You are on page 1of 225
FINITE-TEMPERATURE FIELD THEORY JOSEPH I. KAPUSTA Professor of Physics, University of Minnesota, Minneapolis CAMBRIDGE UNIVERSITY PRESS : Cambridge New York New Rochelle Melbourne Sydney Published by the Press Syndicate of the University of Cambridge ‘The Pitt Building, Trumpington Street, Cambridge CB2 1RP 32 East 57th Street, New York, NY 10022, USA 10 Stamford Road, Oakleigh, Melbourne 3166, Australia © Cambridge University Press 1989 First published 1989 Printed in the United States of America Library of Congress Cataloging-in-Publication Data Kapusta, Joseph I. Finite-temperature field theory / Joseph I. Kapusta. p. cm. ~ (Cambridge monographs on mathematical physics) Bibliography: p. Includes index. ISBN 0-521-35155-3 1. Quantum field theory. 2. Particles (Nuclear physics). 3. Many. -body problem. L Title. Il. Series. QC793.3.F5K36 1989 530.14 - del? 88-21653 CIP British Library Cataloguing in Publication Data Kapusta, Joseph I. Finite-temperature field theory. 1. Particles, Relativistic quantum theory L Title 539.721 ISBN 0 521 35155 3 11 12 13 14 15 1.6 Contents Preface Review of quantum statistical mechanics Ensembles One bosonic degree of freedom One fermionic degree of freedom Noninteracting particles in a box Bibliography Exercises Functional integral representation of the partition function Transition amplitude for bosons Partition function for bosons Neutral scalar field Bose-Einstein condensation Fermions Remarks on functional integrals Bibliography Exercises Interactions and diagrammatic techniques Perturbative expansion Diagrammatic rules for A¢* theory Propagator First-order corrections to II and In Z Summation of infrared divergences Yukawa theory Remarks on real-time perturbation theory Bibliography Exercises Page ix OowNANH SE vi 4 41 4.2 43 44 45 5 5.1 5.2 5.3 5.4 5.5 5.6 5.8 61 6.2 63 64 65 6.6 71 13 14 15 16 8.1 8.2 8.4 8.5 Contents Renormalization Renormalizing A¢* theory Renormalization group Application to the partition function Bibliography Exercises Quantum electrodynamics Quantizing the electromagnetic field Blackbody radiation Di tic . Photon self-energy Loop corrections to In Z 5.5.1 Two loops 5.5.2 Ring diagrams 5.5.3 Three loops White dwarf stars Bibliography Exercises Linear response theory Linear response to an external field Screening of static electric fields Exact formula for screening length in QED Plasma oscillations Bibliography Exercises Spontaneous symmetry breaking and restoration Charged scalar field with negative mass-squared Goldstone’s theorem Loop corrections Higgs model “Bibliography Exercises Quantum chromodynamics Quarks and gluons Asymptotic freedom . Perturbative evaluation of partition function Instantons . Gluon propagator and linear response 101 101 107 115 118 118 121 125 131 135 140 Al A2 A3 145 Preface x Apart from a brief reference in Chapter 8, I do not discuss lattice gauge theory. This is due partly to lack of space, partly to my limited knowledge, and partly to the even more highly technical nature of that topic. I would encourage those who wish to study lattice gauge theory further first to read this monograph and then to go on to one devoted exclusively to lattices, of which there are now several. I found, as many authors probably do, that it is not always an easy task to transform notes into text. I am thankful to a number of friends and colleagues for their helpful suggestions, critical comments, and careful Teading of the manuscript, especially P. Ellis, S. Gasiorowicz, U. Heinz, K. Kajantie, P. V. Landshoff, and T. Toimela. I wish to thank E. Heide for preparing Figures 10.1 to 10.3 and N. K. Glendenning for preparing Figures 10.4 and 10.5. Finally, I thank S. Smith for faithfully typing this manuscript. J. Kapusta Minneapolis, 1988 1 Review of quantum statistical mechanics 1.1 Ensembles In equilibrium statistical mechanics, one normally encounters three types of ensemble, The microcanonical ensemble is used to describe an isolated system which has a fixed energy E, a fixed particle number N, and a fixed volume V. The canonical ensemble is used to describe a system in contact with a heat reservoir at temperature T. The system can freely exchange energy with the reservoir. Thus T, N, and V are fixed. In the grand canonical ensemble, the system can exchange particles as well as energy with a reservoir. In this ensemble, 7, V, and the chemical potential » are fixed variables. In the latter