a first book of quantum field theory - amitabha lahiri and palash b. pal

393

Click here to load reader

Upload: arindam-kumar-chatterjee

Post on 22-Jan-2018

454 views

Category:

Science


18 download

TRANSCRIPT

  1. 1. Preface to the second edition There has not been any drastic changes in the present edition. Of course some of the typographical and other smaH mistakes, posted on the internet page for the book, have been corrected. In addition, we have added OCC8S- sional comments and clarifications at some places, hoping that they would make the discussions more accessible to a beginner. The only serious departure from the first edition has been in the no- tation for spinor solutions of the Dirac equation. In the previous edition, we put a subscript on the spinors which corresponded to their helicities. In this edition, we change the notation'towards what is more conventionaL This notation has been introduced in 4.3 and explained in detail in 4.6. We have received many requests for posting the answers to the exercises. In this edition, we have added answers to some exercises which actually require an answer. The web page for the errata of the book is now at http://tnp.saha.ernet.in/~pbpal/books/qft/errata.html If you find any mistake in the presen~ edition, please inform us through this web page. It is a pleasure to thank everyone who has contributed to this book by pointing out errors or asking for clarifications. On the errata page we have acknowledged all correspondence which have affected some change. In particular, comments by C. S. Aulakh, Martin Einhorn and Scan Murray were most helpful. Amitabha Lahiri Palash B. Pal July 2004
  2. 2. Preface to the first edition It has been known for more than fifty years that Quantum Field Theory is necessary for describing precision experiments involving electromagnetic interactions. Within the last few decades of the twentieth century it has also become clear that the weak and the strong interactions are well described by interacting quantum fields. Although it is quite possible that at even smaller length scales some other kind of theory may be operative, it is clear today that quantum fields provides the appropriate framework to describe a wide class of phenomena in the energy range covered by all experiments to date. In this book, we wanted to introduce the subject as a beautiful but essentially simple piece of machinery with a wide range of applications. This book is meant as a textbook for advanced undergraduate or beginning post-graduate students. For this reason, we employ canonical quantization throughout the book. The name of the book is an echo of various chil- dren's texts that were popular a long time ago, not a claim to primacy or originality. Our approach differs from many otherwise excellent textbooks at the introductory level which set up the description of electrons and photons as their goal. For example, decays are rarely discussed, since the electron and the photon are both stable particles. Howeverl decay processes are in some sense simpler than scattering processes, since the former has only one particle in the initial state whereas the latter has two. We felt that the basic machinery of Feynman diagrams could be intro- duced through decay processes even before talking about the quantization of spin-l fields. With that in mind, we start with some introductory ma- terial in Chapters 1 and 2 and discuss the quantization of scalar fields in Chapter 3 and ofspin-4 fields in Chapter 4. Unlike many other texts at this level, we use a fermion normalization that should he applicable to massless as well as massive fermions. After that we discuss the generalities of the S-matrix theory in Chapter 5, and the methods of calculating Feynman diagrams, decay rates, scattering cross sections etc. with spin-O and spin-! fields, in Chapters 6 and 7. The quantization of the spin-l fields, with special reference to the pho- vii
  3. 3. Vlll Preface ton, is taken up next in Chapter 8. After this, we have a detailed intro- duction to quantum electrodynamics in Chapter 9, where we introduce the crucial concept of gauge invariance and give detailed derivations of impor- tant scattering processes at their lowest orders in perturbation theory. Discrete symmetries can serve as a good guide in calculating higher or- der corrections. With this in mind, we discuss parity, time reversal, charge conjugation, and their combinations, in Chapter ]O. Unlike most other textbooks on the subject which describe the P, T, C transformations only in the Pauli-Dirac representation of the Dirac matrices, we present them in a completely representation-independent way. Since other representations such as the Majorana or the chiral representations are very useful in some contexts, we hope that the general formulation will be of use to students and researchers. We calculate loop diagrams in Chapter 11, showing how to use symme- tries of a problem to parametrize quantum corrections. In this chapter, we restrict ourselves to finite contributions only. Some basic concepts of renor- malization are then described, with detailed calculations, in Chapter 12. Only the electromagnetic gauge symmetry is used in these two chapters for illustrative purpose. But this leads to a more general discussion of symme- tries, which is done in Chapter 13. This chapter also discusses the general ideas of symmetry breaking, and the related physics of Nambu-Goldstone theorem as well as its evasion through the Higgs mechanism. This is fol- lowed by an introduction to the Yang-Mills (or non-Abelian) gauge theories in Chapter 14. Finally, a basic introduction to standard electroweak the- ory and electroweak processes is given in Chapter 15. In keeping with our overall viewpoint, we discuss decays as well as scattering processes in this chapter. We have tried to keep the book at the elementary level. In other words, this is a book for someone with no prior knowledge of the subject, and only .a reasonable familiarity with special relativity and quantum mechanics. To set the stage as well as to help the reader, we visit briefly the relevant parts of classical field theory in Chapter 2. A short but comprehensive introduction to group theory also appears in Chapter 13, since we did not want to assume any background in group theory for the reader. The most important tools to help the reader are the exercise problems in the book. These problems are not collected at the ends of chapters. Instead, any problem appears in the place of the text where we felt it would be most beneficial for the reader to have it worked out. Working it out at that stage should also prepare the reader for the ensuing parts of the chapter. Even if for some reason the reader does not want to work out a problem at that stage, we suggest strongly to at least read carefully the statement of the problem before proceeding further. Some of the problems come with notes or hints, some come with a relation that might be useful later. A few of the problems are marked with a * sign, implying that they
  4. 4. Preface IX might be a little hard at that stage of the book, and the reader can leave it at that point to visit it later. A few have actually been worked out later in the book, but we have not marked them. The book was submitted in a camera-ready form to the publishers. This means that we are responsible for all the mistakes in the book, including typographical ones. We have spared no effort to avoid errors, but if any has crept in, we would like to hear about it frum the reader. We have set up a web site at http://tnp. saha. ernet. in/rvpbpal/qftbk .html containing errata for the hook, and a way of contacting us. The book grew out of courses that both of us have taught at several universities and research institutes in India ami ahroad. We have ben efited from the enquiries and criticisms of OUl' students and colleagues. Indrajit Mitra went through the entire manw;cript carefully and offered numerous suggestions. Many other friends and colleagues also read parts of the manuscript and made useful comments, in particular Kaushik Bhat tacharya, Ed Copeland, A. Harindranath, H. S. Mani, Jose Nieves, Saurabh Rindani. Those who have taught and influenced liS through their lectures, books and papers are too numerous to name sepa.rately. We thank them all. We thank our respective institutes for extending various facilities while the book was being written. Amitabha Lahiri Palash B. Pal April 2000
  5. 5. Notations JJ., II, Space-time indices of a vector or tensor. i,i, Spatial indices of a vector or tensor. gpol.J (Components of) metric tensor, diag(l, -1, -1, -1). pJJ contravariant 4-vector. P,.. covariant 4-vectof. p 3-vector. p Magnitude of the 3-vector p, i.e., Ipl. We have used p2 and p2 interchangeably. The magnitude of the co-ordinate 3-vector has been denoted by r. a b Scalar product of 3-vectors a and b, a . b = aibi == I:aibi . a b Scalar product of 4-vectors a and b, a b = al-'b,.. =- L: aP-b,.. = aObo - a b. ;. "'(/Jaw !L' Lagrangian density, frequently called Lagrangian. L Total Lagrangian (; f d3 x Z). PI Action (; f dt L ; f d4~ Z) . .Ye Hamiltonian density, frequently called Hamiltonian. H Total Hamiltonian (; f d3 x'). lA, Blp Poisson bracket of A and B. lA, BI_ Commutator AB - BA. lA, BI+ Anticommutator AB + BA. a(p), at(p) Annihilation and creation operators for antiparticles of particles created by at(p) and annihilated by a(p), first encountered in 3.6. 8(x) Unit step function, defined in Eq. (3.13). :[000]: Normal ordered product, first defined in 3.4. 5' Ij Time ordered product, first defined in 3.7. xi
  6. 6. Xli Notations AT Transpose of t.he matrix A. At For any matrix A, At =::. !'oAfl'D . (Ti Pauli matrices, given in Eq. (4.55). r i Same as (Ti, hut thought of as generators of some internal SU(2) symmetry. le- (p, s)) Electron state of 3-mornentum p and spin s, first defined in 6.2. 6.F(p) Feynman propagator for scalar field in momentum space. SF(p) Feynman propagator for fermion field in momentum space. DJw{p) Feynman propagator for vector hoson field in momentum space. e Electric charge of proton. Electron carries charge -e. a Fine structure constant, first defined in 1.5. e- Electron. "'Y Photon. S/i S-matrix element between initial state Ii) and final state If}. .4lfi Feynman amplitude between initial sLate Ii} and final state If). 1_4f12 Magnitude squared of Feynman amplitude after making spin and polarization sums and averages. lit Polarization vector for a vector boson., P Parity transformation operator, defined in Ch. 10. C Charge conjugation operator. T Time reversal operator. P Matrix of parity transformation acting on a fermion. C Matrix of charge conjugation acting on a fermion. T Matrix of time reversal acting on a fermion. x Parity transformed system of co-ordinate::;, i = (t, -x). r It General electromagnetic vertex, first. defined in 11.1. E For dimen::;ional regularization in N dimensions, E = 2 - !N. dN Rank of I'-rnatrices in N-dimensional space-time, tr hpl'v) = dN 9lw ' Ta Generators of a Lie group. fabc Structure constants of a Lie group, IT", Tbl- = ifnbcTc. DI Gauge covariant derivative D,! = a/. + i9I:,A~.
  7. 7. Contents Preface to the second edition v Preface to the first edition vii Notations xi 1 Preliminaries 1 1.1 Why Quantum Field Theory 1 1.2 Creation and annihilation operators 3 1.3 Special relativity . . . . . . . 5 1.4 Space and time in relativistic quantum theory. 8 1.5 Natural units . . . . 9 2 Classical Field Theory 12 2.1 A quick review of particle mechanics 12 2.1.1 Action principle and Euler-Lagrange equations 12 2.1.2 Hamiltonian formalism and Poisson brackets 14 2.2 Euler-Lagrange equations in field theory . 15 2.2.1 Action functional and Lagrangian 15 2.2.2 Euler-Lagrange equations 17 2.3 Hamiltonian formalism. 19 2.4 Noether's theorem . . . . 21 3 Quantization of scalar field3 28 3.1 Equation of motion . . . . . . 28 3.2 The field and its canonical quantization 29 3.3 Fourier decomposition of the field. 30 3.4 Ground state of the Hamiltonian and Ilormal ordering 34 3.5 Fock space. .. 36 3.6 Complex scalar field . 37 3.6.1 Creation and annihilation operators 37 3.6.2 Particles and antiparticles . . . 39 3.6.3 Ground state and Hamiltonian 40 3.7 Propagator .. ..... 41 xiii
