invariant, us) ny dynamical system has (2NV — 1) nontrivial constants of motion that fdstie€the system to a [2N ~(2N ~1)]=1 dimensional surface in phase oe is one-dimensional surface is the trajectory C(t) Most of the (2N — 1) constants of motion describ less in the sense that these constants can be det of motion are integrated (and hence are of For certain physical systems there might exis fa(Q. P), A=1,2, ..., k that are indeper ‘general symmetry considerations. Such dent of time, are called integrals of mi -ven among the integrals of motion, all are not of equal importance. Cer jegrals of motion isolate the time evo- lution of the system to a restricted region in phase space, thereby simplifying the physics. Such integrals are called isolating integrals. The integrals of motion that do not isolate the motion to a well-defined region in phase space are called nonisolating. are, however, use- 1eonly after the equations studying the dynamics), set of k constants of motion and that can be detected from ts of motion, which are indepen- 2.5 Integrable Systems 61 To illustrate these concepts, let us consider the example of a two-dimensional harmonic oscillator with the Lagrangian 1 1 L= 50 +5) — s(0ix? + oy") (2.66) In this particular case, we can easily integrate the equations of motion and obtain the solutions X(t) = Acos(wst +€x), y(t) = Beos(wyt + €y), (2.67) where (A, B, €,, €,) are the four constants that are to be determined in terms of [+(0), »(0), p.(0), py(0)]. From the above solution we get x= Acos(ort +6), Px = —Aos sin(oxt + €), 2.68) Y=Beoswyt +6), py=—Booy sin(wyt + €5). We can now express (A, B, €,,€y) in terms of (x, pry ys Pys!) Maret, BP 2.69) cS by using the arbitrariness in the origin ofr, we can eliminate one of them. (~~) Because the Lagrangian depends on both x and y we expect only the to be a conserved quantity. This will define a three-dimensional dimensional phase space in which the motion takes place. The ab shows, however, that it is possible to write two independent inte! for the system because the x and y motions are uncoupled. We hav (A, B) or, equivalently, (Ey, Ey), where = py planes. Be- 2 dimensional , given by Eqs. (2.71), surface in the phase space. The equations of thi can be written parametrically as 2E, p= ViBsina, r= fz /2E, sin B, This is the two-dimensional surface of a torus on which the motion takes place. @ and F). Inverting this relation, we can express the Hi jn (which is the same as F) in terms of I, and 2, The rest of the analysis will proceed as described above. As a particular case, let us consider the motion under the potential a_B voy= 2-5. (2.98) leading to the actions n the turning points in r. Given 2.5. Integrable Systems 67 The two actions are h=F,, (2.99) Yama 1 1 h Ef rear r= [nbs te( 1” —y —2mB + FE +a( . (2.100) err (a) e100) Solving for H = F; in terms of fy, I, we get ma = z (1+ (i =2mp) From the general theory, we know that the corresponding angular coordinates are linear functions of time: ©, = oat + €a, where the two frequencies are given by Hh. b)= (2.101) aH oh aH ma? hb Ch ow = 2H —— oh (1+ (8 -2mp) /B -2mp ~ ‘We can easily express the old coordinates r, 6 in terms of @4 fase of a ‘two-dimensional oscillator and thus solve the problem comp! Expressions (2.102) and (2.103) have some interesting fea fe know that, o (2.10% in general, the motion will fill the surface of a two torus (ception arises when the two frequencies are commensurable, ie, their ‘ational number, This ratio is given by (2.104) If B 0, then, this ratio can be rational, 1 — 2mB/13)"? is rational. If that happens, then the motion will be e¢ to a closed curve on the surface of the two torus, making the motion one dimensional rather than two dimensional This reduction in the dimensionality of motion implies the existence of another isolating integral. To find this integral explicitly, we can proceed as in the case of the two-dimensional oscillator. By using Eq. (2.63) we can find the orbit of 68. 2 Dynamics the particle to be (1 - ten" + Qeos [6 - Bt)" es]. (2.105) where Q=[1 — (2mB/J?)]-{a2m?/J* + (2mE/J?)1 = 2mB/J?)}}"? and is another constant. Solving for 6, we get “1 2) \] (2.106) noe (1 Bt) cw [ 44 on This quantity 6 is clearly an integral of the motion. If itis single valued, there will exist another isolating integral of motion, reducing the motion from two- to one-dimensional space. In general, cos~! z will introduce a 277m factor, giving ana (1-22) "owt 2) Jo where Cos~'z (with an uppercase C) denotes the principal value 6 can be single valued only if (1 — 2m /J2)"? is a rational nurft we have an extra isolating integral. The nature of the orbit in real space reflects the dimensi in phase space. In general, the projection of phase-space st two-dimensional torus) will fill a two-dimensional regigniin th, 6 plane. But when the frequencies are commensurable, the ses ‘aone-dimensional region in phase space and its projection will be a cloféd opt inthe r, @ plane. The situation is more interesting in the case of 6 =BewWhich corresponds to the Kepler problem. In this case, the two frequenei always spans a one-dimensional region of the, orbit is a closed curve in real space. Once ai integral of motion, which can be found alot Consider a particle moving under a differentiation, itis easy to see that identical and the motion ace; correspondingly, the jere must exist an additional following lines. nce fir) = — V'(r)F. By direct foexaya—mpnr, ge Da —mMfoOre. (2.108) where p, J, and r are the momentum, the angular momentum, and the position of the particle. For the special case of inverse-square-law force, a = f(r)? isa 2.5. Integrable Systems 9 constant and Eq, (2.108) implies that Aspxd- Sr (2.109) is conserved. This conserved vector, called the Runge-Lenz vector, has three components, but it satisfies the constraints A= ImPE+arm, Av 0. (2.110) The first one shows that the magnitude of A can be expressed in terms of other constants of motion; and the second one shows that A lies in the orbital plane. These two constraints reduce the number of independent components in A from three to one, This extra, independent constraint lowers the dimensionality of motion from two to one. In this particular case, we can easily find the orbit of the particle by taking the dot product of A with r and by using the identity r -(p x J)=J-(r x p)=J?. This gives A or, in more familiar form, 1 am A 1-214 Aone) ang We see that A is inthe direction of the major axis ofthe ellipse. For future reference, the explicit solution for the Kepler problem in Ce cases is given here. By using the first equation of Eqs. (2.110), we a trajectory of Eq, (2.112) as o ta} tecos, (4D, San ral — Di where the semimajor axis a and the eccentricity ¢ of the ao by € z (2.114) = Arcos@ = J? — amr, Qa) §—§) (E> 0), (2.115) —esing) (E <0). The above examples clearly illustrate that, if the system is integrable, then the motion takes place on an N-dimensional torus and — in general — fills the surface. The exceptional situation arises when the ratio between any two of the 70 2 Dynamics frequencies is a rational number. In such a case, the motion is confined to a space with a lower dimension than N. When such a situation arises, there will exist another isolating integral of motion. In the case of the two-dimensional harmonic oscillator, this is given by Eq. (2.75); for the Kepler problem, itis the Runge-Lenz vector. Exercise 2.13 Hodograph for Kepler problem: The curve traced by the tip of the velocity vector of a particle moving under some force law is called a hodograph. Show that the hodograph ‘ofa particle moving in the Kepler problem isa circle in the case of bounded motion and, part ofa circle in the case of unbounded motion. Exercise 2.14 Kepler motion of a binary system: Consider two stars of masses m, and m2 orbiting around the common centre of mass (in an elliptical path) under their mutual gravitational attraction. The plane ofthe orbits inclined at an angle ito the line of sight connecting the observer to the focus. From the Doppler shift of the radiation emitted by star 1, it is possible to determine the velocity of that star along the line of sight. Show how this information will help us to determine the combination & 2 sin? msi ae t (my + m2? ©) ‘This procedure is of importance in determining the mass of an invisible coon ey binary stellar system 2.6 Adiabatic Invariance & ‘The Hamiltonian of any physical system depends on some Cyne that characterise the system and are treated as constants ye om evolution. For example, the frequencies co, of a two-dimensional for or the mass of the Sun in planetary motion belong to this set. The ‘sical situations in which such parameters may change slowly with time ‘than remain strictly constant (for example, the mass of the Sun may slowly