two ensembles T~! = 8 may be thought of as a Lagrange multiplier which determines the-mean energy of the system. Similarly, » may be thought of as a Lagrange multiplier which determines the mean number of particles in the system. In a relativistic quantum system where particles can be created and destroyed, it is most straightforward to compute observables in the grand canonical ensemble. So we therefore use this ensemble in this monograph. This is without any loss of generality, since one may pass over to either of the other ensembles by performing a Laplace transform on the variable » and/or the variable £. (See Appendix A.3.) Consider a system described by a Hamiltonian H and a set of conserved number operators N,. In relativistic QED, for example, the number of electrons minus the number of positrons is. a conserved quantity, not the electron number or positron number separately. These number operators must be Hermitian and must commute with H as well as with each other. Also, the number operators must be extensive in order that the usual macroscopic thermodynamic limit can be taken. The statistical density matrix is B= exp[-8(H - 2.%)] (1.1) where a summation over i is implied. The ensemble average of an 2 _ 4. Review of quantum statistical mechanics operator Ais Trpd A= Ts (1.2) The grand canonical partition function is Z=Trp : (1.3) The function Z = Z(V, T, Hy #2)---) is the single most important func- tion in thermodynamics. From it all other standard thermodynamic properties may be determined. For example, the pressure, particle num- ber, entropy, and energy are, in the infinite volume limit, P= pinz N poz awe’ Op, * a(TinZ g = HER?) E=-PV + TS + p,N, 4) 12. One bosonic degree of freedom As a simple example consider a time-independent single-particle quan- tum mechanical state may be occupied by bosons. Each boson in that state has the same ‘w. There may be 0, 1, 2, or any number of bosons in that state. There are no interactions between the particles. This system may be thought of as a quantized simple harmonic oscillator. It will serve as a prototype of the relativistic quantum field theory systems to be introduced in later chapters. We are interested in computing the mean particle number, energy, and entropy. However, since the system has no size or “ volume,” physically speaking there is no pressure. Denote the state of the system by |”), which means that there are n bosons in the system. The state |0) is called the vacuum. Properties of these states are (al) = By orthogonality (1) E In)dol = 1 completeness 6) One may think of the bras (n| and kets |n’) as row and column vectors, respectively, in an infinite-dimensional vector space. These vectors form 1.2 . One bosonic degree of freedom 3 a complete set. The operation in (1.5) is an inner product and the 1 in (1.6) is the infinite-dimensional unit matrix. It is very convenient to introduce creation and annihilation operators, at and a. The creation operator creates one boson. Its action on a number eigenstate is atin) = (n + 1){n +1) (1.7) Similarly, the annihilation operator annihilates one boson, a|n) = nt[n -— 1) (1.8) unless n = 0, in which case it annihilates the vacuum, . a|0) =0 (1.9) Apart from an irrelevant phase, the coefficients appearing in (1.7) and (1.8) follow from the requirements that at and a be Hermitian conju- spies and that aa be the number operator Fi That is, Nin) = ata|n) = n\n) (1.10) It follows that the commutator of a with at is aat — ata = [a, at] =1 (1.11) We can build all states |”) from the vacuum 10) by repeated applica- tions of the creation operator In) = —| (at) 10 (1.12) 1 (at) Next, we need to have a Hamiltonian. Up to an additive constant, it must be w times the number operator. Starting with a wave equation in nonrelativistic or relativistic quantum mechanics, one usually is moti- vated to write H = ha(aat + ata) = o(ata+3)=0(N+4) (1.13) The additive term 4w is the zero-point energy. Usually this term can be ignored. Exceptions arise when the vacuum changes due to a background field, such as a gravitational field, or an electric field as in the Casimir effect. We shall ignore this term here and take H = wN. The states |n) are simultaneously number and energy eigenstates. Therefore we can assign a chemical potential to the particles. The 4 1. Review of quantum statistical mechanics partition function is easily computed Z = Tre~PH—e®) we Tre Pew h ey 2 =z (ne Fo ny = De few a=0 n=0 1 eo = [lien ‘ (1.14) ‘The mean number of bosons is found from (1.4) to be 1 * few] and the mean energy is E = wN. Note that N ranges continuously from zero to infinity. N (1.15) _. 