  8. 8. XIV Contents -4 Quantization oj Dirac fields 47 4.1 Dirac Hamiltonian . . . . . 47 4.2 Dirac equation .. . . . . . 51 4.3 Plane wave solutions of Dirac equation. 54 4.3.1 Positive and negative energy spinor::; 54 4.3.2 Explicit solutions in Dirac-Pauli representation 56 4.4 Projection operators . .. 59 4.4.1 Projection operators for positive and negative energy states . . . . . . . 59 4.4.2 Helicity projection operators . 60 4.4.3 Chirality projection operators. 61 4.4.4 Spin projection operators . 62 4.5 Lagrangian for a Dirac field . . 63 4.6 Fourier decomposition of the field. 65 4.7 Propagator . 69 5 The 5 -matrix expansion 72 5.1 Examples of interactions 73 5.2 Evolution operator 75 5.3 5-matrix. 80 5.4 Wick; theorem . . 82 6 Prom Wick expansion to Feynman diagrams 87 6.1 Yukawa interaction: decay of a scalar . 87 6.2 Normalized states. . . . . . . . . 94 6.3 Sample calculation of a matrix element. 97 6.4 Another example: fermion scattering 101 6.5 Feynman amplitude 105 6.6 Feynman rules 106 6.7 Virtual particle~ 110 6.8 Amplitudes which are not S-matrix elements 112 7 Cross sections and decay rates 115 7.1 Decay rate. . . . . . . . . . . . . . . . . . . . . . . . . . 115 7.2 Examples of decay rate calculation. . . . . . . . . . . . 117 7.2.1 Decay of a scalar into a fermion-antifermion pair 117 7.2.2 Muon decay with 4-fermion interaction. 122 7.3 Scattering cross section 130 7.4 Generalities of 2-to-2 scattering 133 7.4.1 eM frame. . . 135 7.4.2 Lab frame. . . . . . . . 137 7.5 Inelastic scattering with 4-fermion interaction 140 7.5.1 Cross-section in CM frame 142 7.5.2 CroSl:i-section in Lab frame 143 7.6 Mandelstam variahles 144
  9. 9. Contents 8 Quantization of the electromagnetic field 8.1 Classical theory of electromagnetic fields. 8.2 Problems with quantization . . . . 8.3 Modifying the classical Lagrangian 8.4 Propagator .... . . . . . . . .. 8.5 Fourier decomposition of the field ". 8.6 Physical states ..... . ... 8.7 Another look at the propagator 8.8 Feynman rules for photons. 9 Quantum electrodynamics 9.1 Local gauge invariance . 9.2 Interaction Hamiltonian 9.3 Lowest order processes . 9.4 Electron-electron scattering 9.5 Electron-positron scattering 9.6 e-e+ --+ ~- ~+ . 9.7 Consequence of gauge invariance 9.8 Compton scattering . 9.9 Scattering by an external field . . 9.10 Bremsstrahlung . lOP, T, C and their combinations 10.1 Motivations from classical physics . 10.2 Parity . 10.2.1 Free scalar fields 10.2.2 Free Dirac field . 10.2.3 Free photon field 10.2.4 Interacting fields 10.3 Charge conjugation. 10.3.1 Free fields . 10.3.2 Interactions. 10.4 Time reversal . . . . 10.4.1 Antilinearity 10.4.2 Free fields . 10.4.3 Interactions 10.5 CP . 10.6 CPT . 11 Electromagnetic form factors 11.1 General electromagnetic vertex 11.2 Physical interpretation of form factors 11.2.1 Charge form factor F1 . . . . . 11.2.2 Anomalous magnetic mOE1ent F2 11.2.3 Electric dipole moment F2 . . . . xv 146 146 149 150 153 156 158 162 164 166 166 170 172 174 180 182 184 185 194 197 200 200 201 201 202 204 205 207 207 211 212 212 213 216 217 218 222 222 224 224 228 228
  10. 10. xvi 11.2.4 Anapole moment F3 . . . . . . . . . 11.3 Anomalous magnetic moment of the electron 11.4 Charge form factor . ... 11.5 Electron-proton scattering . . . . . 12 Renormalization 12.1 Degree of divergence of a diagram . 12.1.1 Superficial degree of divergence . . . . . 12.1.2 Superficial vs. real degree of divergence 12.2 Specific examples in QED 12.3 Outline of the program . 12.4 Ward-Takahashi identity . 12.5 General forms for divergent amplitudes . 12.5.1 Fermion self-energy. 12.5.2 Vacuum polarization . 12.5.3 Vertex function . . . . . . . . . 12.6 Regularization of self-energy diagrams 12.6.1 Vacuum polarization diagram 12.6.2 Fermion self-energy diagram. 12.7 Counterterms . . .. . . 12.7.1 Vacuum polarization diagram 12.7.2 Fermion self-energy diagram. 12.7.3 Vertex function . 12.8 FUll Lagrangian . 12.9 Observable effects of renormalization 12.9.1 Modification of Coulomb interaction 12.9.2 Running coupling constant ..... 12.9.3 Cancellation of infra-red divergences 13 Symmetries and symmetry breaking 13.1 Classification of symmetries 13.2 Groups and symmetries . 13.2.1 Symmetry group . 13.2.2 Examples of continuous symmetry groups 13.2.3 Generators of continuous groups 13.2.4 Representations . 13.3 Approximate symmetries. . . . . . . 13.4 Spontaneous breaking of symmetries 13.4.1 Discrete symmetry . 13.4.2 U(I) symmetry . 13.4.3 Non-Abelian symmetry 13.5 Goldstone's theorem . 13.5.1 Appearance of massless states. 13.5.2 Examples of Namhu-Goldstone hosons. Contents 229 230 239 241 245 245 245 248 250 252 253 256 256 257 259 260 260 265 267 267 270 272 273 275 275 276 279 283 283 284 284 286 289 292 293 294 295 299 300 301 301 303
  11. 11. Contents xvii 13.5.3 Interaction of Goldstone bosons . 305 13.6 Higgs mechanism . . . . . . . . . . . . . 308 14 YangMilis theory oj non-Abelian gauge fields 312 14.1 Gauge fields of non-Abelian symmetry . 312 14.2 Pure gauge Lagrangian. . . . . . . . . . . . 315 14.3 Interactions of non-Abelian gauge fields . . 318 14.3.1 Gauge interactions of other particles 318 14.3.2 Self-interactions of gauge bosons . . 319 14.4 Equations of motion and conserved currents 321 14.5 Quantization of non-Abelian gauge fields. 322 14.6 Quantum Chromodynamics . . . . . . . . . 323 15 Standard electroweale theory 326 15.1 Gauge group ..... . . . 326 15.1.1 Choice of gauge group. 326 15.1.2 Pure gauge Lagrangian 328 15.2 Spontaneous symmetry breaking 330 15.2.1 Introducing the Higgs boson multiplet 330 15.2.2 Gauge boson masses 332 15.2.3 Scalar modes . . . 333 15.3 Fermions in the theory. 335 15.3.1 Gauge interactions 335 15.3.2 Electron mass. . . 338 15.3.3 Yukawa couplings. 339 15.3.4 Other fermions in the model 340 15.4 Gauge boson decay . . . . . . . . . . 342 15.5 Scattering processes 346 15.5.1 Forward-backward asymmetry 346 15.5.2 Low energy weak interactions 349 15.5.3 High energy scattering. . 351 15.6 Propagator for unstable particles 355 15.7 Global symmetries of the model. 356 A Useful formulas 358 A.l Representation of 1'-matrices 358 A.2 Traces of 1'-matrices . . . . 360 A.3 The antisymmetric tensor . . 363 A.4 Useful integration formulas . 364 A.4.1 Angular integrations in N-dimensional space 364 A.4.2 Momentum integration in loops. 366 B Answers to selected exercises 368 Index 371
  12. 12. Chapter 1 Preliminaries The era of quantum mechanics began when Planck postulated that the modes of radiation of angular frequency w could be counted as particles with energy E = fu.J, and derived the iaw of blackbody radiation without running into infinities. The physical implications of this fundamental postulate became clearer when Einstein showed that quantization of radiation also explained the frequency depen- dence of photoelectric effect. This soon led to a proliferation of new ideas due to Bohr, Dirac, Born, SchrOdinger, Heisenberg and many others, who appiied the idea of quantization to particles and found remarkable success in describing subatomic phenomena. Dirac even constructed a relativistic theory of the electron, which was in ex- cellent agreement with subatomic experiments. Despite its success, there were some essential shortcomings of quantum mechanics of par- ticles. 1.1 Why Quantum Field Theory One of these shortcomings was philosophical. Quantum mechanics started with the description of light in terms of photons, but the clas- sical description of light was in terms of propagating electromagnetic fields. So a theory of photons required a prescription or how to quan- tize fields, a missing link between the two descriptions. As we shall see in the next several chapters, the structure required to quantize electromagnetic fields can be used to describe all elementary parti- cles as quantum fields. Even the relativistic wave equation of Dirac was found to have the simplest description ill terms of quantum fields 1
  13. 13. 2 Chapter 1. Prelirninades representing electrons and positrons. Another problem of particle quantum mechanics is that it is valid in the non-relativistic regime by definition. This is not just because it uses non-relativistic Hamiltonians to solve various problems. In fact, the whole design of non-relativistic quantum mechanics defies relativity. For example, it uses the concept of potentials, which is untenable in any relativistic theory since it assumes the transfer of in- formation at an infinite speed. Moreover, space and time are treated very differently in non-relativistic quantum mechanics. The spatial co-ordinates are operators, whereas time is a parameter1 and we typ- ically study the evolution of different operators, including the spa- tial co-ordinates, in time. Dirac equation, although covariant, treats space and time on different footings. In a truly relativistic theory, space and time should merge into a space-time, and one cannot make such fundamental distinction between the spatial part and the tem- poral part. There is a practical problem as well. In llature there are processes in which new particles are created or annihilated. For example {3- decay can be thought of as the decay of a neutron into a proton, an electron and an anti-neutrino: (1.1) The neutron is annihilated in this proce~s, while the proton. the electron etc are created. Quantum mechallic~ of particles deals with stable particles and their motions in various potentials. Without quantization of fields, any calculation involving the creation or an- nihilation of particles was essentially ad hoc if not downright wrong. Quantum Field Theory, by incorporating creation aud annihilation of quanta or particles as its essential feature, allowed meaningful calculations of experimentally verifiable results. When one goes to the relativistic regimc, onc cncounters the pos sibility of creating new particles from ellcrgy. For example, if an e1ectron and a positron collide with sufficient cllergy, the collision process can create extra electron-positron pairs: (1.2) This of course call happell only if the total killctic energy of the initial e-e+ pair is larger than the rest mass energy of the extra pair to be
  14. 14. 1.2. Creation and annihilation operators 3 created. To allow for such possibilities, one requires some mechanism to describe particle creation. Even when the initial and the final states of a problem involve the same particles, a quantum process can go through intermediate states involving other particles which are created and destroyed in the process. Such intermediate states can contribute to the ampli- tude of the process and affect the result. It is important to consider these possibilities when one wants to make precision tests of quantum theory. The creation and annihilation of the intermediate particles in such cases can be handled only by going over to Quantum Field Theory. 1.2 Creation and annihilation operators We now recall one problem in non-relativistic quantum mechanics which can be solved by the introduction of operators which create and annihilate quanta. This is the problem of the simple harmonic oscillator. Here we review this topic since it will be very useful for building up Quantum Field Theory in subsequent chapters. The Hamiltonian of the one-dimensional oscillator is given by H = 2~ (p2 +m 2 w 2x2) . (1.3) We define the following combinations of x and p: a=J1 (P-imwx),at =J 1 (p+imwx),(1.4) 2mliw 2mliw sa that the Hamiltonian is H = ~Iiw (ala + aat) (1.5) If p and x were purely classical objects, we could have written H = aat1iw = atanw. But they are not. In Quantum Mechanics, they are operators which satisfy the commutation relation [x,pj_ == xp - px = ih. (1.6) This implies that the objects a and at, defined in Eq. (1.4), are also operators. Unlike x and p, these are non-hermitian operators, and they satisfy the relation [a,atl_ = 1. (1.7)