decrease because of solar wind), We would ike to understand how the say ehge in the parameters ofthe Srmalism developed above ~ hay in this case. dimensional system executing a peter in the Hamiltonian that varies 2) energy E of the system would be conserved; when 2. varies slowly, E would not bbe conserved but we expect the rate of change of energy with time & to be small. If we now average this rate over a period T,, the rapid variations will be averaged out and the resulting value of (£) will give the secular slow change in E that is 2.6 Adiabatic Invariance 1 due to the variation i. Because the variation of 2 leads to an average variation of E, we can express the averaged F as some function of A; equivalently, we can express this relation as the constancy of some function of E and 4. Such functions, which remain constant when the parameters of the system are varied slowly, are called adiabatic invariants. We now show that the actions J, ate adiabatic invariants and remain unchanged when the parameters of the system are varied slowly. Let H[q, p, A(t)] be the Hamiltonian of the system that depends on the slowly varying parameter A. The rate of change of energy of the system is aH dE _@H _dHdd cere 2.07 dt ar aA dt t y On the right-hand side, 3. is a slowly varying quantity but (9H /9A) depends on the rapidly varying quantities q and p. To determine the steady, secular change in E we average this expression over the period of motion. Because 4. varies slowly, we can take it outside the averaging and write dE di (OH (ez) <2 (244, ome where we treat 2. as a constant in the function (9/24) that is being were) Expressing the time average explicitly as aH 1p’ aH (l= ae ww and transforming from dt to dq by dt =dq/{(H/ap)], we find Ge) () (2) §@H/9A)dq/(@H [apy TAS (2.120) fdq/(@H/p), The circle on the integral sign denotes that the meres is taken over the complete range of variation of the coordinates durirfgne)period. Because these integrals are taken along paths in which the Hi mn has a constant value E, we can express the momentum as a definit ion of the variable coor- dinate q and the two constant parameters E Setting p= p(q;£,2) and differentiating the equation H(q. p. ect to A, we get (AH /8A)= —[8H /ap)IL4p/94)) or (@H/92) ap (@Ajap) 121) By using this in Eq. (2.120) and writing the integrand of the denominator as, n 2 Dynamics (@p/9E), we ger ae =-(2) senna, 122) dt dt ) f@p/AE)dq that is, Bp [dE\ | apa 4, _ $ (2G) +23 dq =0. 2.123) This is equivalent to the condition (df /ar) = 0, where - yf MED f rae 58, (2.124) and the integral is taken over a path with given E and 2. This result can be directly generalised to systems with more degrees of freedom and shows that the actions introduced by Eq. (2.83) are adiabatic invariants, ‘Note that adiabatic invariance of actions requires that the change in the para- meter be small during a complete cycle of motion, that is, we must have Ti. < 1, where T is the period. We can imagine situations in which this condition is violated even though the external parameter changes slowly in some other pr specified manner. Adiabatic invariance will not be a valid approximation ung such circumstances. As a first example of adiabatic invariance, consider a harmonic whose frequency is varying slowly. Because the action is E /o, we can Kone}\de that the energy changes in proportion to the frequency. ‘A more interesting example is given by the Kepler problem, orbit is an ellipse with the equation 1 al tecose, I=, € x where 2! = 2a(c? — 1) is the latus rectum, ¢ is the <=) [see Eq. (2.113)], and J and E are the conserved angular momentum an , respectively. The actions in this case are 1” h h 2.126) [see Eqs. (2.99) and (2.100) with 6 = 0] 488asy to see that hy 7 (2.127) “ (s +h ) a Suppose now that the parameter m or a changes slowly with time. Because the actions /, and f; remain invariant, the eccentricity of the orbit does not change. On the other hand, the scale of the orbit determined by / changes inversely 2.7 Perturbation Theory for Nonintegrable Systems B as ma. These conclusions, which are straightforward to obtain from action- angle variables and adiabatic invariance, are not so easy to derive from more direct approaches. 2.7 Perturbation Theory for Nonintegrable Systems Fora closed system with 2.V-dimensional phase space, the only obvious constant of motion is the Hamiltonian. This restricts the motion of the system toa (2N — 1) submanifold. On the other hand, for a fully integrable system, the motion is confined to an NV-dimensional submanifold, which is of lower dimensionality compared with (2N ~ 1) except in the trivial case of N= 1. If the system is made of several particles, each moving in three-dimensional physical space, then homogeneity of time and space and isotropy of space will lead to the existence of a total of seven integrals of motion corresponding to the conservation of energy, three components of linear momentum, and three components of angular ‘momentum. For a system with k particles, the phase space is 6k dimensional and the existence of the above conserved quantities confines the motion to a (6k — 7) dimensional space. If the system is integrable, then the motion will be confined to 3k-dimensional space, which is of lower dimensionality forall k > 2. Itis clear that integrability puts a severe restriction on the region of phase space that t dynamical system can explore. This prompts the following question: Is integrability a rule or an ex« In other words, do all Hamiltonian dynamical systems in 2 -dimension, space possess NV integrals of motion? The question is nontrivial heca take the point of view that all systems are integrable and our, ¥ to find N integrals of motion is only due to lack of mathematical exp In that case, the Grew he structure of the a region of phase space of Hamiltonian is a very small subset of the class of all possible Hamil smallest perturbation to an integrable system will di tori, and the system will — in due course of time ~ higher dimensionality than M. These issues are of importance in the study, especially in the study of the long-term evolug moving under the inverse-square-law fore, system, and its trajectory in phase space, Jaw were changed by the addition of cube term, say, the system would still be integrable. (In the first case, the trajectory is closed on the torus, whereas, in the second case, it will fill the torus.) However, from the point of view of celestial mechanics, it is important to know what happens to the trajectory when the perturbations from the other planets are taken into account. The system can i astrophysical systems, f planetary systems. A planet un constitutes an integrable ified to a torus. Even if the force 4 2 Dynamics no longer be described by a central potential and hence we cannot, by inspection, obtain NV integrals of motion, We would like to know, in this case, whether the perturbation destroys the tori completely or whether they survive. Detailed investigations have shown that most of the tori persist under pertur- bations, although in a distorted form, Some of them are destroyed (and these do not forma set of measure zero), with the measure of destruction growing with the perturbation. The destroyed tori are distributed among those that are preserved in a very pathological manner. These assertions were rigorously proved in what is now called the Kolmogorov-AmnoldMoser (KAM) theorem. Some features of this analysis are now described. Letus consider the effect ofa perturbation on an integrable system by assuming that the total Hamiltonian is given by H(I, ©) = Ho(1) + €M(I, ). (2.128) Here Ho(1) is the Hamiltonian corresponding to an integrable system and J, © are the action-angle variables for this system. [Both / and © denote N tuplets (Ia, 2) with a= 1, 2,..., N. We suppress the subscripts when no confusi likely to arise.] The second term €#,(J, @) is a perturbation characterized by. the strength €. Because the total Hamiltonian depends on © (through the a turbation), / and © are no longer action-angle variables for the whole syst although they are perfectly good canonical coordinates. If tori exist for t turbed system, then there must exist a new set of action-angle varie) such that HU) = Hl’), 129) and the new variables must be related to the old by a canonic: ‘mation produced by some generating function S(, 1’ e with as b= 7, % aa Q (2.130) ‘Substituting these relations into Eq. (2.129), we ey \dition that $ must satisfy: H{VeS(@, 1'), 0] = Whether the tori exist or not is thus reduced t solutions to this equation. Letus examine whether a solution can: bation parameter €. The zeroth-order the identity transformation, Therefor S= OK +50.) +--. (2.132) 2.131) Substituting this expression in Eq. (2.131), we get Holl! + €VoSi ++) ber (I+, O)= HW), (2.133) 