71.3 One fermionic degree of freedom - Now consider the same problem as above except with fermions instead of bosons. This will be thé prototype for a Fermi gas, and will later on help us to formulate the functional integral expression for the partition function involving interacting fermions. These will include electrons and positrons in QED, quarks in QCD, and dense nuclear matter. The Pauli exclusion principle forbids the occupation of a single-particle state by more than one fermion. Thus there are only two states of the system, |0) and |1). The action of the fermion creation and annihilation operators on these states is at}0) = |1), all) = 0), at{1)=0, a|0)=0 (1.16) Thus, these operators have the property that their square is zero, aa = atat = 0 (1.17) Up to an arbitrary phase factor, the coefficients in (1.16) are chosen so that a and at are Hermitian conjugates and that a’« is the number operator N: N{0) = atal0) = 0, NO) = ate|1) = (1) (1.18) It follows that the creation and annihilation operators satisfy the anti- commutation relation ‘ aat + ata = {a, at} =1 (1.19) 4.4 Nonwnteracting parucies in a box > The Hamiltonian is taken to be H = 1a(ala — aat) = o(N — 1) (1.20) Notice that the zero-point energy is equal in magnitude but opposite in sign to the bosonic zero-point energy. In the following discussion we drop this term (see Section 1.2). The partition function is computed as in (1.14) except that the sum terminates at n = 1: 1 Ze Tre~P 1M) we YY (nfo Min) mw 1 +e A-*) (1.21) n=0 The mean number of fermions is 1 N= Gena (1.22) Note that N ranges from 0 to 1- The mean-energy is wN. 1.4 Noninteracting particles in a box Put identical particles, either bosons or fermions, in a box with sides of length L. We neglect their mutual interactions, although in principle they must interact in order to come to thermal equilibrium. One can perform a gedanken experiment by first allowing them to come to equilibrium and then slowly turning off the interactions. Such a simple thermodynamical description is often useful for electrons in a metal, blackbody photons in a heated cavity, phonons at low temperature, neutron matter in a neutron star, and many other examples. For defi- niteness we impose the boundary condition that the wavefunctions vanish at the surface of the box. This means that an integral number of half-wavelengths must fit in the distance L, A, =2L/l,, Ay=2IL/l,, d,=2L/l, (1.23) where /,,/,, 1, are all positive integers. The magnitude of the x compo- nent of the momentum is |p,| = 27/A, = a/,/L, and similarly for the y and z components. Amazingly, quantum mechanics tells us that these relations hold for both nonrelativistic and relativistic motion, for both bosons and fermions. The full Hamiltonian is the sum of the Hamiltonians for each mode. We use the shorthand notation where i represents the triplet of numbers 6 1. Review of quantum statistical mechanics (1, 1,, 1,) which uniquely specify each mode. Thus H=YH, N= IN (1.24) i i Due to (1.24), the partition function, Z, is the product of the partition functions for each mode: Z= 12, _ (1.25) Each mode corresponds, of course, to the single bosonic or fermionic degree of freedom discussed previously. According to (1.4), it is In Z which is of most direct application. From (1.25), mz- 5 EF iz,,, (1.26) Gealieal gad In the macroscopic limit L —> 00, i.e., when L becomes large compared to all other physical length scales, it is permissible to replace the sum from /, = 1 to © with an integral from /, = 0 to oo. (The first-order correction to In Z given in (1.27) is proportional to the surface area L?,) Also, di, = Ld|p,|/2. Hence, — InZe= z fav. [Paws [svimz, - 2 (127) In all cases to be dealt with here the mode partition function depends only on the magnitude of the momentum components. Then the integra- tion over p, can be extended from —co to +o if one divides by two. a nz=v{f P nz, (1.28) (22) Note the appearance of the total phase space integral fax @p/Qa)? in (1.28). ik ene InZ=V le — yin + e7Ae-m)#t (1.29) where the upper sign (+) refers to fermions and the lower sign (—) 1.4 Noninteracting particles in a box 7 refers to bosons. From (1.4) and (1.29) we obtain P=Tmnz/v "fay ar rT +1 EnV aT =H +1 (1.30) The formulas for N and E have the simple interpretation of being phase space integrals over the mean particle number and mean energy of each mode, respectively. The dispersion relation w = w( p) furnishes the connection between p and w. For relativistic particles = (p? + m?)