  15. 15. 4 Chapter 1. Preliminaries Moreover, they have the crucial property which can easily be checked directly from Eqs. (1.3) and (104): [H,ul- = -fu.Ju, IH,atl_ = /twa!. (1.8) If we use the commutation relation of Eq. (1.7), we can rewrite the Hamiltonian as (1.9) The relations in Eq. (1.8) implies that, if In) is an eigenstate of H with eigenvalue En. i.e., if H In) = En In) , (1.10) then the states a In) and at In) are also eigenstates: Ha In) = (En - fu.J)a In), Hat In) = (En + fu.J)at In) . (1.11) In other words, the operator a seems to annihilate a quantum of energy, of amount hw, from the state. On the other hand, at creates a quantum of energy. In this sense, they are the annihilation and the creation operators, respectively. Of course, in any physical process energy must be conserved, so the operators a or at cannot appear alone in the expression of any measurable quantity. For example, in the expression for the Hamiltonian they appear together1 as we see in Eq. (1.9). The ground state can be denoted by 10). Since this is the state of lowest energy, the annihilation operator a, acting on itl cannot produce a state of lower energy. Thus, this state must be totally annihilated by the operation of a: alO) = 0. Using Eq. (1.9) now, we can easily find its energy eigenvalue: H 10) = ~fu.J 10) . (1.12) (1.13) The first excited state, which we will denote by II), can be defined by the state whose energy is larger than the ground state by one quantum, i.e., II} = at 10) . (1.14)
  16. 16. 1.3. Special relativity The normalized n-th excited state is defined by 5 It is easy to see that the energy eigenvalue of the state In} is given by 1 En =(n+ 2)1iw. (U6) o Exercise 1.1 If the ground state is norma.lized, i.e., if (O 10) = 1, ShOUl tha.t the excited states defined bll Eq. (1.15) a.re also normal- ized. o Exercise 1.2 The number opera.tor is de.fined b1J JV=ata. Sh.om tha.t its eigen'Uo.lue in the state giuen in Eq. (1.15) is n. (1.17) o Exercise 1.3 Construct the Hamiltonian for n fS'ystem of N simple harmonic oscillators of frequencies Wi. i = 1, ... , N. Construct the normalized state in which the i-tll. oscilla.tor is in the nith excited state. Construct the number operator and ShOUl that its eigen1Julue in the nbo'lle stnte is 2::i nj. o Exercise 1.4 Suppose in a. system there nre operators which abell a.nticommuto.tion rela.tions [an a11+ == aral + a~ar = OrlJ and lar,a.lll+ = O. foT' T.S = 1.,N. Construct the generic norm.o.l Zed. excited state. ShOUl that no state can huve two Of' mof'C qua.nta of the same s'Pecies. 1.3 Special relativity Since our goal is to discuss relativistic field theories. here we give a quick summary of the mathematical structure of special relativity, which will help in setting up the notation. In special relativity, the three spatial co-ordinates and time define the four components of the position 4-vector, which we will denote by xp . In keeping with established practice, we shall use Greek indices J.l., v, A, to denote components of a four-vector. and Latin indices i,l, k, to denote its spatial components. In other words, Greek
  17. 17. 6 Chapter 1. Preliminaries indices take on values 0,1,2,3, while Latin indices take on values 1,2,3. Thus, x" == (xO,xi ) == (cl,x) , (1.18) (1.19) where the factor of c is put in so that all components have dimensions of length. Here c is the speed of light in the VlUOuum, which is frame- independent according to one of the fundamental axioms of special relativity. The distance I between two points x and y in space-time can be written in terms of the Cartesian coordinates as 3 [2 = (xo _ yO)2 _ ~)Xi _ yi)2 1=1 We can write this in a compact form by defining the metric tensor 9"v by (1.20) Here, the values of the various components of the metric tensor are 9"v=diag(I,-I,-I,-I), (1.21) which means that 900 = +1, 911 = 922 = 933 = -I, and 9"v = 0 if J.I. # v. In addition, we have used the summation convention, which says that any index which appears twice in the same term is summed over. We can use the metric to lower the indices, e.g.: (1.22) . Similarly, one can use the inverse of the matrix gpoV} which will be denoted by gpo-v} to raise the indices: xlt = glt V xv The inverse matrix must satisfy gP-V gv>. = 6~ =gJJ.. } where o~ denotes the Kronecker delta which is defined by o~ = { 1 if I" = A > 0 if I" ;l A. (1.23) (1.24) (1.25)
  18. 18. 1.3. Special relativity Under a Lorentz transformation, the co-ordinates transform as 7 (1.26) where the specific form of the matrix A is not very important for our purposes. These transformations are defined to be the ones which leave x~xJl invariant: Since this must equal 9pq X PX(1, we obtain the relation A~ A" -9p.v p (1 - gpa I which the Lorentz transformation matrices must satisfy. (1.27) (1.28) o Exercise 1.5 Using the definition in Eq. (1.22), Lorentz tra.nsjo1"'ffio.tion rule for X JJ is giucn b1J show that the (1.29) o Exercise 1.6 [f 0. primed frame of reference moues with respect to an unprimed onf'. with a. uniform 'llelocitll v along the common x- nris, the Lorentz transformation equa.tions between the co-ordinates in the two frames a.re gillen by t' = '"((t - vx/c'), x' = '"((x - vt), y' = y, z' = z, (1.30) 'Where "t = (1 - v2/c2)-1/2. Show that these tra.nsfonna.tions can be 'Written in the joT'Tl oj Eq. (1.26) with A/w = g/UI +WjJ.V for infin.ites- imal unlues of v, wheTe WjJ.V = -WVjJ.' Any object that remains invariant like xjJ.xjJ. is a scalar quantity. Any four-component object that transforms like xl-'- is called a contravari- ant vector, whereas a four-component object transforming like x!' is a covariant vector. The scalar product of two vectors A~ and B~ will be written as g~"A~B" = A~B~ = A~B~ = AoBo+ A,B' = AOBo - A B, (1.31) where A B is the usual three-dimensional scalar product. If (x) is a scalar function, so is a small change in it: 6 = aa 6x~. (1.32) x~
  19. 19. 8 Chapter 1. Preliminaries Since the right side is a scalar while 6xp.. is a contravariant vector, It must be a covariant vector, which will be denoted. by 8p..rP, with the identification (1.33) For some arbitrary vector Ap..I The wave operator 0 is a scalar operatorI _l&" 2_ o = c2 &12 - V - a~{)J' . The momentum 4-vector of a particle is given by (1.34) (1.35) (1.36) where E is the energy of the particle. The invariant scalar product ~P~ is denoted hy (1.37) where m is the rest-mass of the particle, which we will call simply mass for the sake of brevity. o Exercise 1.7 Write Maxwell's equa.tions and the LorenbjoTce!O'r- mula in 4-'UectoT nota.tion. Get the signs right! 1.4 Space and time in relativistic quantum theory In non-relativistic quantum mechanics, the status of time and of the spatial ccrordinates are very different, as we mentioned earlier. In a relativistic theory, they must be brought to par before proceeding. This is achieved by treating space on the same footing as time, i.e., by treating the spatial co-ordinates as parameters as well. Thus, the quantities that we can talk about in relativistic quantum theories are functions of both space and time, Le., of space-time. Such functions
  20. 20. 1.5. Natural units 9 are called fields even in classical physics. For example, in classical physics, when we talk about the electromagnetic field, we talk about the electric field and the magnetic field which are functions of the spatial co-ordinates and time. Similarly, in Quantum Field Theory, we will talk about various kinds of fields, including the electromag- netic field. In classical or quantum mechanics, we could talk about objects which were functions of time alone. But the necessity to talk about space and time on the same footing automatically requires that we talk about fields. It is important to realize that the uncertainty relations, which form the basis of a quantum theory, should be interpreted differently for this purpose. In non-relativistic quantum mechanics, the relation (1.38) is interpreted. in terms of the uncertainties in measurements of the ccrordinate x and the corresponding momentum pz. On the other hand, the time-energy uncertainty relation, (1.39) is interpreted by saying that if one tries to make a measurement of energy to an accuracy ~E, the measurement process should take a time ~t which is related to ~E by the above relation. Now that x is also a parameter, the interpretation of Eq. (1.38) must also follow the line of that of Eq. (1.39). In other words, we should understand that the measurement of momentum to an accuracy ~Px requires that the measurement be performed in a region whose spatial extent is larger than ~x given by Eq. (1.38). 1.5 Natural units Elementary courses in physics start with three independent units, those of length, time and mass. As one proceeds, one introduces extra units, e.g., those of electric charge, temperature etc. In discussions on Quantum Field Theory, it is customary to use a system of units in which there is only one fundamental unit, which we can take to be the unit of mass. The units of length and time are defined by declaring that, in this system of units, 1i=I, c=l. (1.40)
  21. 21. 10 Chapter 1. Preliminaries Table 1.1: Dimensions of various physical quantities in natural units. To go from energy units (MeV or GeV) to conventional units, we need to multiply by the conversion factor. For electric charge, 'conventional' will mean Heaviside. Lorentz unit for us. Quantity Mass Length Time Energy Momentum Force Action Electric charge Dimension [M] IMt' [M]-' [M] [M] [M]' [M]D [M]D Conversion factor l/e2 nc n 1 l/e l/ne n v'nC The resulting units are called natural units, which will be used in the rest of this book. The dimensions of various quantities in these units are presented in Table 1.1. We have given the conversion factors for quantities expressed in energy units, which is universally preferred in calculations. Readers should note that most other texts give con- version factors for mass units. The electric charge is defined in such a way that Coulomb's law about the force F between two charges q, and q, takes the form F = q,q, r. (1.41) 47Tr3 Thus electric charge is dimensionless, as ~hown In Table 1.1. The unit of electric charge will be taken to be the charge of the proton, e. The fine-structure constant 0:: is related to this unit by the following relation: e > O. (1.42) o Exercise 1.8 In natural units, the inllerse lifetime of the muon is gillen b'y C' 5-1 F m T = 192"." ' (1.43) where m is the muon mass, 106 MeV. Whut is the dimension of GF in natural units'? Put in the factors of rt und c so that the equation
  22. 22. 001 be interpreted in con:le.nlional. units o.s well. From th.is, jind the. lifetime in seconds if GF = 1.166 X 10- 11 in MeV units. [Hint: Remember that MeV is a. unit oj ene'r9ll.1 r I II 1.5. Natural units 11
  23. 23. Chapter 2 Classical Field Theory Before getting into the question of quantization, we discuss in this chapter some of the basic techniques involved in the covariant for- mulation of classical field theories. 2.1 A quick review of particle mechanics 2.1.1 Action principle and Euler-Lagrange equations All theories of classical physics can be described by starting from a wonderful principle called the principle of least action. Suppose we are dealing with a non-relativistic system of particles and rigid bodies. We identify a set of coordinates q,.(t) which, together with their time derivatives 4,(t), adequately describe the configuration and the evolution of the system. If we are clever enough1 we can use our knowledge of the symmetries of the system to reduce this set to the smallest number of independent coordinates. On the other hand, in almost all the interesting cases it is useful not to make the reduction but instead choose a set of coordillates which have simple interrelations, making explicit the sym!netric:-; of the system. We now define two math~matical objects mlled the Lagrangian L and the action .fl1, related by I t2 .fl1 = dt L(q,(t),4,(t).t). ,, (2.1) The Lagrangian L is a function of the co-ordinates q, and the veloc- ities qrl and tl and t2 are the initial and final times between which we are studying the evolution of the system. If there is a source or 12