2.7 Perturbation Theory for Nonintegrable Systems 8 with the notation VeS =(3S/3@,), etc. To first order in €, Ho(U') + [Vr Holl’) VoS\ + Ail’, OY] =H). (2.134) We now note that apa(’) = [8 Ho(I')/A 1] is the frequency of the unperturbed ‘motion and that Hi and Si can be expanded as periodic functions in ©: Hy(I',@)= )> Him(Z')exp(im - ©), = 2.135) 510.1) = D> Sin(d')exp(im ©). aa mz (Note that H; and S; are functions of q and p, which in turn are periodic fune- tions of ©.) Substituting Eq. (2.135) into Eq, (2.134) and equating the Fourier coefficients, we get HI!) = Hol) + Holl’) +++ (m= 0), (2.136) Sia(t’) = +i Hiol) a £ 0) (2.137) maooa() Thus, to the lowest order, the generator of the new tori is given by Him(!') S(O, 1')= 0-1’ +ie > expt Ot. 2, mn It might appear now that we have achieved what we originally ee mana" tothe lowest order in perturbation theory, we have a generating fu lead to a new set of action-angle variables for the full system a new set of tori. By repeating this process order by order in €, full generating function and the tori, thereby showing that pert integrable system still integrable. These conclusions, however, are wrong. The key dil 8 from the denominators in Eq. (2.137) that have the terms maa. wer these quanti- ties vanish, the perturbation theory breaks down. This mn arises when the smijjhsurable — that is, if elicies are in resonance. le, itis always possible IL This raises the doubt as ve to worry about both the a8 well as the convergence of the frequencies of motion on the unperturbed torus are the original orbits are closed and the fundamental In fact, even when the frequencies are incor to find a set of mg such that maina is arbit to whether the series ever converges. Note convergence of the sum over individual series in powers of €. In the study of planetary systems etc., tie unperturbed motion usually does not have commensurable frequencies (although there are some very important and interesting exceptions in the solar system, as we shall see in Vol. II). In that case, the trouble arises for only large values of ma. For such large values of ma the corresponding Fourier coefficients Him are very small, If we truncate the sums 16 2 Dynamics before these large values of m, come into play and work with a few low-order terms, we can predict the motion for a long time. The success of these predictions is important from the practical point of view of celestial mechanics, but does not help in answering the key theoretical question of arbitrarily late-time behaviour of the system, Significant progress in this problem can be made if the nature of the perturba tive analysis is changed. Instead of expanding in a power series about the original Hamiltonian, we can do a perturbation series with the nth term defined as a per- turbation around the system defined by all terms up to the (n — 1)th term. From such an analysis, we can prove the following result: The process of generating (perturbed) tori does converge almost always for a small but finite €. So most of the trajectories stay on an V-dimensional manifold for all times and do not explore the 2/V — 1 dimensional energy surface. More precisely, al initial tori, whose frequency ratios are sufficiently irrational, are preserved. The condition for sufficient irrationality may be stated (in the case of N =2, for example) as, the following condition on the ratio of the two frequencies wp) and «n>: Bae 2.139) for all integers r and s, where K is a number that is independent of r and s tends to zero as ¢ tends to zero, Note that this condition says the frequen should be sufficiently far away from any rational number r/s. The tor} excluded by this condition are the ones that are likely to be seriously aftessed the perturbation, & Let us estimate the measure of these tori. Taking 0 < (op / without any loss of generality), we remove from the interval [0, 1] all pat it violate condition (2.139), i. all patches that satisfy jor} Ke) co es) 2.140) jon 5] 8 cae ‘Thus we take off a region of length (K /s*5) aurea rational r/s in the interval [0, 1]. This amounts to taking away a totaf\length that is less than Ky which tends to zero as the perturbation denotes the number of r values wit ‘on 0 to 1; actually, the result in Eq, (2.141) is an overestimate because We have double counted neighbourhoods which overlap.] Clearly, we are left with a set of nonzero measure, even after we remove the tori that are close to rationals. What happens to these tori, which originally had commensurable [or nearly commensurable, in the sense of condition (2.140)] frequencies? Numerical = 2.6K, (2.141) ler. [The factor s in Eq, (2.141) 28 Surface of Section 1 investigations suggest that these tori are destroyed and the orbits in them are redistributed in phase space. The low-order resonances in the original system are almost always destroyed by the perturbation (and can lead to interesting phys- ical effects, which we will study in Vol. II), although the fate of higher-order resonances is difficult to study. 2.8 Surface of Section It is clear from the discussion in Section 2.7 that the nature of motion in the phase space for nonintegrable systems can be extremely complex. In principle, ‘we can attempt to study such a system by numerically integrating the equations of motion starting from some initial condition. Although this approach is of great utility (and is often resorted to) in understanding the behaviour of specific systems, it does not, by itself, help us to understand the general characteristics of motion, To do so, it is convenient to represent the key features of the motion by use of a technique called surface of section. To illustrate this technique, let us consider the motion of a particle in the x-y plane under the action of a potential (x, y). The phase space for system is four dimensional. Conservation of energy will restrict the motion to a three-dimensional surface in phase space. The path of the system C(t) on thi three-dimensional surface will cut the y = 0 plane in a two-dimensional surf which can be characterised by the two coordinates x, and p,. We can valuable insight into the nature of the motion by plotting the intersectio (0! curve C(t) with the y=0 surface. In general, the motion will beggpreathover a bounded region in the x, pr plane, as we expect the motion dimensional surface. If the system has any other hidden integr then the motion can become one dimensional and the points C(0) with the y =0 surface will lie on a curve. For a system with N=2, the phase space is four dimerf§iomatwith the co- =e ordinates x, y, px, and py. Conservation of energy res three-dimensional surface defined by stat 2.142) 1 3 (Pi + Py) +009) = B= (We shall take m = 1.) The intersection of thi the surface y =0 defines the surface of secti which any point has the coordinates x, an 0, we can determine py as fimensional surface with x the dynamical problem, on ven (x, px) and the condition py = +{2[E 980] — p23}, (2.143) ‘where we have chosen the positive square root by convention, Given the energy E and suitable initial conditions, the equations of motion can be integrated to ive the orbit of the particle. This orbit will repeatedly cut the surface of section B 2 Dynamics at different values of x and ps. Ifthere are no other integrals of motion, we expect, these points to be scattered in a two-dimensional region bounded by the curve bet +00 SE. (2.144) In case the system has some other hidden integrals of motion, the orbit will be confined to a lower-dimensional surface in phase space and hence will form a smooth curve in the xp, plane. In general, very different kinds of surface of section are generated by different Hamiltonians. Even for a given Hamiltonian, the nature of the points (induced by the motion) on the surface of section can be widely different for different initial conditions. Toillustrate these concepts, we consider the motion of three-point particles un- der their mutual gravitational attraction subject to the approximations discussed in Section 2.3. As in that section, it is convenient to use a system of axes (1x, y) that corotates with my and m2, with the x axis pointing toward m, say. In this frame m, and m are stationary with fixed coordinates (~i1, 0) and (I — 1,0), respectively, where j= m2/(m; +m). If the coordinates of m3 are (x, y) then the equations of motion are given by Eqs. (2.35). This system can be brought, into @ Hamiltonian form with 2 degrees of freedom by choosing N=, pay, paytx, @Q 1 l-u pala — mn LZ, 3 (Pi + PB) + Pian ~ pa = = =F integral of motion; itis conventional to define it in terms of a q Jacobi integral, which is given by C= — 2H. In terms of the off C is given by cary é Pi as Here C differs from that in Eq. (2.37) by only a consi -H). No other integral for this system is known, and hee the trajectories have to be computed by numerical integration. Let sidér the surface of section defined by y =0, j > 0 forthe case of equal = 1/2). Figure 2.3 shows the surface of section for C h the system is nearly inte- grable and all accessible regions are cove lI-connected regular curves. The otbit itself shows great regularity; 3 staying comparatively close (2.147) ‘C =4,the surface of section changes to that shown in Fig. 2.4. Now the curves do not fill the accessible region in a regular fashion. All the isolated points correspond to a single trajectory that has ‘a chaotic character. For a still lower value of C =3.5 (see Fig. 2.5) the chaotic region increases in extent, The orbit has a very disordered character. In a typical 28 Surface of Section ” i 05 0 08 1 Fig. 2.4. Surface of section for restricted three-body problem for C= 4. 