!, which has the nonrela- tivistic limit w = m + p?/2m. (The chemical potential in our convention is always measured with respect to the total particle energy including the mass. The chemical potential yj,, aS customarily defined in nonrela- tivistic many-body theory, is related to our p by py, =m — m.) For phonons, the dispersion relation is w = c,p, where c, is the speed of sound in the media. . For bosons which carry no conserved. charge, such as photons or 2° mesons, » = 0. For electrically charged bosons, such as * and w~ mesons, we can associate a chemical potential with the electric charge. The w* and =~ form a particle—aitiparticle pair. The +* has charge plus one and chemical potential y, while ~ has charge minus one and chemical potential —y. (Antiparticles have a chemical potential of opposite sign to particles.) The total conserved charge is 1 on vf (sent — —1 - = (131) and the total energy is @p o o - "lop (ROLE +aenrz) 0) If the bosons have nonzero spin s, then the phase space integrals must be multiplied by the spin degeneracy factor 2s + 1. An analogous discus- sion may be given for fermions. % 4. sxevIeW OF quantum statsuca: mecnanics 1.5 Bibliography Thermal and statistical physics: Reif (1965) Landau and Lifshitz (1959) Many-body theory: Abrikosov, Gorkov, and Dzyaloshinskii (1963) Fetter and Walecka (1971) 1.6 Exercises 1.1. Prove that the state |) given in (1.12) is normalized to unity. 1.2 Referring to (1.16), let |0) and |1) be represented by the basis vectors in a two-dimensional vector space. Find an explicit 2 x 2 matrix representation of the abstract operators a and at in this vector space which satisfy (1.16), (1.17), and (1.19). 1.3 Calculate the partition function for noninteracting bosons including the zero-point energy. From that ‘partition function calculate the mean energy, particle number, and entropy. Repeat the calculation for noninteracting fermions. 2 Functional integral representation of the partition function The customary approach to nonrelativistic many-body theory is to proceed with the method of second quantization begun in the first chapter (Fetter and Walecka 1971). There is an alternative approach, the method of functional integrals, which we shall follow. Of course, what can be done in one formalism can be done in another. Nevertheless, functional integrals seem to be the method of choice for most elementary particle theorists nowadays, and they seem to lend themseives more readily to nonperturbative phenomena like tunneling, instantons, lattice gauge theory, etc. For gauge theories they are practically indispensible. Unfortunately, there is a certain amount of formalism which must be developed before we can discuss physical applications. In this chapter, we will derive the functional integral representation of the partition function for interacting relativistic nongauge field theories. As a check on the formalism, and to obtain some feeling for how functional integrals work, we will rederive some well-known results on relativistic ideal gases of bosons and fermions. 2.1 Transition amplitude for bosons Let $(x,0) be a Schridinger-picture field operator at time ¢ = 0 and let #(x, 0) be its conjugate momentum operator. The eigenstates of the field operator are labeled as |) and satisfy (x, O)l6) = o(x)1¢) (2.1) where $(x) is the eigenvalue, actually a function of x. We have the usual completeness and orthogonality conditions fde@ier
    (2.34) Here D = B?(— 97/1? — vy? + m?) in (x, 7) space and D = («2 + «) in (p, @,) space, and (¢, Dg) denotes the inner product on the function space. The formula (2.34) follows from the following formula for Riemann integrals (with a constant matrix D): f * dx, +++ dx, e727 = (#)"(det D)? (238) Use of (2.34) also leads to (2.33). Thus far we have InZ= -3Y Din[A?(o? + o?)] (2.36) =P Notice that In[(2en)? + 6207] = -("" oon + In[1 + (2an)"] (2.37) The last term in (2.37) is B-independent and thus can be ignored. 16 2. Functional integral representation Furthermore, « 1 _. n+ (6/22) + -1 zi) (2.38) nzZ= ~Z fae; + = i) (2.39) Carrying out the @ integral, and throwing away a f-independent piece, we finally arrive at z= rf otei-ise - in(1 - e-**)] (2.40) This is identical with the bosonic version of (1.29), with = 0, except that (2.40) includes the zero-point energy. Both a ap : By~ ~ gg Vf ase (2.41) and _@ Aya Ts; nZy= —E/V (2.42) should be subtracted, since the vacuum is defined: to be that state with zero energy and pressure. 