  24. 24. 2.1. A quick review of particle mechanics 13 sink of energy in the system under consideration, the Lagrangian can also depend explicitly on time. The principle of least action then states that, among all trajec- tories that join q,(tIl to q,(t2), the system follows the one for which .PI is stationary. To see what this implies, consider infinitesimal variations around a trajectory defined by q,(t), 4,(t): q,(t) ~ q,(t) + oq,(t) , 4,(t) ~ 4,(t) + :toq,(t). (2.2) This introduces the following variation on the action defined in Eq. (2.1): (2.3) where we have used the summation convention for the index T. We will assume that the system is conservative, so the Lagrangian does not have any explicit time-dependence. Integrating by parts, we can then write 1.'2 [8L d 8L] 8L 1'2o.vl = dt -8 - -d -8' oq, + 8---,--oq, . tl qr t qr qr tl (2.4) Since we are looking at paths connecting given initial and final con- figurations q,(tIl and q,(t2), we shall not vary the end points. So we set (2.5) and the last term of Eq. (2.4) vanishes. The principle of least action states that, near the classical path, the variation of the action should be zero. Since this is true for any arbitrary variation of the path, Le., for arbitrary oq,(t), o.vl = 0 implies that d (8L) 8L dt 84, = 8q, . (2.6) These are the equations of motion, also known as the Euler-Lagrange equations, corresponding to the co-ordinates (}r.
  25. 25. 14 Chapter 2. Classical Field Theory For a particle of mass m moving in a time-illdependent potential V(x), a convenient choice for the Lagrangian is L = ~lTlX2 - V(x), (2.7) and the Euler-Lagrange equation as derived from this Lagrangian is :tlTlX = -V'V, as expected from Newton's second law. (2.8) 2.1.2 Hamiltonian formalism and Poisson brackets The velocity-independent part of the Lagrangian can be thought of as generalized potential energy. So we can make the Euler-Lagrange equations look like Newton's second law by defining generalized me>- menta 'conjugate to q/ as p, = "a L(q, q). uqr (2.9) There will be one such equation for each value of the index r. Since the Lagrangian is a function of the positions and the velocities, the right hand sides of these equations are in general functions of the q's and the (j's. Let us assume that these equations are invertible, Le., the q's can be solved in terms of the positions and the momenta. Then the Hamiltonian is defined by the Legendre transformation H(p,q) = Prqr(p,q) - L(q,q(p,q)). Indeed, by differentiation, we find . . (aL aL. )dH = q,dp,. +p,dq,. - n-dq,. + oo-dq,. uqr U{Jr (2.10) (2.11 ) The terms with dqr cancel each other owing to the definition of the momentum in Eq. (2.9). The term with dq,. can be simplified by the use of the Euler-Lagrange equation and Eq. (2.9), so that we can finally write dH = qrdpr - p,.dq,. , (2.12)
  26. 26. 2.2. Euler-Lagrange equations in field theory 15 confirming the fact that the Hamiltonian is indeed a function of the positions and the momenta. MoreoverI it directly leads to Hamilton's equations of motion: . aH Pr=-aqr (2.13) (2.14) The Poisson bracket of any two dynamical variables is defined in the following way: [II. hlp = alI ah _ af, ah. aqr apr apr oqr Using this definition, it is then trivial to check that and that the Hamilton'5 equations can be rewritten as qr = [qro H]p. Pr = jpro H]p More generally, if f is any dynamical variable. we can write . df af f= dt = at +lJ,HJp. where ~ occurs if f has an explicit time-dependence. (2.15) (2.16) (2.17) 2.2 Euler-Lagrange equations in field theory 2.2.1 Action functional and Lagrangian Going over to a field theory is trivial. Let us consider our system to be a field or a set of fields. A field is essentially a set of numbers at eacb point in space-time. We still want to describe this system by an action, and we still want to write the action as the time integral of a Lagrangian. But time is not the only independent variable in this case. We have to incorporate the information from every point in space into our action as well. We can do this by writing the Lagrangian as a spatial integral of some function of the fields. This has the added advantage of putting space and time on the same footing in the integral. The integrand has then the dimensions of Lagrangian density. However, in discussions of field theory, this is
  27. 27. 16 Chapter 2. Classical Field Theory always called the Lagrangian, and the word 'density' is implied for the sake of brevity. We will always use the word Lagmngian in this sense from now on. However, we will denote it by a different symbol, viz, .!'. If, at any point, we need to use the word Lagrangian in the sense used in particle mechanics, we will call it the total Lagrangian and will denote it by L. The Lagrangian should depend on the fields, which we denote generically by A Here, different values of the index A can denote completely independent fields, or the different members of a set of fields which are related by some internal symmetry, or maybe the components of a field which transforms non-trivially under Lorentz transformations, e.g., like a vector. In any case, the field or fields denoted by A take the role of the co-ordinates qr of Eq. (2.1). What then replaces the velocities tIr of Eq. (2.1)? These should be the derivatives of the fields. However, we cannot have only the time derivative in this case, since spatial co-ordinates are independent parameters as well. We therefore should expect the Lagrangian to depend on the derivatives 8/1 A . And finally, since we are interested about fundamental fields, we should not expect any source or sink of energy or momentum in the system. Thus, the Lagrangian should not depend explicitly on the space-time co-ordinates. So finally we can write the action in the form PI = Ind'x.Y (A(x), a~A(x)) , (2.18) where the region of integration n is the region of space-time we are interested ill. For any physical experiment this region extends far enough away from the equipment and far enough into the past and future of the actual experiment so as to make the effect of all events outside that region negligible on the experiment. Usually n is taken to be the entire space-time for convenience. A few things are worth noticing at this point. The action must be Poincare invariant in a covariant theory. which means that it should be invariant under Lorentz transformations as well as space-time translations. Since the space-time volume element d4x is Poincare- invariant by itself, the Lagrangian as defined in Eq. (2.18) must also be invariant. The Lagrangian density .Y is a function of x through its depen- dence on the fields A and their derivatives. The action on the other
  28. 28. 2.2. Euler-Lagrange equations in field theory 17 hand is a number, i.e., it is a rule that associates a real number with a given configuration of the fields. If the field configuration is changed the number also changes in general. Such objects are called function- also The action is a functional of the fields. The total Lagrangian L is a function of t but a functional of the fields at a given t. 2.2.2 Euler-Lagrange equations Given the Lagrangian, the classical equations of motion can be de- rived from the principle of least action. This states that the system evolves through field configurations which keep the action PI sta- tionary against small variations of the fields q,A which vanish on the boundary of n. Consider therefore the variations A(X) ~ q,A{x) + 6q,A{x), a~A{x) ~ a~A{x) + a,,6q,A(x), (2.19) such that 6A vanishes on the boundary, which we denote by an. Using the functional dependence of the action denoted by Eq. (2.18), we can write (2.20) We now invoke Gauss theorem, which, for 4 dimensions, reads rd"x a F~ = r dS FI' in It Jan J1. ' (2.21) where F~ is any well-behaved vector field, dS" is the outward point- ing volume element on the boundary an. Using this, we can write the last term in Eq. (2.20) as an integral over an. However, the in- tegrand now contains 64>A, which vanishes everywhere on af!. Thus, the last term in Eq. (2.20) vanishes.
  29. 29. 18 Chapter 2. Classical Field Theory As for the other terms, we notice that they must vanish for arbi- trary variations of the fields. This can be true only if the expression multiplying o4>A vanishes, which implies (2.22) These equations, one for eoch 4>A, are called the Euler-Lagrange equa- tions for the system. o Exercise 2.1 The Lagrangian (density) of an electromagnetic jield interQcting 'With a c'u:rTent jJl is given by U) _ _ ~F"VF -}' A".z-4 Jlll / ' , (2.23) where F'lii = 8j.iAv - 8v AJJ> Treating the All's as the fields, find the EulerLagro.nge equations and show that they give the inhomo gcnoous Maxwell's equations. o Exercise 2.2 For a real scalur field A are defined in analogy to systems with finite number of degrees of freedom: (2.26) Here ciA is the time-derivatiye of 4>A. This is a partial derivative, since I>A depends also on the spatial co-ordinates. MoreoverI L is a functional, so that the derivative indicated in Eq. (2.26) is really the derivative of a functional with respect to a function, Le., a func- tional derivative. We will avoid using the partial derivative sign to remember this fact and will use the symbol J instead. In order to perform the functional differentiation in Eq. (2.26), we need to know the derivatives, with respect to cl>'s, of the fields cl>1 as well as of 4>A and '17cl>A, since these are the building blocks of L. We take the cue from particle mechanics, where the generalized coordinates qi and their corresponding velocities qi are independent variables in the Lagrangian formalism. In particular, oq;/oqj = Jij , whereas oq;/oqj = O. For fields, we replace the discrete index by the space co-ordinate x and the index A which characterize the field. So the fundamental functional derivatives are given by (2.27) whereas the functional derivatives of the ficlds and their spatial derivatives with respect to 4>A vanish identically. In Eq. (2.27), no- tice that the functional derivative of 4> with respect to itself is not dimensionless, but has the dimensions of an inverse volume. We now introduce the Hamiltonian of the system, which also should be understood to really mean the Hamiltonian density and will be denoted by the symbol . The volume integral of , which is usually called the Hamiltonian in the case of particle dynamics, will be called total Hamiltonian and denoted by H. The Hamiltonian can be defined in terms of the canonical momenta in the following way: (2.28)