80 2 Dynamics ‘motion, m3 will describe a few loops around m, then jump to m2, thep my, ete, This example shows clearly that the dynamical evolution of pig Took- ing systems can be highly irregular. What is more, the behaviouSfth Poincaré section changes drastically when some parameters inthe proble1 slightly, The motion of the particle as depicted in the Poincaré sect fges from a fairly regular form to an aperiodic, chaotic form when ters of the problem change. ‘Some aspects of this transition to chaos can be under: lines. The points in the surface of section are gener: en the equations of ‘motion are integrated for a particle, Mathematically we Sah model such a system by giving a map of the form xy..1 = f(%,), Whereadigives the coordinates of the nth point in the surface of section and the fu is a rule that allows us to determine the (n + 1)th point from the ntl his suggests that we should study the evolution of the coordinates un fe maps in order to understand the dynamics of complex systems. Si ig the model still further, let us consider one-dimensional maps of t net = Fn, He), Where 1 is some parameter in the problem. Given a particular value of i, the form of the function ‘Fle, 4), and a starting value xo, this rule will generate a succession of points ‘Xo, }, 19... Weare interested in studying the general features of this sequence. To begin with, it is clear that the sequence will repeat itself with a period k if Xn¢k~1 = f(%n, 14). Given the functional form of f, we could examine whether 2.8 Surface of Section 81 such nontrivial solutions exist for specific values of x and yx. Further, if k=1, if tq = fn, 12), sucha value of xy is called a fixed point of the map. The structure of mappings can usually be understood in terms of the fixed points and the periods in the sequences. ‘Asan example, consider the mapping generated by the function f(z, 1) =4uz (1 ~2), where jis a parameter in the range 0 <1 <1. This function maps the interval (0, 1) to itself. The fixed points of this map are the roots of the equation z=4yz(1 ~2) and are given by (1) z=0 and (2) 2=[1 — (4u)"']. The first solution is trivial and exists for all values of j2, whereas the second solution exists in the relevant interval (0,1) only if > (1/4). ‘These fixed points are the solutions to the equation z = f(z, 4). Letus consider the effect of varying the value of x slightly around the fixed point. We do this by writing x» = (z +&,) and expanding the map in a Taylor series in €. This leads to the relation & +1 = f'(@, 14)En, where the prime denotes the first derivative with respect to z. This equation admits solutions of the form &, = cL f"(z, 1)J", where c is a constant. The perturbation &, will decay with successive iterations if and only if | f(z, )| < 1. Such a fixed point is called stable. For the map discussed above, the nontrivial fixed point in (2) is stable if < (3/4). Thus in the range (1/4) <1 <(G/4) there exists a nontrivial, stable fixed point for our mapping, at 2=[1 —(4.)""]. Figure 2,6(a) shows a geometrical way of identifying successive values of the coordinate for a particular value of 1 in this range. i) thick curve is described by the equation for our map y= 4j.x(1 —x). We with a particular value xo in the x axis and identify the point to whi mapped by drawing a vertical line at x = xo and finding the point of jn fn with the thick curve. The diagonal dashed line corresponds to xan hedce a line parallel to x axis from the point of intersection will cut thi Hine ate that corresponds to xj =4y.x9(1 — x9). The can be now repeated from x) to find x, etc. Itis clear that the intersei ie dashed line and the thick curve will correspond to a fixed point. Th n proceeds towards the fixed point xy) and winds around it. ‘When the value of «is beyond 3/4, a new feature in the evolution i se from the original ep, two values, x2— and lutions of the equations 3 solution is given by [see Fig. 2.6(b)]. A new solution with period 2 takes stable fixed point. The sequence now oscillates b i — 3)"/?]. (2.148) This behaviour persists as long as . < (1 .86, When jis increased above this value, another new solution appears that has a period of 4. This pattern of evolution goes on ad infinitum; there is a series of critical values 14 =0.5, 0.86, ..., such that as 4. crosses these values the period of the oscil- lation doubles. 82 2 Dynamics (a) 4=0.7 (b) 4=0.775, (0) w= 1. Finally, letus consider the maximum allowed value: quence from a generic starting value will lead to a complex, a shown in Fig, 2.6(c). The actual terms in the sequence depend on the starting value, but the overall nature of the evolution remains, ie. This example illustrates that very complex evolution c: ibited by a seemingly simple system. It also shows that period doubli ical values of the parameter can lead to chaotic behaviour eventually. se 1e dynamical system considered in Eq, (2.146) (the reduced three-| iblem) is at least as complicated as this example, it is understandable that juence of points in ‘the Poincaré section shows complex behaviour is gensitive to the value of the input parameter. Exercise 215 Loren: equations: One of the early discover 108 was in a model for atmospheric convection that was governed by three v: .«, y, and z evolving according to the equations & = 10(y ~ x), §=28r —xz ‘£= xy — (82/3). Integrate these equa- tions numerically for different initial conditions and show that x, y, and z follow nonre- peating orbits in the phase space that can be chaotic. 3 Special Relativity, Electrodynamics, and Optics 3.1 Introduction This chapter begins with a rapid overview of concepts from special relativity and develops the four-vector notation. Several aspects of electrodynamics are then introduced using special relativity and four-vector notation. Finally, this formalism is used to discuss some aspects of principles of optics that are relevant to astronomy. The concept of distribution functions, developed in Section 3.6, will be used extensively in several later chapters. This chapter will also be needed, in the development of radiative processes (Chaps. 4 and 6), general relativit (Chap. 11), and in the study of cosmology in Vol. II. 3.2 The Principles of Special Relativity = ‘The description of physical processes requires the specification psp and temporal coordinates of events that may be combined into a singleaui'charac- terised by the four numbers x' = (¢, x). Throughout this chapte in indices 4b, +4, j, efe.,run over 0, 1, 2, and 3, with the 0 index dy he “fourth” dimension and 1, 2, and 3 denoting the standard space diembisions? The actual dinate system which is used. We pay special attention tOvesilbset of all possible coordinate systems called inertial coordinate systems, Sugh coordinate systems are defined by the property that a material particle, influences, will move with uniform velocity in, convenient and useful but is inherently flaws verify this criterion. In fact, there is no of coordinate systems should be preferged.over others except for mathemati cal convenience. In the development I relativity in Chap. 11 we shall drop this restrictive assumption and develop'the physical principles, treating all coordinate systems as equivalent. For the purpose of this chapter, however, we shall postulate the existence of inertial coordinate systems that enjoy a special values of x', attributed to any given event, will sy specific coor- ‘we can never operationally al reason why any one class 83 84 3 Special Relativity, Electrodynamics, and Optics status. It is obvious from the definition that any coordinate frame moving with uniform velocity with respect to an inertial frame will also constitute an inertial frame, Experiments then show that the following two statements are true: (1) All Jaws of nature remain identical in all inertial frames of reference, that is, the equations expressing the laws of nature are invariant in form with respect to the