2.4 Bose-Einstein condensation A more interesting system is obtained by considering the theory of a charged scalar field ©. The field © is a complex field which describes bosons with positive and negative charge, i.e. they are antiparticles of each other. The Lagrangian density is L= 3,0°9") — m>O*d - A(@*0)? (2.43) This theory has an obvious U(1) symmetry, {8 > 0 = bem (2.44) where a is any real constant. This is a global symmetry since the field (x) is multiplied by a phase factor which is the same at all points in spacetime. 2.4 Bose-Einstein condensation 17 It was shown by Noether that with each continuous symmetry of the Lagrangian there is an associated conserved current. In this case, the current may be found by letting a depend on x and then considering it as an independent “field”. Under the transformation (2.44), LL = A, (O* l)) a"(He-) — m2" — A(H*O)? = L+ 0°0,0 "a + 18,0(b*9"d — 9H") (2.45) The equation of motion for the “field” a(x) is 0L' OL" (2.46) Pn 0(d%) da Since 0.£'/da = 0, it follows that the “current” 0.2'/0(%a) = o°00,0 — i9,@* + i@*0,® is conserved. We can recover our original theory without the “field” a(x) simply by setting a = constant. Hence we have found the conserved current density jy = (099,06 — $9,0*) ~ ay, =O (2.47) This conservation law may be verified independently by using the equation of motion for ®. The full current is J, = fd°xj,(x) and the conserved charge is Q. = { d°x jo(x). It is convenient to decompose ® into real and imaginary parts, ® = (4, + ig,)/2!, The conjugate momenta are-m, = 99,/0t and m= 9,/8t. The Hamiltonian density and charge are easily found to be om 3 fn? + a2 + (Vos)? + (V4)? + m4 + m3] +(e + 8) (2.48) and Q= [ &x(ex, - 91m) (2.49) ‘The partition function is Z= fldlanl f_ [aaltaeslen| (ar farx(in, a tim, =~ * — (My My dy 2) + w( Gam — #))| (2.50) where we have inserted-a chemical potential » associated with the 18 2. Functional integral representation conserved charge Q. The integration over the conjugate field momenta can be done following the example of the neutral scalar field. The result is oy > Z= (Pf [ablldtalem (Par fox [-: (F - ine.) a -(= + ints) — H(vos)? — 2(¥6,)? — dnt —dmigt — 2A(93 + a)| (2.51) where N’ is the same divergent normalization factor. Notice that the argument of the exponential in (2.51) differs from one’s naive expecta- tion of 2( $1, $2, by Ib25 # = 0) + Hjo(Pr $21 991/97, i 94/97) by an amount p’@*®. This is due to the momentum dependence of jg. The expression (2.51) cannot be evaluated in closed form unless A = 0. Inthe latter case, the functional integral becomes Gaussian and can be worked out analogously to the free neutral scalar field. The components of ® can be expanded in a Fourier series, = 2! cos 6 + (=) LE errr, a(P) = saosaee( Sy eextg, (Pp) (2.52) . ET Here { and @ are independent of (x,r) and carry the full infrared character of the field; that is, $,,9(p = 0) = ¢,9(p = 0) = 0. This allows for the possibility of condensation of the bosons into the zero-momen- tum state. Condensation means that in the infinite volume limit a finite fraction of the particles reside in the n = 0, p = 0 state. Setting A = 0, and substituting (2.52) into (2.51) after an integration by parts (see (2.29), we find 2 = (W|TTEL J (0 40, (| 2 aye _ _ $1, a(P) S- BV(u? — m?)¢ ALL (1: -of PDs ba; -al po(e9)] 2 _ 2 _ 2 + w? - 2po, (2.53) D = p? a 2po, w+ w? - p? 2.4 Bose-Einstein condensation 19 Carrying out the integrations, In Z = BV(u? — m?)¢? + In(det D)* (2.54) The second term can be handled as follows: -1indet D = ~4i0(T1TTA*[ (a2 +ot— p+ 4uu3]} ~ ~tin{ FT TT8"[ 02 + (0 - #4} -rm(TTTTs%[ot+(o+ay]} 55) Putting these together, In Z = BV(u? - m?)g? — 29) In{ B*[o2 + (w - 2)']} ne -FDL inf 6?