  30. 31. 20 Chapter 2. Classical Field Theory (2.31) (2.30) Notice that the Lagrangian is a function of the fields and their par- tial derivatives with respect to all space-time co-ordinates. In the Hamiltonian, the field derivatives with respect to time have been re- placed by the canonical momenta. However, the Hamiltonian can still be a function of the spatial derivatives. The total Hamiltonian is a functional in the same sense as the total Lagrangian. o Exercise 2.4 TT'eo.ting the AI"8 as the fields, find the co.nonico.l momenta Q.ssocinted with them, u.sing the Lo.grnngio.n of Eq. (2.23). Ca.n these equations, expressing the momenta in teT1lS oj AJ."s and their deri'vatives, be inuerled to 80lue the AJ.I's'? II not, can 1I0U think. of Q Teason wh'll'? o Exercise 2.5 There is some freedom in defining the electroma.g- netic 'Potentials AJ.'. Usin.9 this freedom, we can make AO = O. This is an example oj gauge fixing. In this gauge, treating the compo- nents of A as the independent fields, find out the momenta. Can these equations be inuerted to soLve the Ai's, wheT'e the index i runs oueT' spatial components onlll'? The Poisson bracket of any two functionals FI and F2 can be defined as J3 (JF' JF2 JF, JF2 ) [F"F2]p = d X JA(t,x) JITA(t,x) - JIIA(t,x) JA(t,x) (2.29) These derivatives can be reduced to the following fundamental ones which are similar to Eq. (2.27): J a a 3 JA(t,x) (t,y)=JAJ (x-y), J A 3 JITA(t,x) IIa(t,y) = JaJ (x - y), whereas the functional derivatives of the fields with respect to the momenta, as well as vice versa, vanish. Thus, for example, [ A ] J3 (JA(t,X) JITa(t,y) (t,x),IIa(t,y) p = d z JC(t,z) JITc(t,z) _ JA(t,x) JIIa(t,y)) JITc(t,z) JC(t,z) = Jd3 z (J~J3(X - z)J~J3(y - z) - 0) = J~J3(X - y). (2.32)
  31. 32. 2.4. Noether's theorem 21 o Exercise 2.6 Find the. HamillonQ:n faT the complex srotnT field whose Lagrangian has been gi:uen in Ex. 2.3 (p 18). a Exercise 2.7 Con.sider 0. rodiation field in 0. onedimensl.onQ,l box 01' equi'Uo.lenUy displa.cements on a. string of Length I with jUed ends. (This wo.s the "'11 first attempt at fieLd theo'1l.) L = l'dx [ (::)' _ c' (:~) '] (2.33) Let 'U.s e:tpres8 the field u in a. Fourier series: (2.34)W, = bee/l. ~ ,,(x,t) = Lq.(t)sin (W:X) , k;;:l 0.) CalcuLate the 'momentum' Pk, canonically conjugate to qk. b) Write down th.e Ha.miltonia.n as Q function oj q's Qnd p's. c) Take qk and Pk to be operotoTS with Iqk,Pi] = ihtSkj . Define opera.loTS ak, ak bll q = Jn [a e-'wo< + at eiWO '] (2.35)k 2lwk k ~.. Find th.e commutators oj ak, at 'mith themsdlles Qnd one an other. d) CaLculate the Hamiltonian in tenns oJ a's and at's. 2.4 Noether's theorem Some Lagrangians are known fwm basic physics. But generally a system may not be as well understood, and we have to construct a Lagrangian that we expect to describe the system. In such construc- tions, it is often useful to rely on symmetry, i.e., transformations which leave the system invariant. It so happens that associated with each symmetry of a system is a conserved quantity. Thus if we know some conserved quantities of a system we can work backwards to find the symmetries of the system and from there make a guess at the form of a convenient Lagrangian. The result which relates symme- tries to conserved quantities is known as Noether's theorem after its discoverer, while the conserved quantities are called Noether charges and Noether currents. Let us consider infinitesimal transformations of the c~ordinate system xll --I> x'Po = xll + dxP 1 (2.36)
  32. 33. 22 Chapter 2. Classical Fjeld Theory under which the fields transform as (2.37) The change in the action resulting from these transformations is given by oJil = in,d"x' ~(4)'A(X'),a;.,A(X')) -ind4x ~ (4)A(X), 8"4>A(X)) , (2.38) where Of is the transform of n under the co-ordinate change. If os;( = 0, we say that the theory is invariant under the transformation. In the first integral x' is a dummy variable, so we can replace it by x and write oJil =in,d"x ~ (4)'A(x), a;.4>'A(x)) - ind'x ~ (4)A(x), 8"4>A(x)) = ind4x [~(4>'A(x), a;.4>'A(x)) - ~ (4)A(x), 8"4>A(x))] + r d"x ~ (4),A(x),a;.4>,A(x)) (2.39) Jrr-n The last term is an integral over the infinitesimal volume fl' - fl, so we can replace it by an integral over the boundary an, = r dS" oX''~(4)A,8"4>A) Jon = ind"x 8" (ox"~ (4)A, 8"4>A)). (2.40) Here we have suppressed the space-time index x as we no longer need to distinguish between x and x'. To get the first equation we have replaced 4>,A by 4>A as the difference is of higher order in Ox". In the second equation we have used Gauss theorem as shown in Eq. (2.21). At this point, it is convenient to define the variation for fixed X, which gives the change in the functional forms only. For any function f(x) whose functional form changes to j'(x), we can write bf(x) =j'(x) - f(x) = [j'(x') - f(xlJ - 1J'(x') - J'(x)J = of(x) - 8"f(x)ox" , (2.41)
  33. 34. 2.4. Noether's theorem 23 where we have again ignored terms of higher order in 6x"" in writing the last equality. Note that this variation J commutes with the par- tial derivatives op. as the variation is taken at the same space-time point. This property of J allows us to simplify the integrals of Eq. (2.39) since we can write (2.42) (2.43) (2.44) using Euler-Lagrange equation, Eq. (2.22), to get the last equality. Putting this in Eq. (2.39) and using Eq. (2.40), we obtain 6d = ind"x a~ [ a(:~A) JA + 2'6x~] = ind"x a~ [ a(:~A) 6A - T~v6xv] where we have used Eq. (2.41) again to bring back 6A, and defined T~v = a2' ,,"...A g~V '" a(a~A)u", - ..L, which is called the stress-energy tensor. We can now define a current (2.45) and write (2.46) This relation holds for arbitrary variations of the fields and co- ordinates provided the equations of motion are satisfied. Even though the right hand side of Eq. (2.46) is the integral of a total divergence, it does not necessarily vanish, because the current need not vanish on the boundary. If however there is a symmetry,
  34. 35. 24 Chapter 2. Classical Field Theory 6Jif vanishes for arbitrary n even without the use of the equations of motion. In that case, we obtain a conservation law (2.47) when the equations of motion are satisfied. This result is called Noether's theorem, and the current J" in this case is called the Noether current. Note that J" is defined only up to a constant factor. It may so happen that under the symmetry transformations of Eqs. (2.36) and (2.37), the action does not remain invariant but instead changes by the integral of a total divergence, (2.48) Then the expression for the corresponding conserved current will be the same as in Eq. (2.45), plus V". It is easy to see that Eq. (2.47) implies the conservation of a charge, called Noether charge. It is defined by an integral over all space, (2.49) provided the current vanishes sufficiently rapidly at spatial infinity. This is because (2.50) (2.51) The last expression can be converted into a surface integral on the spatial surface at infinity, on which it should vanish. Thus, dQ =0 dt ' i.e., the charge Q is conserved by the dynamics of the system. In what follows, we will find the specific form of the conserved current for various kinds of symmetries. Space-time translations The action constructed from relativistic fields is invariant under Lorentz transformations as well as translations of the co-ordinates.
  35. 36. 2.4. Noether's theorem 25 Let us look at translations first. Suppose we make small changes in the space-time co-ordinates, defined by (2.52) where a~ is constant, i.e., independent of X~. The fields do not change at any point, so Since Eq. (2.47) is true for arbitrary a~, we conclude that a~T~v = 0 (2.53) (2.54) This equation implies the conservation of the quantity p~ = Jd3 x T'~ , (2.55) which is the 4-momentum of the field. Using the definition of the stress-energy tensor from Eq. (2.44), it "is easy to see that pO is the total Hamiltonian defined through Eq. (2.28). o Exercise 2.8 Find the stress-eneT91l tensor JOT the following fields: a) rool scalar, as in "Ex. 2.2 (p 18); b) complex 'calar, as in Ex. 2.3 (p 18); c) electromagnetic jeld, a.s in Eq. (2.23) with f ~ O. Lorentz transformations In this case the infinitesimal transformations are (2.56) where the w~/,I are independent of x, and they are also antisymmetric because x~x~ is invariant. The field variations under this kind of operation are written in terms of a spin-matrix L by (2.57) The factor of 4appears because otherwise each independent compo- nent of w is counted twice because of antisymmetry. Putting these in Eq. (2.45), we obtain a [ a.:f 1 ("Ap)A->.B T~A p] - 0 (2.58) ~ a(a~~A) 2WAp L. B'" - WApX -
  36. 37. 26 Chapter 2. Classical Field Theory Pulling out the w~p which are constants, and also using the antisym- metry property of them, we can write this as & M"~P - 0 " -, where The conserved charges for this case are J~p = Jd3x MO~p (2.59) (2.60) (2.61) The space components of this, Jij I are related to angular momentum by J k - ~A and their derivatives. The index r is not summed over, it indicates the type of symmetry. In other words, there may be several independent symmetries in a system. We can treat them separately by defining the conserved current for the roth symmetry as " &.!L' o4>A Jr = &(&,,4>A) O, q;,t -+ e,q8 t.
  37. 38. 2.4. Noether's theorem 27 a.) Use Noether'g theorem to Jind the colTesponding conserued cur- rent p'. Verifll 0s.j" = O. b) Ca.lcula.te the conserued Noether turn.',nt for the, sa.me transfor- mation when the electroma.gnetic field is also included, and the combined La.gra.ngian is !L' = _~F"vFvV + [(a" - iqA,,)tll(a,. + iqAv)] _m't - V(t) (2.65) c) The EuleT-La.gT'unge equations for AI' should gi've the Maxwell equations: OIJ,FJlV = j"'. ShoUl that the 'current' ~ppea.ring in this equation is the same as the CUTTent obtained in part (b) abolle. INote: Since the conserued current inuolues A~. the Lagrangian cannot be written like Eq. (2.23) for this case.I o Exercise 2.10 Consider the Lagrangian oj a. real sca.lar field gi'lJen in Ex. 2.2 (p 18), with m = 0 and V() = >.'. Verijy that the "ction has the 8'Ymmetrll x ~ bx, (2.66) Find the conserued CU1Tent corresponding to this 8'ymmetry. o Exercise 2.11 Suppose the action of a certain fl.eld theory is in- 'llarlant under space-time translations as well us dilatations x-tbx, (2.67) Show that the stTess-energ-y tensor in this cuse 'LS traceless, i.e., T" -0"- .
  38. 39. Chapter 3 Quantization of scalar fields Having outlined the necessary aspects of classical field theory in Ch. 2, we now turn to quantization of fields. In this chapter, we take up the quantization of the simplest kind of field which is de- scribed classically by one Lorentz invariant quantity at each space- time point. Such fields are called scalar field". From n0W on, we will employ the natural units outlined in 1.5. 3.1 Equation of motion For free particles of mass m, the 4-momentum satisfies the relation pI'p~ - m 2 = O. (3.1) In quantum theory, momentum is treated as an operator whose co- ordinate space representation is given by PIA ---+ ia~ I (3.2) which operates on the wavefunctions. Thus, if we tried to repre- sent the particles by a single wavefunction >(x), we would simply substitute Eq. (3.2) into Eq. (3.1) and obtain (0 + m2 ) >(x) = O. (3.3) This is called the Klein-Gomon equation after the people who sug- gested it. It looks simple enough, but there is a problem. The energy eigen- values satisfy (3.4) 28
  39. 40. 3.2. The field and its canonical quantization 29 (3.5) Notice that energy can be negative, and its magnitude can be ar- bitrarily large since the magnitude of p, which we denote by p, is not bounded. Thus, the system will have no ground state. The sys- tem will collapse into larger and larger negative values of energy and will be totally unstable. Hence, a wavefunction interpretation of the Klein-Gordon equation is not possible. We will see that this problem will be evaded once we interpret (x) as a quantum field, with a proper prescription for quantization. 3.2 The field and its canonical quantization So we want to interpret (x) as a field. The classical equation of motion of this field will be given by Eq. (3.3). According to Ex. 2.2 (p 18), it can be derived from the Lagrangian .!/ = i(8")(a~) - i m22 D Exercise 3.1 Use the dejinition oj the Lagrangian in Eq. (2.18) to find the mass dimensions of .!L' in natuTul units, assuming the space- time is N-dimensional. Use this to show that the mass dimensions oj the jield 4> is (N - 2)(2. o Exercise 3:.2 From the Lagrangian in Eq. (3.5), shoUJ that the mo- mentum canonical to