coordinate transformation connecting any two inertial frames. (2) Interaction between material particles does not take place instantaneously and there exists a ‘maximum possible speed of propagation for interactions. We denote this speed by the letter c and shall show later in Section (3.12) that ordinary light waves, described by Maxwell's equations, propagate at this speed. Anticipating this result we may talk of light rays propagating in straight lines with the speed c From statement (1), it follows that the maximum velocity of propagation c should be the same in all inertial frames. This fact has profound consequences. To begin with, this result rules out any absolute nature for simultaneity; two events that appear to occur at the same time in one inertial frame will not, in general, appear to occur at the same time in another inertial frame. Consider two inertial frames K and K’, with K" moving relative to K along the x axis. Let A, B, and C be three points along the common x axis, with AB = AC in the prim frame K’. Two light signals start from point A and go towards B and C and reach B and C at the same instant of time as measured in K’. But the two: Ses namely the arrival of signals at B and C, cannot be simultaneous for an o in K. This is because, in frame K,, point B moves towards the signal wf ‘moves away from the signal, but the speed of the signal is postyated Yee the same in both the frames. Sco The second consequence of the constancy of speed of light lowing: Consider two events P and Q with coordinates x! and x! fe define a quantity ds — called the space-time interval — between thes TW events by the relation ds = edt oak GBD Ifds =0 in one frame, it follows that these two infinksisphlly separated events P and Q can be connected by a light signal. Because light travels with the same speed c in all inertial frames, ds’ =0 in afg@Sther inertial frame. Treating (ds) as a function of (ds’)’, we can expan ‘a Taylor series in ds”, as ds? =a +ads" + ---. The fact that ds = ds’ =0 implies that a =0; the coefficient a can be a function of the relate etocity V only between the frames. Further, homogeneity and isotropy of space requires that only the magnitude |V| enter into this function. Thus ide that ds? =a(V) ds”, where the coefficient a(V) can depend on the absolute value of the relative velocity between the inertial frames. Now consider three inertial frames K, Ky, K2, where Ky and K, have relative velocities V; and V2, respectively, with respect to K. It is easy to see that a(V2)/a(V1)=a(Viz), where Vip is the relative velocity of Ky with 3.3 Transformation of Coordinates and Velocities 85 respect to K2. This is, however, impossible because the relative velocity Viz depends not only on the magnitudes of V; and V2 but also on the angle between the velocity veetors. Hence the function a(V) must be a constant; further, this constant should be equal to unity to satisfy the above relation. It follows that the quantity ds has the same value in all inertial frames; ds? = ds", ie., the space-time interval is an invariant quantity. Events for which ds*is greater than, equal to, or less than zero are said to be separated by timelike, null, or spacelike intervals, respectively. To understand the physical meaning of ds, consider a clock that is moving relative to an inertial frame K in an arbitrary trajectory. During a time interval dr, the clock moves through a distance {dx as measured in K- In another inertial coordinate system, K’, which is moving with respect to K with the same velocity as the clock at that instant of time f, the clock is momentarily at rest, giving dx’ =0. Ifthe clock indicates a lapse of time dt’ when the time interval measured in K is dt, the invariance of space-time intervals shows that ds? = 7 dt? — dx? — dy? — dz? = ds? = c*at™ (3.2) or, Hence ds/c may be thought of (for timelike intervals) as the lapse of me le moving clock; this is usually called the proper time along the trajegtor clock. ey In defining ds? we have used a positive sign for cds? and a ne} for the spatial terms dx”, etc. The sequence of signs is called si id it is usual to say that the signature of space-time is (+ — — —). Weiains¥quivalently, use the signature (— + ++), which will require a chang in several expressions. We should keep this point in mind while formulas in different textbooks. 3.3 Transformation of Coordinat Velocities Because the concept of simultaneity has no invs of any two inertial frames will be related by a: time coordinates will, in general, be differ we study the general question of determi arbitrary observer moving along the x Let(t, x, y, 2) bean inertial coordinate system. Consideran observer travelling along the x axis in a trajectory x = f(r), f=h(r), where f and h are specified functions and r is the proper time in the clock carried by the observer. We would like to assign a suitable coordinate system to this observer. Let P be some event ificance, the coordinates formation in which space and future applications in mind, the coordinates appropriate for an 86 3 Special Relativity, Electrodynamics, and Optics with inertial coordinates (t, x). The observer sends a light signal from the event A (at r= 14) to the event P. The signal is reflected at P and reaches back the observer at event B (at r = rg). Because the light has travelled for atime interval (tp — 14), it is reasonable to attribute the coordinates 1 1 p(tetta, x= 5 (te ta)e G4) to the event P. To relate (1’,x') to (f, x) we proceed as follows: Because the events P(t, x), Alta, xa) and B(tg, xg) are connected by light signals travelling in forward and backward directions, it follows that c(t — te) G.5) (20m xtet = xp-tete = flts) + ch(te) = (r+ =) sen( + “) & Given f and h, these equations can be solved to find (x, 1) in terms of This procedure is applicable to any observer. For an observer moving with uniform velocity V, the trajectory 3) a with the proper time given by r =s[1 — (V2/c*)}'? [see Eq¢{3.3)] 580 the parameterised trajectory is, ° «koe te > — x4 = c(t = ta), x= Xa ~ 6ta = f(ta)— ch(ta) alk v = 3.8 [= FH G8) giving pa onge raans(va2) aei(4%) 1 Vv | (vr toe’) + (cr et’. mores) G9) Solving these equations, we get the tranformation between the inertial frames, called the Lorentz transformation (LT} xayQl4Ve), y= ' ("+ Be). rsa? G.10) 3.3. Transformation of Coordinates and Velocities 87 This transformation leaves the quantity s? = cs? — |x|? invariant, as this is the space-time interval between the origin and any event (¢, x). [This fact can also be directly verified from Eqs. (3.10),] A quadratic expression of this form is similar to the length of a vector in three dimensions that — as is well known — is invariant under rotation of the coordinate axes. This suggests that the transforma- tion between inertial frames can be thought of as a rotation in four-dimensional space. For two inertial frames K and K’ with a relative velocity V, we can always align the coordinates in such a way that the relative velocity vector is along the common (x, x’) axis. Then, from symmetry, it follows that the transverse direc- tions are not affected and y’ = and 2’ =z. The rotation must be in the rx plane characterised by a parameter, say y. The LT in Eqs. (3.10) can be written as Jeoshy tet’ sinhy, ct =x'sinhy +et/coshy, (3.11) where tanh yr = V/c, which determines the parameter yy in terms of the relative velocity between the two frames. In general, the LT mixes up space and time coordinates and reduces to the stan- dard Galelean transformations of nonrelativistic mech y’=y,2'=2) in the limit of e— 00, Exercise 3, Lorentz transformations in arbitrary directions: Consider a LT along a directi cated by the unit vector fi. (1) Show that the transformation of coordinates: in the matrix form x" = A',x/, where the components of the matrix A 1- py, Ala Ag=—ypn", pay ata 48 ss ea Verify tat the Jacobian of this transformation is unity. (2) the inverse ofthe matrix A is obtained when f is changed to ~B. ° Exercise 3.2 Four vectors and matrices: To any event x! =.x'o}, where ap is the identity matrix and 0, the direction i with speed V, the event x! goes, 1, do not change.) Show that P’= LPL", and tanh a =(V/o). associate a 2 x 2 matrix uli matrices. Under a LT along. P goes to P’. (By convention = cosh(a/2)+(n ) sinh(ar/2) Given the LT, we can compute the transformation law for any other physical ‘quantity that depends on the coordinates. As an example, consider the transfor- mation of the velocities between two inertial frames: Taking the differential form. 88 3 Special Relativity, Electrodynamics, and Optics of the LT, Vv dx =y(dx'+Vdt), dy=dy’, dz=d?, dt= (a + Se). 