[o2 + (0 + w)]} (2.56) The last two terms in (2.56) are precisely of the form of (2.36). All we need to do is recall (2.40) and make the substitutions w > w — » and w > w + p for the two terms in (2.56), respectively. We obtain In Z = BY (ut — m?)g? Vf pleat — aA (257) There are several points to note about (2.57). The momentum integral is convergent only if |u| FS) nam + or(B n= m) ap 1 1 = [Eo gemrg - wer) (259) (The case 4 = ~m is handled analogously.) Here the separate contribu- tions from the condensate (the zero-momentum mode) and the thermal particle excitations (finite-momentum modes) are evident. If the density p is held fixed and the temperature is lowered, » will increase until the point p = m is reached. If the. temperature is lowered even further, then p*(B, » = m) will be less than p. Therefore { is given by 2 P= eB, = m) p.foeecm 0.60) when p=m and T< T.. ‘The critical temperature T, j is determined implicitly by the equation p= p*(B.. p= m) (2.61) In the nonrelativistic limit, one finds the well-known result 297 pe \i n=S(5 7 ) > p m? (2.63) In the limit that m — 0, also |p| — 0, and T, - co. When m = 0 then all of the charge resides in the condensate, at all temperatures, and none is carried by the thermal excitations. There is a second-order phase transition at T,. This can be shown rigorously by a careful examination of the behavior of the chemical potential »(p, T) as a function of T near T, with p fixed. This analysis is left as an exercise. A more intuitive way to see this is to invoke the general Landau theory of phase transitions (Landau and Lifshitz 1959). The order parameter { drops to zero continuously as T. is approached from below and remains zero above T,. Physically the reason for a phase transition is as follows. At T = 0, all of the conserved charge can reside in the zero-momentum mode on account of the bosonic character of the 2.5 Fermions 21 particles. (This is not possible for a charge carried by fermions.) As the temperature is raised, some of the charge is excited out of the conden- sate. Eventually the temperature becomes great enough to completely melt, or thermally disorder, the condensate. The phase transition is second order because there is no reason for { to drop to zero discontinu- Although the subject of Bose-Einstein condensation is an interesting and illustrative one, the reader should be reminded that Bose-Einstein condensation of ideal gases has never been observed. The presence of particle interactions, no matter how weak, qualitatively alters many of the aspects of the condensation phenomenon (Bogoliubov 1947). This includes the behavior of the specific heat near T, and the dispersion relation for collective excitations. 2.5 Fernions . ‘All discussions in this chapter ‘have so-far referred-to bosons. We now turn our attention to (Dirac) fermions. In relativistic quantum mechanics one finds that electrons or muons are described by a four-component spinor . The components are called y, where a runs from 1 to 4. The motion of free fermions can be described by a wavefunction ¥(x,1) = (1/VA)EL(m/e)! . ps x[b(p, s)u(p, s)e7”* + d*(p,s)o(p,s)e'?*] (2.64) Here u and v are the positive and negative energy plane-wave spinors, respectively. The sum on s refers to the two possible spin orientations of the spin-} Dirac fermion. The expansion coefficients b(p,s) and d*(p,s) are complex functions in relativistic quantum mechanics but become operators in a field theory. As usual, p - x = p'x, = Et — p*x. Usually (2.64) is normalized as fev ues) = DL [l6(2, 5) +14(2,5)F] =1 (265) Pa In the absence of interactions the Lagrangian density is L=V(id-m)y (2.66) The Dirac matrixes, y*, which are defined by (y*,y”} = 2g", are in 22 2, Functional integral representation standard convention (29) eon Each of these is a 4X 4 matrix, with 1 denoting the 2 x 2 identity matrix and o the triplet of Pauli matrixes. In (2.66), ¥ = ¥ty° and 9 = 7'8, = y"d/Ax". Written out explicitly, L= vin’ +iyev- my (2.68) The Lagrangian has a global U(1) symmetry where y > ye7** and yt — pte. According to Noether’s theorem there is an associated con- Served carrent. To find it-we proceed--as in the charged scalar field theory. We allow a to depend on x and treat a(x) as an independent field. Under the above phase transformation > $+ }[Ja(x)]y. Us- ing the equation of motion for a(x), namely 3,(3.