    (x) , and that the Hamiltonia.n is gi1Jen b1:l ' = ~(n' + (V4' + m'4>'). (3.6) (3.7) In particle mechanics, the Poisson brackets between the co- ordinates and the canonical momenta were derived in Eq. (2.15). In order to go over to Quantum Mechanics, one can just use the replacement (38) whereby one obtains the commutation relation of Eq. (1.6). Similarly, in the case of the fields, once we obtained the canonical momentum II, we can use the same prescription to replace Eq. (2.32) by 1(t,x),II(t,y)L = iJ3 (x - y) (3.9)

  40. 41. 30 Chapter 3. Qual'tization of scalar fields This is called the canonical commutation relation, which implies that and II can no more be considered as classical fields. They should now be treated as operators. The quantization procedure is a direct consequence of this commutation relation, as we will gradually see. It might be a little dissatisfying to see that although we are try- ing to build up a relativistically covariant theory, the commutation relation of Eq. (3.9) does not seem to be a covariant equation. It seems that it somehow gives a preferential treatment to the time component and therefore cannot be valid in a general frame of refer- ence. However, appearances are sometimes deceptive, and Eq. (3.9) can really be valid in any frame. The commutator is non-zero only if we consider the field and the canonical momentum II at the same space--time point. If two space-time points coincide in one frame, they will coincide in any other frame. Thus, at least for this part, there is no difference if we go to any other frame. Next, consider the Lorentz transformation properties of the commutator bracket on the left hand side. Since is a scalar and II = if, = 8o, the non- trivial Lorentz transformation property of the left hand side comes only from 80. In other words, the left hand side transforms like the time component of a 4-vector. As for the right hand side, notice that d4x 53(x - y) = dt when integrated over a spatial region containing the point y. By definition, dt transforms like the time component of a 4-vector. Since d"x is a scalar, this means that 53 (x - y) also transforms like the time component of a 4-vector, and therefore Eq. (3.9) is covariant. 3.3 Fourier decomposition of the field A classical field, being a function of space-time co-ordinates, can be Fourier decomposed. For the free field, we can write (3.10) Here the factor outside the integral sign is just a normalizing fac- tor, whose convenience will be seen later. Inside the integral sign, apart from the usual Fourier components A(p), we have put a factor
  41. 42. 3.3. Fourier decomposition of the field 31 5(p' - m') to ensure that satisfies Eq. (3.3), (0 + m') (x) = (2:)'12 J d"p (-p' + m')5(p' - m')A(p)e-ip . X = O. (3.11) The 5-function, in association with the factor (-p' +m'), guarantees that the integral will vanish. Let us now assume that rI>{x) is a hermitian operator, i.eo, classi- cally speaking the value of the field at any point is a real number. The generalization to complex scalar fields will be taken up in 3.6. The hermiticity condition, (x) = t(x), then implies that the Fourier components satisfy the condition A(-p) = At(p). (3.12) (3.13) We now introduce the step function e, defined for any real variable z by { I if z > 0, e(z)= O! ifz=O, if z < O. Obviously, e(po) + e(_pO) = I. Inserting this factor in Eq. (3.10) and changing the sign of the integrated variable p in the second term, we can write (x) = 1 Jd"P 5(p' _ m')e(po) (A(p)e-'P'x + At(p)e'P'x) (2,,) 'j, (3.14) This form, then, involves only the states of positive energy. In fact, the Fourier decomposition can be written in a somewhat simplified form if we eliminate pO from Eq. (3.14). We will use the result that given a function f(z) which vanishes at the points Zn for n=1,2, ... , ( ) "" 5(z - zn) 5 f(z) = ~ Id!/dzl,=,,, ' (3.15) provided the derivatives themselves do not vanish at the points Zn. Thus we can write 5(p' - m') = 5 ((pO)' - En = 21~OI [5(pO - Ep ) + 5(l + Ep )] (3.16)
  42. 43. 32 Chapter 3. Quantization of scalar fields (3.18) where from now Oll, we reserve the symbol Ep for the positive energy eigenvalue: Ep =+VP2 +m 2 (3.17) When Eq. (3.16) is substituted into Eq. (3.14), the second 6-function does not contribute to the integral since this is zero everywhere in the region where 8(po) i' o. The other 6-function contributes to the integral and gives J d3p ( . t . )(x) = a(p)e-'px + a (p)e'px V(21r)32Ep where and we have introduced the notation a(p) = A(p) JIE;. (3.19) (3.20) (3.22) This is the form in which the Fourier decomposition will be most useful for us. The canonical momentum can also be written in a similar form involving integration over the 3-momentum only: II(x) = (x) = Jd 3 p iV2(~;)3 (_a(p)e-iPX + at(p)eiPX ) (3.21). In the discussion about simple harmonic oscillators in L2, we started with the operators x and p, but switched to another set of operators, viz, a and at. The situation is quite similar here. We started with (x) and II(x), hut now we can do everything in terms of a(p) and at(p). Just as the commutation relation between x and pin Eq. (1.6) led to the commutation relation of Eg. (1.7) between a and at 1 here also we can use the canonical commutation relation of Eg. (3.9) to derive the commutation relation between the a's and the at's. For this, we need to invert the Fourier transform equations of Eqs. (3.18) and (3.21) to express a(p) and at(p) in terms of (x) and II(x). From that, using Eg. (3.9), one can show that [a(p) ,at (p')L = 63 (p - p'), [a(p), a(p')L = 0, [at(p) ' at (p')L =0
  43. 44. 3.3. Fourier decomposition of the field 33 (3.25) These commutation relations are reminiscent of the corresponding equations obtained for a linear harmonic oscillator in Eq. (1.7). If one takes a different prefador in Eq. (3.10) than the one we used, the right hand side of the commutator of a(p) and al(p) contains some extra factors. We can derive Eq. (3.22) from the canonical commutation relation of Eq. (3.9), but we leave that for Ex. 3.4 (this page). Here, let us prove the converse. If we assume Eq. (3.22), using the Fourier decomposition of Eqs. (3.18) and (3.21), we can write [q,(t,x),IT(t,y)J- = 2(2~)3 Jd 3 p Jd 3 p' j~ x {[a(p),al(p')L e-ip.x+ip'.y - [al(p),a(p')L eiP.X-iP'.Y} (3.23) where xO = yO = t. The commutators give a-functions, and so the integration over p' can easily be performed. Note that once this is done, it forces p = pi, which also means E'l = Epl. In the exponents, the time components in the dot products therefore cancel, and we are left with This immediately reduces to Eq. (3.9) if we recall the integral repre- sentation of the 6-function of the form: 63(x _ y) = Jd 3 p e-ip.(,,-y) (2")3 o Exercise 3.3 Use the commutntOi5 oj Eq. (3.22) to ShOUl tha.t [4>(t,x),8i 4>(t,y)!_ = 0 (3.26) o Exercise 3.4 Write a(p) and at (p) us inuerse Fourier tra.nsjonns of a.nd fi, a.nd deriue the commuta.tion relations oj a, at from the canonica.l commuta.tors oj and n. We can now derive the total Hamiltonian of the system. Let us start with the volum.e integral of ,J'l' given in Eq. (3.7), replace q, and IT by the expressions obtained in Eqs. (3.18) and (3.21) and perform
  44. 45. 34 Chapter 3. Quantization of scalar fields the integral over all space. Each of the exponential factors will give us a momentum cS"-function. If we now integrate over one of the momenta, say pI, we get for the different terms in the Hamiltonian the following results: J ,flx rr2 (x) = J ,flp r; [_u(p)u(_p)e-2iEp, + u(p)at(p) +at(p)a(p) - at(p)at(_p)e2iEp,] , (3.27) J d3x (V.p(x))2 = J d3p i::. [a(p)u(_p)e-2iEp, + a(p)at(p) +at(p)a(p) + at (p)at(_p)e2iEp,] , (3.28) m2Jd3xq,2(x) = Jd3p m 2 [a(p)a(_p)e-2iEp'+a(p)at(p) 2E. +at(p)a(p) +at(p)at(_p)e2iEp'] . (3.29) This yields H = ~ J d3p E. [at (p)a(p) + a(p)at(p)] (3.30) Once again, the similarity with Eq. (1.5) is obvious. Moreover, one can check the following relations: [H,a(p')I_ = -E.,a(p') [H,at(p')J_ = +E.,at(p') , (3.31) which are similar to the relations in Eq. (1.8). The analogy between the simple harmonic oscillator and the field is now complete. It is clear that we can interpret a(p) as the annihilation operator and at(p) as the creation operator for a field quantum with momentum p. What was the positive energy component of the classical field now annihilates the quantum, and the negative energy component now creates the quantum. This quantum is what we call a particle of positive energy. 3.4 Ground state of the Hamiltonian and normal ordering The ground state of the field, which is called the vacuum in field- theoretical parlance, can now be defined in analogy with Eq. (1.12).
  45. 46. 3.4. Ground state of the Hamiltonian and normal ordering 35 Here, it reads a(p) 10) = for all p. We will assume that this state is normalized, Le., (010) = 1. (3.32) (3.33) (3.34) However, there is one problem here. which was not present in the case of the linear harmonic oscillator. To see this, let us use the commutation relations of Eq. (3.22) to rewrite the expression of Eq. (3.30) in the form H = JtPp Ep [at(p)a(p) + ~83(0)p] , which would be the analog of Eq. (1.9) for this case. The subscript p after the 8-function in this equation is just to remind us that it is a 8-function for the zero of momentum, not the zero of co-ordinate. The distinction is not really relevant here, but the notation will be useful later. What comes out from this way of writing is that, if we consider the ground state energy, Le., the expectation value of the total Hamiltonian in the ground state, we obtain (3.35) which is infinite! The point is that, here we have one oscillator for each value of the momentum. Thus, we have an infinite number of oscillators. Since each oscillator has a finite ground state energy, the contributions from all such oscillators add up to give us an infinite result for the total ground state energy. This, by itself, is not a catastrophe. After all, energy differences are physical quantities, absolute values of energies are not. We can just redefine the zero of energy such that the ground state energy vanishes. For this, we need a consistent prescription so that we do not run into trouble with other variables. This prescription is called normal ordering. Stated simply, it means that whenever we encounter a product of creation and annihilation operators, we define a normal-ordered product by moving all annihilation operators to the right of all creation operators as if the commutators were zero. Once
  46. 47. 36 Chapter 3. Quantization of scalar fields we have done the rearrangement, we treat them again as operators with the usual commutators. Using this algorithm on the expression of Eq. (3.30), we find that the normal-ordered total Hamiltonian, denoted by :H:, is given by This expression immediately shows two things. First, for any state 11lI), (1lIIH: I Ill) = Jd3 p Ep (1lIIat(p)a(p)1 Ill) = Jd 3 p Ep Ila(p) 11lI) W (3.37) which is always non-negative. Second, using Eq. (3.32), we obtain (OI:H:IO) =0, (3.38) so that the vacuum defined earlier is really the state of lowest energy, i.e., the ground state. D Exercise 3.5 a.) For 0. real massiue scalar fleld, calculnte the 4-momentum PIJ = rd3 x ']'O1l in terms of creo.tion and a.nni- hila.tion operators. Show tha.t nonnal ordering is not required jor P'. b) Show that Ie/>, P"I_ = Wee/> 3.5 Fock space So far, we have defined only the vacuum state, which is a state with no particles. We shall need other states, in particular states with specified particle content, when we start describing physical events. We can define such states in analogy with excited states of oscillators. For example, we will define a one-particle state as Ip) =at (p) 10) (3.39) This state contains one quantum of the field 4> with momentum pi' = (Ep , p). Such states have positive norm, since (3.40)
  47. 48. 3.6. Complex scalar field 37 which comf'.-8 from the commutation relations and the definition of the vacuum. Similarly, we can define many-particle states. If a state has N particles with all different momenta PI, P2, ... PN, it is defined by (3.41) On the other hand, if we want to construct a state with n particles of momentum P, it will be given by 1 ( )"Ip(n)) =v'n! at(p) 10), (3.42) where the prefactor is needed for proper normalization. Such multi-particle states distinguish field quantization (also called second quantization) from single-particle quantum mechanics (or first quantization). The vacuum, together with single particle states and all multi-particle states, constitute a vector space which is called the Pock space. The creation and annihilation operators act on this space. D Exercise 3.6 The numbeT O'PeratoT in lhe Fock RpaCf'. can be de- lined as Show that 1A",at'(k)l_ = a'(k), 1A",a(k)l_ = -a(k). (343) (3.44) Show that this operator corrccttll counts the number of particles in the states given in Eqa. (3.41) and (3.42). 3.6 Complex scalar field 3.6.1 Creation and annihilation operators The Lagrangian for a complex scalar field (x) is given by .:t: = (8"t)(o"2. From this form, it is straightforward to calculate the momenta conjugate to , and 2, and impose the canonical com- mutation relations. We can perform Fourier decomposition of both , and 2 in the way outlined in 3.3, and define the annihilation and creation operators corresponding to each of them. The canonical commutation relations in this case imply (3.48) all other commutators being zero. To express these relations in terms of creation and annihilation operators associated with the complex field , let us define which mean 1 . a(p) = j2(a, (p) +W2(p)), alp) = ~(al(p) - ia2(p)), at(p) = ~(at(p) - ia~(p)), at(p) = ~(at(p) + ia~(p)) (3.49) . (3.50) (3.51) We can now start with Eq. (3.18) for both 1 and 2 and use the definition of Eq. (3.46) to write the Fourier expansion of the field : (x) = J d 3 p (a(p)e-ip.X +at(p)eiP'X ) , )(27[)32Ep which also implies (3.52)