14) x! /dt', we find the transformation law and forming the ratios y=dx/dt, v’ for the velocities to be vetV oY _ vw T+ (%,V/e)" eve "OY Te wv G.15) Up From this result, itis easy to verify the following facts: (1) The transformation law reduces to the familiar addition of velocities in the limit of e+ 00. But in the relativistic case, none of the velocity components can exceed c. (2) The addition law for the x component of the velocity is equivalent to the additivity in yy = tanh™'(V/c), viz., yo = Wi + Yo for two successive rota- tions in the rx plane. Even the transverse velocities are, however, affected by, motion along the x axis; this is quite unlike the situation in nonrelativis physics. (3) The velocities v’ and V combine in an asymmetrical way except in the, case in which they are both along the same direction. This is a of the fact that the net effect of two Lorentz transformations (noncollinear) axes depends on the order in which the trangformat performed. (4) The transformation of velocities shows that the direction o ticle will appear to be different in and vy =v sin are the components in the coordinate (with primes denoting corresponding quantities in frame K’), the Sy to see from Eqs. (3.15) that — wane © G16) cos EV tan@ = For a particle moving with relativistic velocitigeQXe*e) and for a ray of light, this formula reduces to @.17) We shall see later (in Section 3.12.1), in radiative processes. formula has important applications Exercise 33 Superluminal motion: Consider a blob of plasma moving with at a speed v along a direction that makes an angle y- with respect to the line of sight. (I) Show that the 3.4 Four Vectors 89 apparent transverse speed of the source will be related tothe actual speed by vsiny oarreg From this, conclude that the apparent speed can exceed the speed of light. How does vagp ‘vary with y fora constant value of v? (2) Ifthe blob of plasma is ejected from a source ‘with equal probability in all directions in the rest frame of the observer, show that the probability that the observer will see the blob move with an apparent velocity greater than Ugg is given by G.18) 1 JT = 7) Exercise 3.4 Thomas precession: (1) Let A(v) denote the matrix corresponding to a Lorentz boost along the direction ¥ with speed v. Show that A(vi)A(v:)= R(@A)A(Ws), where R(OA) is the spatial rotation matrix for rotation by an angle @ about the direction a. This shows that the effect of two LTS is (in general) the same as that of an LT and a rotation. Find 9, 8, and vs in terms of v; and v2. (2) Consider a particle undergoing acceleration comoving in the comoving, instantaneous rest frame. Because of the acceleration, the comoving frames forthe particle at two different instants of time ¢ and (¢ + df) will be different. In the lab frame, the combination of two boosts with velocities v and (v +8¥) ‘will not be a pure boost but willbe a combination of pure boost and rotation by an an ‘Use the results of (1) to show that sone =e ian, r= (son?) (nae Fn ee Innit et sein Gly N= (1/2c(a x y). (Hint; Use the result of Exercise 3.2.) Light echoes: A large, planar sheet of dust exists between an of radiation with the plane inclined at an angle i to the line of, source is scattered by dust in this plane towards the observer{ Gupppse the source flares up in intensity at = 0 for some reason. Show that the observer see, at time f > 0, a circle of scattered light with the radius o 3.4 Four Vectors Equations like F=ma, which are written in vector notation, remain valid in any three-dimensional coordinate system without change of form. For example, consider two Cartesian coordinate systems with the same origin and the axes 90 3 Special Relativity, Electrodynamics, and Optics rotated with respect to each other. The components of the vectors F and a will be different in these two coordinate systems but the equality between the two sides of the equation will continue to hold. Ifthe laws of physics are to be expressed in a form that remains invariant under LT, we should similarly construct vectorial quantities with four components and treat LTs as rotations in a four-dimensional space. Such vectors are called four vectors and will have one time component and three spatial components. The spatial components, of course, will form ordinary three vectors and transform as such under spatial rotations with the time component remaining unchanged Let us denote a generic four vector as A‘ with components (A°, A) in some inertial frame K. The simplest example of such a four vector is the space-time coordinates of an event x' = (ct, x). (We have used ct rather than ¢ in order to give the same dimensions to all four components.) Knowing the transformation properties of the four vector x‘, we define the transformation law for a general four vector to be Abs y(ae+ tar), Al =r(a" + Yar), a 2 Boas, (3.22) Itis obvious from our construction that, under these transformations, the sq ofthe length ofthe vector defined by (A®)? ~ JAP? remains invariant. =< It is convenient to introduce at this stage two different types of comple) of four vectors denoted by A‘ and Aj, with A! =(A®, A) and 4; =(AY2A)-In ‘other words, lowering of the index changes the sign of the spatig] comhon Given this definition, we can write the squared length of the vector as A'al, with the understanding that any index that is repeated in an expres summed ‘over, that is, AA; stands for the expression 3 ANA =)” A‘A) = AAg + ANA, + APAY Ga (3.23) i == This quantity need not be positive definite. A four were led timelike, null, or spacelike depending on whether this quantity is posite, zero, or negative, respectively More generally, given two four vectors Aé define a dot product between them by a rule. AB; = APBy + A'B, + A? The dot product is invariant under LE&ean be easily verified. The squared length of the vector, of course, is j {ot product of the vector with itself. To illustrate the above formalism, consider two examples of four vectors. Let x'(s) denote the space-time coordinates of a particle when the proper time carried by a clock moving with the particle is The four velocity of the particle can be defined to be u! =dx' /ds. Because ds? = dx'dx; is a scalar, it follows ea ) and Bi = (B°, B), we can to that for A'B,, = A°B°-A-B. (3.24) 3.4 Four Vectors 91 that w! transforms as a four vector. The components of this vector are y 7 (» y ) 3.25) e From this definition it follows that uu; =dx'dx;/ds? = locity has only three independent components. Ina similar manner we can define the four acceleration to be a‘ =d?x‘ /ds? = du’ /ds. Differentiating the relation u'u; = 1 with respect to s, we easily see that a'u;=0, Just as the concept of the three vector has been generalised to that of a four vector, we can also generalise the gradient operator to four dimensions. The ordinary three-dimensional gradient, V = [(8/8x), (8/8y), (2/92)], transforms as a vector under spatial rotations. To define the four-dimensional gradient, we have to only note that the differential of a scalar quantity d@ = (36/dx')dx! is also a scalar. Because this expression is a dot product between dx‘ and (39/83), it follows thatthe latter quantity transforms like a four vector under LT. Explicitly, the components of the four gradient of a scalar are given by oer (t 20 ve 6.26) ax = (cor As an example, consider the notion of a normal to the surface. A tht dimensional surface in four-dimensional space is given by an equation oflte)) form f(x')=0. The normal vector nj(x*) at any event x* on this surffce- given by n =(Af /Ax"), which is an example of a four gradient. Iris c tional to call a surface spacelike, null, or timelike at x°, wet ether so that the four ve- ‘n; is timelike, null, or spacelike at x*. In a similar manner, the fo) ence of a four vector is defined to be (3A' /@x') with summation over Gauss’s theorem in three dimensions is generalised to four. oe as aat fore at @E (n"Ag), (3.27) ly ax Soy where V isa region of four-dimensional space bounded bytatilee surface AV and 4°E is an element of a three surface with a normal isa volume integral and the right-hand side is a sur of a four-dimensional region V can be conver following components: (1) Two three-dimen: both of which are spacelike; the coordina spatial coordinates (r, 6, 6). (2) One timekik&eurface at a large spatial distance (r=R= oo) at any time satisfying f; the coordinates on this three- dimensional surface could be (t, 8, @). On thie right-hand side of Eq, (3.27) the integral has to be taken over the surfaces in components (1) and (2). Ifthe vector field Aq vanishes at large spatial distances, then the integral over the surface in component (2) vanishes for R—> oo. For the integral over the surfaces in torg®. The left-hand side itegral. The boundaries taken to be made of the surfaces at ¢=t) and t=, jese surfaces are the regular 92 3 Special Relativity, Electrodynamics, and Optics component (1), the normal vector n® has components (1, 0, 0,0). It follows that [ase =[exao-f, Bx Ao, 6.28) ax! with the minus sign arising from the fact that the normal must always be treated as, outwardly directed. Itfollows that if(9.A‘/ax') = 0, then the integral of Ao( = A") over all space is conserved in time. We can obtain the same result, of course, by writing the equation (@'/x') = 0 in the form [9° /a(ct)] + V-A=0 and integrating the terms over all space. ‘At the next level of complication, we can define four tensors as quantities that transform like the product of four vectors. For example, consider an entity Cix defined to be Cx = A;Be, where A; and By are four vectors. Knowing the trans- formation law for the four vectors, we can easily find out how the components of Cix get mixed under a LT. A second-rank tensor Tix is defined to be a set of 4x 4= 16 quantities that transform like the product A; By of two four vectors under a LT. Because we have defined two types of components for four vectors, ‘we can have second-rank tensors like T"*, 7j, or Tix differing in the position of indices. Whenever an index corresponding to a spatial coordinate is rai ‘or lowered, the sign of the component changes. Higher-rank tensors, with my indices, can be defined along similar lines. ‘A tensor is symmetric (antisymmetric) on two indices (a, b) if interchg the indices leaves the value of the component same (changes the component). Any tensor can be written as the sum of a symmetric symmetric part with respect to any pair of indices. For example, (7 — T#)/2is antisymmetric and S! = (7 + where A! ric in i,k). and A‘ =n'* Ag, (We stress the convention that indiges.w whch are repeated are summed over in such expressions.) It can be trivially teats that (1) this relation reproduces our original definition of the components, a4(2) the components /* and mx have the same numerical value in all inertial ames (see Exercise 3.6). The raising and the lowering of tensor indic the obvious generalisation ofthe above rule, ¢.g., Tj = na T'4, etc. jduct between two vectors can now be expressed as jx A’ B* and the not rector by nix 4; A*; in particular, ds* =nixdxidx*. These concepts and will be of tremendous use when we develop general relativity in Chaps Exercise 3.6 Index gymnastics: (1) Show that ny and n* have the same numerical value forthe com- ponents in all inertial frames. (2) The four-index tensor cis defined to be completely 3.5. Particle Dynamics 93 antisymmetric in any pair of indices with ¢oy23 = 1. Show that this tensor also has the same components in all coordinate frames. We can also define a thre-index object €apy (here the Greek subscript take the values 1,2,3) long similar lines. tis antisymmetric in any pair of indices and ¢123 = 1. Show that, in three-dimensional space, éqyy A? BY gives the ath component of A x B. (3) Write down the transformation law for a two- index object A under LTs, i. express the components A" explicitly in terms of Ai* How does it simplify if A = —A¥? (4) Show that A’M#™---Sjap++ =O if AM is antisymmetric in the pair (i,k) and Six, is symmetric in the pair (i,k). This result ill be used extensively in future discussions 3.5 Particle Dynamics ‘To determine the laws governing the motion of a free particle, we need an expres- sion for the action that can be varied. This action should be constructed from the ‘trajectory 2 (s) of the particle and should be invariant under LTs. The only pos- sibility is some quantity proportional to the integral of ds, so the action must be a 2 we A -a fia -f acyl Sat, wu where is a constant, In arriving at the second equality, we have expressed ds in terms of dr by using Eq. (3.3), which shows that the Lagrangian is by L=dA/dt =—acy/i— v2/e2. When c—> oo, this Lagrangian redu L=av?/2c+ constant. Comparing this with the Lagrangian (1/2)mvt(for free particle in nonrelativistic mechanics, we find that a = mc, whet Az ‘where the second equation identifies the Lagrangian wigs he Ive. This action in relativistic mechanics has a clear geomettic-}eaning, unlike its nonrelativistic counterpart. To determine the dynamics, we vary the action xi(s) and get o spect to the trajectory sAeome fas =me » mame f ujdbx! If we now assume that 5x! vanishes at the end points, we obtain the equation of motion du! /ds =0, which is a generalisation of force-free motion to relativistic as, G31) 94 3 Special Relativity, Electrodynamics, and Optics ‘mechanics. Further, by treating the action as a function of the end points of the trajectory that satisfies the equation of motion, we obtain (3.A/Ax') = —mew; Because the derivatives of the action with respect to the coordinates define the ‘momenta, the four-momentum vector is given by aA E Fo mows = (ye, yy) (=. -r). 32) E vy tome, ymo)= (sp), y (1-3) 3.33) To obtain the physical significance of the time component, E = ymc?, we note that, in the nonrelativistic limit, E mc? + mv? /2. This suggests that E corre- sponds to the relativistic energy of the particle. Such an identification is further vindicated by the fact that, for a Lagrangian L=—me?/y, the Hamiltonian 1H=p-v—L is numerically the same as E. We thus conclude that the three mo- ‘mentum p = ymv and energy (divided by c) form the components ofa four vector. Because wu; = 1, it follows that p! p; =mc?, giving the following relations connecting momentum, energy, and velocity: E=Vpetmc, p #(3)- 0h The first relation allows the existence of massless particles with m —> 0, ‘meu and v=(pe?/E)=c. Finally, the Hamilton-Jacobi equation for the relativistic partigle cki-bdob- —(8.A/8x;) and the condition me; we L(aAy?_ (ad aA ) a(t) -(&) -G) -(@) Oo a To obtain the nonrelativistic limit of this equation, we om = met + ‘S(x*) into Eq. (3.35). Simplification then gives a Ese + © (3.36) where the last equality arises in the limit ofc Jacobi equation for the free particle in the his is exactly the Hamilton lativistic theory. Practice with four vectors: An observer Wifour velocity w* measures the properties of a particle with four momentum p*. Express (1) the energy, (2) the magnitude ofthe three momentum, and (3) the magnitude of the three velocity attributed to the particle by the observer in terms of w* and p®. [Answers: (1) Consider the dot product pu, that is Lorentz invariant and has the same numerical value in all frames. We evaluate it in 3.6 Distribution Functions and Moments 95 the rest frame of the observer in which i, = (1, 0, 0, 0). In this frame pu. = p°=E/c, where E is the energy of the particle, Because p*u., is Lorentz invariant, it follows that the energy is E = cp*u,, evaluated in any frame. The rest follows easily from the relation pe= JEP = mick and v= (pe?/E)) 3.6 Distribution Functions and Moments So far, we have discussed the dynamics of a single particle. Often in astrophysics, we need to deal with a large collection of particles undergoing nearly identical physical processes. In nonrelativistic mechanics, we deal with this situation by uusinga distribution function. Itis necessary to generalise this concept ina Lorentz invariant manner to take into account a system of relativistic particles. To do that, we first obtain several relativistically invariant quantities that will serve as basic building blocks. To begin with, an element of four-dimensional volume, d*x = cdtdxdydz is invariant under LTS. Although this is obvious from the definition of LTS as rotations in space-time, it can also be explicitly proved by checking that the Jacobian of a LT is unity (see Exercise 3.1). Let us next consider a set of N particles, each of mass m, described by a distribution function f(p') at any given location in space. The total number GS particles can be written in terms of the distribution function as Nm f atv on soln" mes, where dp =dp°d3p; the Dirac delta function 8p(p*p. — mc?) all the particles have mass m, and the theta function @(p) (whet p® > O and vanishes for p? <0) ensures that p" > 0, The quantit and 3p(p" pa —m?c?) are all individually Lorentz invariant, i Lorentz invariant. Writing the Dirac delta function as ER c Sp(pip! — mc?) sole = 2) ale (" a where £2 = m?c*+ pc? and noting that integratios p° with (Ep/c) because of the condition p? > =f épap’ap) o_ wa fa pap a(p' see [80( 0! Hf [EP poe =] Gp 0" Foie. 639) Because N and f are invariant, the combination (d° p/Ep) must be invariant under Lorentz transformations 96 3 Special Relativity, Electrodynamics, and Optics Finally, from the relations Ey = ymc?, (ds/cdt) = y~', we find that Ep(ds/dt) is Lorentz invariant. Multiplying by d°x in the numerator and denominator, we get ds dx ds Px (5). (3.40) showing that the combination Eyd°x is also invariant, Combined with the earlier result that @p/Ep is Lorentz invariant, we conclude that the product (Epd°x) (@°p/E»)=4°xd'p is Lorentz invariant. In other words, an element of phase