%/8[8,a(x)) — OL/da(x) = 0, we find the conservation law ait = 0 jt ty nee (2.69) Now we set a = constant to recover our original theory. The total conserved charge is Q-= f Pxjr= f ax pty (2.70) For relativistic quantum mechanics in the absence of interactions this is a trivial result because of (2.65). For the field theory we treat y as the basic field. The momentum conjugate to this field is, from (2.68), ae U= say79 7 iy! (2.7) because y°y° = 1. Thus, somewhat paradoxically, y and yt must be treated as independent entities in the Hamiltonian formulation. The 2.5 Fermions 23 Hamiltonian density can be found using the standard procedures, _ a: a - = n+ -g= vlig)e-2- V(-iy+v +m)y (2.2) The partition function is Z = Tre-M4-10) (2.73) Apart from two differences, which could be lost in the formalism if we are not careful, we can follow the steps leading up to (2.19) to write Z= fliayt][ay] x xero| [ar f ax3(-v Oe iy: v- m+ wr] }y] (2.74) We emphasize-again: that in the field- theory yt -and- are independent fields which must be integrated independently. In contrast to boson fields, there is no advantage in attempting to integrate out the conjugate momentum separately from the field. The two differences have to do with the periodicity of the field in imaginary time + and with the nature of the “classical” fields (x, 7) and WG 1) over which we are supposed to integrate. The canonical commutation relations for bosons are [$(x, t), #(y, 1)] =in8(x = y) / [$(x, 4), $y, 0] = [#(x, 1), #0, 9] =0 (2.75) We keep h explicitly in these relations to see the classical limit more clearly. The corresponding relations for fermions are {$.(x 1), HO. 1)} = 8b,98(x - y) {$.0,,1), do.9} = (00), 80.)} =0 (2.76) In the limit A - 0, the field operators are replaced by their eigenvalues. For ‘the case of bosons, these eigenvalues are actually c-number func- tions, as illustrated in (2.1). We have expressed the partition function as a functional integral over these c-number functions, or “classical fields.” For the case of fermions, the limit A — 0 is rather peculiar since the eigenvalues replacing the field operators anticommute with each other! 24 2. Functional integral representation This is intimately connected with the Pauli exclusion principle and with the famous spin-statistics theorem. Note that (2.74) instructs us to integrate over these “classical” but anticommuting functions. The mathematics necessary to handle this situation was studied by Grassmann. There are Grassmann variables, Grassmann algebra, and Grassmann calculus. . For a single Grassmann variable 4, there is only one anticommutator to define the algebra, {a1} =0 (2.71) Because of this, the most general function of n is (using a Taylor series expansion) {(9) = a + by, where a and b are c-numbers. Integration is defined by fan=0, fany=1 - (2.78) “The first of these says that the integral is invariant under the” shift 1 + a, and the second is just a convenient choice of normalization. In a more general setting, we may have a set of Grassmann variables ‘ny i= 1 to N, and a paired set nf. The algebra is defined by {a 0j) = {ap af} = (nt. ah} <0 (2.79) ‘The most general function of these variables may be written as frat Lan, + Loaf + Laynin, + Lo, satnt i i inj inj + Leyalay +--+ téababn «> aw (2.80) as Integration over all variables of (2.80) is defined to be . faskdn ---dildnyf=d (2.81) Integrals over Grassmann variables were introduced for the explicit purpose of dealing with path integrals over fermionic coordinates. The bibliography at the end of the chapter refers the interested reader to more detailed treatments. For our purposes, the only integral we need is fanaa, --- dafydny e¥?* = det D (2.82) 25 Fermions 25 where D is an N X N matrix. This formula is trivial to prove in case N = 1 or 2. The general case is left as an exercise for the reader. As with the bosons, it is most convenient to work in (p, w,) space instead of (x, ) space. In imaginary time we write ¥.(x,7) = (A/V) DL el@=torp,. ,(p) (2.83) np where both n and p run over negative and positive values. For an arbitrary function defined on the interval 0 <1 < B the discrete fre- quency w, can take on the values naT. For bosons, we argued that we must take w, = 2anT in order that $(x, 7) be periodic, which followed from the trace operation in the partition function. This can be verified by examination of the properties of the thermal Green’s function defined by Gs(x,y; 7,0) = Z-" Tr{ 67, [$(x, 7)$(y,0)]} (2.84) Here T, is the imaginary time ordering operator, which for bosons reads T,[$(14)6(1)] = 6(71)6(2)0(m — 2) + $(%2)6(11)0(% 1) : (2.85) Using the fact that 7, commutes with 6 = e~°*, K = H — y@, and the cyclic property of the trace, we find Ga(x,y;7,0) = Z-? Tr[e“*%6(x, +) $(y,0)] = 2" Tr[$(y,0) e- PS (x, 7)] = Z-' Tr[e~ PX eFX$(y, 0) e-PXG (x, 7)] = Z7" Tr[e“P*6(y, B)$(x, 7)] = Z-' Tr{ 67, [$(x, 7) 6(y, B)]} . = Ga(x,y; 1, 8) (2.86) (Notice that $(y, 8) = e"$(y,0)e"* in analogy to the real-time Heisenberg time evolution $(y, t) = e'”(y, 0)e-'”) This implies that (y, 0) = o(y, B) and hence w, = 2anT. For fermions, however, instead of (2.85) one has (in direct analogy to the real-time Green’s functions) T[¥(n)¥(2)] = $(m) 9 (1) 0(7 —n)- $(m)4 (1) O(n -1) . (2.87) 26 2, ' Functional integral representation Following the same steps as in (2.86), one is led to G,(x,y; 7,0) = —Gp(x,y; 7, B) (2.88) This means that ¥(x,0) = —¥(x, 8) (2.89) and hence , o, = (2n+ Ia. The antiperiodicity required of fermion fields is in no way inconsistent with the trace operation in the partition function. The trace only means that the system return to its original state after a “time” 8. Since the sign of y is not an observable, the right side of (2.89) describes the same state as the left side. A relevant analogy would be the fact that a fermion’s wavefunction changes sign if a rotation of 27 is made about some axis. Both the angle 9 and the imaginary time + are defined on compact intervals. : Now we are ready to evaluate (2.74). Inserting (2.83) and using (2.82) we get : 2~ [IIIT fiat.) 44.0] $= LEW (0) Paha) D = -ip[(—ia, +n) — y°y +p — my] (2.90) and Z= det D (2.91) In (2.91) the determinantal operation is to be carried out over both Dirac indices (4 x 4) and in momentum-frequency space. Employing the identity Indet D = Trin D (2.92) and (2.67) one finds In Z = 2D Lin{ 62[(«, + ip)? + o?]} (2.93) np Since both positive and negative frequencies are summed over, (2.93) can 2.5 Fermions 27 be put in a form analogous to (2.55): inZ = YY {inl 62(o2 + (w — w)’)] + n[47(u2 + (w + 2)")]} ap (2.94) Following (2.37) we write In[(2n + 1)°x? + 6(o + w)’] 249 do? - {** Tas + In[1 + (2n + 1)*2?} (2.95) The sum over n may be carried out by using the summation formula = 1 (cot rx — cot ry) n=) ae @9 Phin pitee a - 2 1 1 fi me Qn+1)a +e a(t Oe i) (2.97) Integrating over 8, and dropping terms independent of B and p, we finally obtain 3, coe inz= 2 aay lte + la( $ eA) + In(1 + e-#*M)] : T (2.98) This result agrees with the result that was derived in Chapter 1 using an entirely different method. Notice that the factor of 2 in (2.98) corresponding to the spin-} nature of the fermions comes out automatically. Separate contributions from particles (4) and antiparticles (—) are evident. Finally, the zero-point of the vacuum also appears in this formula. To recapitulate, the difference between bosons and fermions in the functional integral approach to the partition function is essentially twofold. First, for fermions we must integrate over Grassmann variables instead of c-number variables. Contrast the result Z = det D for fermions (2.91) with the result Z = (det D)~* for bosons (2.34). Integra- tion over c-number variables for fermions would have led to a factor of —1 in (2.98) instead of 2. Second, and this is actually related to the first, 28 2. Functional integral representation is the fact that the fermion fields are antiperiodic in imaginary time, with period 8, instead of periodic as is the case for bosons. The consequence is w, = (2n + 1)aT for fermions whereas w, = 2naT for bosons. These two points account for the difference between (2:57) (with [ = 0, of course) and (2.98) 2.6 Remarks on functional integrals The notation used for functional integration (and differentiation!) is deceivingly simply. It must be kept simple, for if we think back on the tremendous progress made in mechanics and electromagnetism in the nineteenth century, it was certainly made easier by the introduction of compact notation for differentiation, integration, and vectors. This also seems to be the case with functional methods in modern quantum physics. However, it is also clear that the mathematical symbols we are using often represent rather exotic entities compared to the repertoire of the uninitiated. For example, (2.3) uses a Dirac délta-function’ whose argument is a difference between two functions. A less formal and compact, but more practical, way to view these objects is to start with a complete orthonormal set of real functions for the physical problem of interest. Call it w,(x) with n any positive integer. Then any function may be written as a(x) = zr a,w,(x) (2.99) Another function may be expresed as (x) = E bamls) (2.100) Then . 8[a(x) - b(x)] = i 8(a, — by) (2.101) and faa) = Tf aa, (2.102) and so on. Most physical problems are defined on a space of a continu- ous-variable-like position. For such problems it is intuitively obvious

You might also like