  48. 50. 3.6. Complex scalar field 39 Moreover, the commutation relations of Eq. (3.48) can be cast into the form (3.53) whereas all other commutators vanish. Since at and a~ create the quanta of the fields (p, and (/>2 which are different, it is clear that there are two different particles in the theory. Since Eq. (3.53) presents the commutation relations in the form of Eq. (3.22), we can interpret at and at to be the creation operators for these two particles. Of course, we could have chosen any two orthonormal combinations of at and a2, even at and a2 themselves, as the annihilation operators for our two particles. There is a special reason for quantizing and t in terms of a and a, as we shall now see. 3.6.2 Particles and antiparticles The Lagrangian of Eq. (3.45) is invariant under the transformation _ e-"o, t _ e"Ot, (3.54) where IJ does not depend on space-time. Following the procedure outlined in 2.4, we can calculate the Noether current for this invari- anee. For infinitesimal (), we have o = -iqlJ, ot = iqlJt . So, a~. a~. t J~ = a(a~) (-.q) + a(a~t) (.q ) = iq [(~)t - wt)] The conserved charge Q is therefore given by Q = Jd3 x JO = Jd3 x iq [(ao)t - (aot)] = qJd3 p [at(p)a(p) - at(p)a(p)] (355) (3.56) (3.57) The first term here, remembering the definition of Eq. (3.43), is just the number operator for the quanta created by at. Let us denote it by Na Similarly, the second term is N. Thus we obtain Q = q(Na - No.) (3.58)
  49. 51. 40 Chapter 3. Quantization of scalar fields Notice the negative sign for the second term. If q is called the charye associated with each quanta created by at, the first term is the total charge in such quanta. The quanta created by iit then possess an opposite charge. These are called the antiparticles for the particles which are created by at. Eq. (3.58) then tells us that the total charge in particles and antiparticles is conserved as a consequence of Noether's theorem. o Exercise 3.1 Write Q in terms oj a,(p),a!(p),a,(p) and a~(p) to uerif'y that it cannot he 'Written in terms of number overotoT' of the fields , and ,. Looking back at Eq. (3.51), we thus see that the operator "'(x) can do two things. The first term describes the annihilation of a particle, the second one the creation of an antiparticle. In either case, if "'(x) operates on any state, it will reduce the charge of the state by an amount q. Similarly, Eq. (3.52) implies that the operator ",t(x) can either create a particle, or annihilate an antiparticle. In other words, if it operates on a state, the charge will increase by q. One may wonder at this point as to how the quantity q is defined. After all, in Eq. (3.54) where it first shows up, q is always multiplied by B, so we can change the value of q by redefining B with a multi- plicative constant. But there is nothing mysterious about this. Aft~r all, any charge is defined only up to a multiplicative constant. For example, take the case of electric charge. If we multiply the values of the charges of all particles by a constant, the law of charge conser- vation will not be affected at all. The Maxwell's equations will also be unaffected, since the electromagnetic fields will be multiplied by the same constant. So, if we consider only one kind of particle in i~ lation, the value of the charge is really arbitrary. Only the ratios of charges of different particles have some physical meaning, and these ratios can be determined only when different kinds of particles are present and there is some interaction among them. Such instances will be encountered when we will discuss interactions in the later chapters. 3.6.3 Ground state and Hamiltonian We can now construct the states of the Hamiltonian, following the procedure outlined in 3.4 and 3.5. The vacuum, or the ground
  50. 52. 3.7. Propagator state, is defined in this case by the relation a(p) 10} = ii(p} 10} = 0, for all p. 41 (3.59) In other words, the vacuum is the state which contains no particles and no antiparticles either. The state Ip} defined in Eq. (3.39) will be a state containing one particle with energy Ep . Similarly, we can define a state (3.60) which will contain one antiparticle with energy Ep . The states with more than one particles or antiparticles can be constructed similarly. The normal-ordered Hamiltonian can now be written as (3.61 ) o Exercise 3.8 Find the Hamiltonian in terms of the field t:P 'Using the Lagrangian of Eq. (3.45). Use the Fourier decomposition of the jield giuen in Eq. (3.51) Qnd the 'Prescription of normal ordering to express it in the Jorm oj Eq. (3.61). 3.7 Propagator The goal of Quantum Field Theory is to describe particle interac- tions. Our interest will be in c81culating cross-sections, transition probabilities, decay rates etc. For this, we need to know how parti- cles move in space-time. We shall use Green's function techniques for this and define a propagator, which will be useful in later chapters. Let us start with the Klein-Gordon equation with a source term: (0 +m2 ) (x) = J(x). (3.62) Since the Klein-Gordon equation is the same whether the scalar field is real or complex, the discussion in this section is equally applicable for both cases. In order to solve this equation, we first introdnce a propagator, or Green's function, denoted by G(x - x'), which satisfies the equation (ox +m2 ) G(x - x') = - c5'(x - x') (3.63)
  51. 53. 42 Chapter 3. QUantization of scalar fields (3.65) Here the notation Ox indicates that the derivatives in the operator have to be taken with respect to the co-ordinates of x, not of x' Clearly, if this equation can be solved, we can represent the solution of Eq. (3.62) by 4>(x) = 4>o(x) - !d"x' G(x - X')J(X'), (3.64) where 4>o(x) is any solution of the free Klein-Gordon equation, Eq. (3.3). The boundary conditions are matched by choosing 4>0 appro- priately. Now, to solve for the propagator, we introduce its Fourier trans- form by the equation G( _ ') =! d 4 p -;p.(x-x')G() x x (2,,)4 e p , where all the components of pi' should be treated as independent variables, not related by the energy-momentum relation of Eq. (3.1). If we act on this equation by the operator appearing in Eq. (3.63), we obtain Remembering that ,j4( _ x') =! d"p e-;p(x-x') x (2,,)4 ' and comparing with Eq. (3.63), we obtain (3.67) 1 G(p) = 2 2 P -m 1 (3.68) where Ep was defined in Eq. (3.17). This expression, however, has an ambiguity. Let us put this ex- pression into Eq. (3.65) and try to perform the integration over pO. The integrand has poles at pO = Ep, and therefore the integration cannot be performed over the real axis for pO. This has nothing to do with quantum theory, since we have not used anything regarding quantization in this section so far. The problem lies even with the classical theory.
  52. 54. 3.7. Propagator 43 To get around this problem, we use a prescription by Feynman and complexify the propagator. Instead of the Green's function of Eq. (3.68), from now on we will use 1 AF(p) = 2 2 +.,p -m 1.E 1 (pO)2 _ (Ep _ i)2 ' (3.69) where , ' are infinitesimally small positive parameters, obviously related to each other by,' = 2Ep'. Of course, has to be taken to zero at the end of calculations. For this reason, we will ignore the distinction between ' and E. Introduction of this parameter makes the propagator complex, and hence inherently non-classical. Let us now rewrite Eq. (3.69) in the form A F (p) = _1_ [ 1 _ ,,-,--,-~I_-:-;-] 2Ep pO - (Ep - i,) pO + (Ep - i,) (3.70) (3.72) (3.73) (3.74) This expression can be put into Eq. (3.65), which gives J d3p e+'p-(z-z') /+00 dpo . 0AF(x - x') = - - - .-.p(,-,')(2..)3 2Ep -00 2.. x [(PO-~p)+i' - (pO+L)-i] (3.71) To perform the integrations over pOI we need the following result from complex analysis: / +00 e-i(t lim de . = -21ri8(t), _0 -00 (+ IE where the 8-function was defined in Eq. (3.13). For the first term in Eq. (3.70), we can shift to the variable pO' = pO - Ep to perform the pO-integration: / +00 dpot e-i(pOI+Ep)(t-t') -- , . = -i8(t - I').-'E,(,-,'} -00 27r pO + IE Similarly, for the other term, we can substitute pO' = _po - EPI which gives for this term / +00 dpoJ e-i(pOI +Ep)(tl-t) -- = i8(t' - I).-'E,(,'-') -00 27T _po' - if
  53. 55. 44 Chapter 3. Quantization of scalar fields Combining the two t.erms then, we can write tJ. (x - x') = -i J d 3 p lett _ t')e-'Ep(t-t')+'P'("-"') F (21f)32Ep +e(t' - t)eiE,(t-t')+il>(,,-,,'l] . (3.75) Changing the sign of the integration variable p in the second term, this can be written as [e(t - t')e-ip.(x-x'l +8(t' _ t)eiP'(x-x')] , (3.76) (3.78) (3.79) where now the components of the 4-vector pi' are not independent, but are related by pO = Ep . o Exercise 3.9 Show that tJ.p(x - x'), given in Eq. (3.76), .ati.jie. the equation (0, + m2 ) tJ.F(x - x') = - .'(x - x'). (3.77) It is therefoTe a. Green's Ju.nction. This form can be easily related to the quantized fields. For the sake of illustration, let us use the real scalar field. From Eq. (3.18), we can write (x) 10) = J d 3 p at (p)e'PX 10) J(21f)32Ep -J d 3 p e'P'x Ip) - J(21f)32Ep , where Ip) is the one-particle state defined in Eq. (3.39). Notice that the annihilation operator in (x) does not contribute here owing to Eq. (3.32). On the other hand, (Ol(x') = J d 3 p' e-'p'x' (p'l , J(21f)32Ep' where only the annihilation operator part of the Fourier decomposi- tion of (x') is important. Using the normalization of Eq. (3.40), we thus obtain (0 1(x')(x)1 0) = J d 3 p e'P'(x-x') (21f)32Ep (3.80)
  54. 56. 3.7. Propagator 45 Thus, we can rewrite Eq. (3.76) in the following form using the field operators: itlF(X - x') = 8(t - t') (0 I>(x)>(x')i 0) +8(1' - t) (0 I>(x')>(x) I0) , which can be abbreviated to itlF(X - x') = (01&1 [>(x)>(x')] I0) , (3.81) (3.82) where &1 [...J implies a time-ordered product, which means that the operator with the later time must be put to the left of the operator with the earlier time. In general, for any two scalar operators A(x) and B(x), the time-ordered product is defined as , _ { A(x)B(x') &1 [A(x)B(x )J = B(x')A(x) ift>t', ift'>t. (3.83) Let us now see how the form given in Eq. (3.81) can be inter- preted. First, suppose t < t', in which case only the second term survives. From the expression in Eq. (3.78), we see that the operator >(x) operating on the vacuum state to the right creates a particle at time t. On the other hand, for (01 >(x'), as commented earlier, only the annihilation operator part contributes, so this part describes the annihilation of a particle at time t'. Thus the matrix element de- scribes the amplitude of a particle being created at a time t, which propagates in space-time and is annihilated at a later time t'. Simi- larly, if t > t', the first term of Eq. (3.81) describes the amplitude of a particle being created at a time t', which propagates in space-time and is annihilated at a later time t. This, in a sense, is the phys- ical meaning of the propagator, which also shows why it should be important in calculating the amplitudes of various processes involv- ing particle interactions. After all, between interactions, the particles just propagate in space-time! The mathematical formulation of these comments will be the subject of Ch. 6. o Exercise 3.10 'or the comple:>: srotar jield, ShOUl that Eq. (3.76) can be written as it.F(X - x') = (Olg [>(x)>t(x')] I0) (3.84) a) Argue that it represents the propngation oj a. partide from t' to t aT the propugation of an antiparticle from t to t'.
  55. 57. 46 Chapter 3. QUaIjtization of scalar fields b) ShOUl that the pTopagatoT ca.n o1so be 'Written (x')tI>t(x)] 10) (3.85) Describe this in tenns oj pa.rtiete and a.ntiparticle propaga.tion.
  56. 58. Chapter 4 Quantization of Dirac fields In Ch. 3, we dealt with the quantization of scalar fields. Such fields, by definition, are invariant under Lorentz transformations. In this chapter, we try something more complicated - fields which describe spin- ~ particles. 4.1 Dirac Hamiltonian The problem with the one-particle interpretation of the Klein-Gordon equation comes from the fact that we encounter negative energies for free particles. We showed how this problem is solved for scalar fields. Dirac tried an alternative solution to this problem, one that led to the correct description of spin- ~ particles. He observed that this problem did not arise in non-relativistic quantum mechanics because the Schrodinger equation was linear in the time derivative. So he tried to construct a Hamiltonian which, in a sense, would represent the square root of the equation (4.1) where H is the Hamiltonian operator and p is the operator for 3- momentum. In coordinate space these operators can be written as derivative operators, H = iBt, P = -iV, as we know from their commutation properties. Dirac wanted an operator linear in the components of momentum which would square to p2 +m2 . But such an operator cannot be constructed using only numbers and functions as coefficients. So Dirac assumed that the square root of this equation 47
  57. 59. 48 should be of the form Chapter 4. Quantization of Dirac fields H =l' = 4gjw , 1'1"1"1.1' = -21p1"1,. (4.42) 4.3 Plane wave solutions of Dirac equation 4.3.1 Positive and negative energy spinors For a single particle interpretation, the t/J(x) appearing in Eg. (4.22) would be a wave function. Although Dirac set forth to avoid the negative energy solutions, the irony of the situation is that the Dirac equation also has such solutions, just like the Klein-Gordon equa- tion. To see that, let us assume that we are trying a plane wave solution in the rest frame of a particle. The time dependence of this