volume is relativistically invariant, even though neither the spatial volume nor the volume in momentum space is individually invariant. This result allows us to introduce distribution functions in relativistic theory in exact analogy with nonrelativistic mechanics. We define the distribution function f such that AN = f(x',pd?xd3p, G41) represents the number of particles in a small phase volume dxd°p. The x! here stands for the four-vector components (t, x),and pis the three-momentum vector; the fourth component of the momentum vector (Ep/c) does not appear as completely determined by p and mass m of the particle. Each of the quanti AN, f, and d°xd'p are individually Lorentz invariant. Given the Lorentz invariant distribution function f, we can construct other invariant quantities by taking moments of this function, Of part ( portance are the moments constructed by integration of the distribution funétion over various powers of the four momentum. We now construct a fetv examples. ‘The simplest Lorentz invariant quantity that we can obtain fromheWesribution function by integrating out the momentum is the harmonic r4an)Par of the energy ofthe particles at an event x'. Ths is defined by the 1 : f G42 Fad) which is clearly Lorentz invariant because: otourpee inet Unfortunately, this quantity does not seem to play any important¥ple.n physics. define the four vector (3.43) = | Ppsei w= (3.44) 1 Jee’) feere!.pw 3.6 Distribution Functions and Moments 97 where we have used the relation (p*/E) = (v /c?). The time component of this vector, J°, gives the particle number density n in a given frame; the spatial components give the flux of the particles in each direction. The factor ¢ was introduced in definition (3.43) to facilitate this interpretation. ‘Taking quadratic moments allows us to define the quantity ’ raac f Fre p. G45) called the energy-momentum tensor or stress tensor of the system. This tensor is clearly symmetric. When one of the indices is zero, we get ® TH!) = Teme [FE (Epo? sew =e favo" s0'.p), 6.46) which is (c times) the total four momentum of the particles per unit volume. ‘The time-time component T(x) gives the energy density and the time-space component T(x‘) gives the density of the « component of the three momentum. ‘The space-space components of this tensor represent the stresses within the medium. The component T°? gives the « component of the momentum that, crosses unit area orthogonally to the f direction per unit time. Given a distribution function, we can construct the four vector J%(x) at given event through definition (3.43). It is also possible, always, to ch Lorentz frame such that the spatial components of this vector vanish th event (ie., (v) =0) so that an observer with that Lorentz. frame, dBes any mean flux of particles around a given event. If the gradient velocity (v) is sufficiently small, then such a Lorentz frame can bd globally for the whole system. (Such a definition is approximaje= when physical processes that depend on the gradient of mest Kinetic energy, etc., are ignored. We shall discuss in Chap.'8 contributions that arise from the gradients.) Let us suppose in such a Lorentz frame and also that the distribution ou is isotropic in ‘momentum in this frame; that is, it depends only gn the=magnitude p ofthe ‘momentum p. In such a frame, fers wasn f° v ;, 1 = fapese'm=ax [rt \ ‘As regards the space-space part of the energy-momentum tensor, it has to be an isotropic, symmetric, three-dimensional tensor. Hence 7 must have the form Tg = P(x')53, as 6f is the only tensor available that satisfies these conditions. G47) (3.48) 98 3 Special Relativity, Electrodynamics, and Optics To find an expression for P(x'), note that Ty = PG!) = seaiae [ope LO, p) = ane en pe +P). (3.49) Hence, nc? pt = —e 3.5 roa If * ange sftp) 6.50) This quantity represents the pressure of the fluid and has simple limits in two ex- treme cases. In the nonrelativistic limit, the energy of the particle is E(p) me? + (p?/2m), Substituting this form for E(p) in the expression for T®, we find that the energy density can be written as 7% = mc2n + éqe, where the nonrelativistic contribution ey to the kinetic energy is oo, gt 7 2" a2 4 a= ar [ PE s@ap=2 f P'fpdp. G1) In the same limit, expression (3.50) for pressure reduces to o anc? an Pg 3 AE se [an r0= [ot sevae Comparing expressions (3.51) and (3.52), we see that éqr=(3/2)Par S the relation between energy density and pressure in nonrelativistf&-gheo} that pressure has nothing to do with interparticle collisions a: it arises from momentum transfer across a surface. In the other extre “of highly relativistic particles, E(p) = pe. In this case, 4 4 a a ln which shows that, for extreme relativistic particles, ‘om and the energy density are related by Qio. (3.53) (3.54) In particular, this equation is exact for, s with zero mass, for which E(p)= pc is an exact relation. Given the components of the stress @abr in the special frame in which bulk. flow vanishes, it is easy to obtain ts in any other frame in which the observer has a four velocity w. The result, obtained by a LT, is Ti = (P+ p)utuy — PS, J? = Mpopt. G55) Here prop is the proper number density — ie, the number density in the frame 3.7 External Fields of Force 9 ‘comoving with the particles — and is a scalar, itis related to m in Eqs. (3.44) by N= YNxop- To avoid misunderstanding, the following fact is stressed. The above form of T% jis valid only when physical processes that arise because of gradients in u”, etc., are ignored so that a global Lorentz frame could be defined. The spatial dependence of quantities like P(x'), etc., should be interpreted with this caveat in mind. When the gradients are not ignorable, there will be additional terms to T* that we will discuss in Chap. 8, Section 8.4. Exercise 3.8 Filling in blanks: Prove the result of Eqs. (3.55). 3.7 External Fields of Force In nonrelativistic mechanics, we can introduce the effect of an external force field by adding to the Lagrangian the term —V(¢, x), thereby adding to the action the integral of —Vdr. Such a modification is, however, not Lorentz invariant and hence cannot be used in the relativistic theory. To find the kind of interactions that are permitted by Lorentz invariance, we may proceed as follows. ‘The action for the free particle is the integral of ds, which is Lorentz invari We can modify this expression to the form C) A= ome f Costu" ds, Ss a where £(x*, u*)is a Lorentz invariant scalar made of the position at Sacein ofthe particle, and still maintain Lorentz invariance. We obtain a choice for £(x*, u*) by taking the polynomial in w* as La 1+ cob(x) + crAiladut + ergan(xutu’ A. 3.57) where $(x) isa scalar, Ai(x) isa four vector, gap(t) isa ink tensor, ete, and co, €2,..., ete., are constants introduced for later lience. (Quantities, like @ depend on the four vector x" but for conver (x) instead of o(x").] In this expansion, 4, Ai, 8a fields that influence the trajectory of the particl It turns out that, in nature, we come actos electromagnetism and a second-rank tens other words, the Taylor series expansion the quadratic term and no higher-deg arise. Further, the scalar field can be included in the term with gai by thé wddition of a part #(x)nap- Thus we need to deal with only A;(x) and gaa(x) as externally imposed fields of force that influence the trajectory of a particle. The study of gap(x) (which describes the gravitational field) is postponed to Chap. 11, and we concentrate on Aj(x) nc9,of notation we write are externally specified a vector field Aj describing Fab(x) describing gravity; in in the variable u® terminates after 100 3 Special Relativity, Electrodynamics, and Optics here, The action for a particle influenced by such a field is given by » a=[ (-meds — £4idx'), 3.58) where we have written the constant c) as —q /cand used the fact that u' = dx! /ds. The quantity q is called the electric charge of the particle, which is a Lorentz invariant number (like m) that characterises the particle. The corresponding Lagrangian is L=—mey1— (ee) — qo + ta-v (3.59) if A’ =(, A) and A; = ($, -A). To find the equations of motion in this particular case, we vary the action of Eq. (3.58) with respect to the trajectory x'(s) and get 3A » i -{ (metas + Laiasx! + fsaax') =0. 3.60) la ds c c By integrating the first two terms by parts and using the relations aA; aA; ‘ k sai = gor, dai = ast, ee f [nest + Uaparutdx! — LapAau'sx! & I, as c c we find that (3.62) \dices i and k Oe f (mee - (3.63) with the definition (3.64) ind points lead to the equations of ‘motion: (3.65) dsc The variation of the action as a function of the end points, with the trajectories 3.7 External Fields of Force 101 satisfying the equations of motion, allows us to identify the canonical momenta as ad 4 4

You might also like