slideshare_object._adQueue.push({ tile : 5, zone : 'slideshare', dart_code: '', width: 300, height: 250, appendTo: 'topRightAd'}); Recommended

  • PowerPoint Tips Weekly Online Course - LinkedIn Learning

Teaching Complex Topics Online Course - LinkedIn Learning PowerPoint Tips and Tricks for Business Presentations Online Course - LinkedIn Learning Numpy cookbook - ivan idris Arindam Kumar chatterjee Digital Electronics Vodafone Albania The latex companion 2ed - Frank mittelbach Arindam Kumar chatterjee Problems and solutions statistical physics of particles - Mehran kardar Arindam Kumar chatterjee Statistical physics of particles - Mehran kardar Arindam Kumar chatterjee Hands on start to wolfram mathe - cliff hastings Arindam Kumar chatterjee Gnu octave beginner's guide - jesper schmidt hansen Arindam Kumar chatterjee

  • English

Espaol Portugus Franais Deutsch

  • About

Dev & API Blog Terms Privacy Copyright Support

LinkedIn Corporation 2018 Share Clipboard Email Email sent successfully..

  • Facebook

Twitter LinkedIn Link Public clipboards featuring this slide No public clipboards found for this slide Save the most important slides with Clipping Clipping is a handy way to collect and organize the most important slides from a presentation. You can keep your great finds in clipboards organized around topics. Start clipping No thanks. Continue to download. Select another clipboard Looks like youve clipped this slide to already. Create a clipboard You just clipped your first slide! Clipping is a handy way to collect important slides you want to go back to later. Now customize the name of a clipboard to store your clips. Name* Description Visibility Others can see my Clipboard Cancel Save