You are on page 1of 267
Supersymmetry and Supergravity Supersymmetry and Supergravity SECOND EDITION, REVISED AND EXPANDED by Julius Wess and Jonathan Bagger Princeton Series in Physics Princeton University Press Princeton, New Jersey COLLEGE OF SCIENCE & CENTRAL LIBRARY ETT Copyright © 1992 by Princeton University Press Published by Princeton University Press, 41 William Street, Princeton, New Jersey 08540 In the United Kingdom: Princeton University Pr All Rights Reserved Oxford Library of Congress Cataloging-in-Publication Data Wess, Julius. Supersymmetry and supergravity / by Julius Wess and Jonathan Bagger. — 2nd rev. and expanded ed. pcm. ISBN 0-691-08556-0. — ISBN 0-691-02530-4 (pbk.) 1. Supersymmetry. 2. Supergravity. I. Bagger, Jonathan. IL Title. QC174.17.89W47 1991 530.1'43—de20 90-26372 Princeton University Press books are printed on acid-free paper, and meet the guidelines for permanence and durability of the Committee on Production Guidelines for Book Longevity of the Council on Library Resources Printed in the United States of America by Princeton University Press, Princeton, New Jersey 13579108642 13579108642 (pbk.) To Traudi CONTENTS PREFACE TO THE SECOND EDITION PREFACE L IL. IIL. IV. Vv. VI. VIL VII. IX. XL. XII. XIII. XIV. XV. XVI. XVIL. XVIII. XIX. WHY SUPERSYMMETRY? REPRESENTATIONS OF THE SUPERSYMMETRY ALGEBRA COMPONENT FIELDS SUPERFIELDS CHIRAL SUPERFIELDS VECTOR SUPERFIELDS GAUGE INVARIANT INTERACTIONS SPONTANEOUS SYMMETRY BREAKING SUPERFIELD PROPAGATORS FEYNMAN RULES FOR SUPERGRAPHS NONLINEAR REALIZATIONS DIFFERENTIAL FORMS IN SUPERSPACE GAUGE THEORIES REVISITED VIELBEIN, TORSION, AND CURVATURE BIANCHI IDENTITIES SUPERGAUGE TRANSFORMATIONS THE 6 = 8 = 0 COMPONENTS OF THE VIELBEIN, CONNECTION, TORSION, AND CURVATURE THE SUPERGRAVITY MULTIPLET CHIRAL AND VECTOR SUPERFIELDS IN CURVED SPACE 11 21 25 30 36 43 51 61 79 88 93 101 109 117 127 132 140 146 viii XX. XX!I. XXII. XXII. XXIV. XXV. XXVI. CONTENTS NEW © VARIABLES AND THE CHIRAL DENSITY THE MINIMAL CHIRAL SUPERGRAVITY MODEL CHIRAL MODELS AND KAHLER GEOMETRY GENERAL CHIRAL SUPERGRAVITY MODELS GAUGE INVARIANT MODELS GAUGE INVARIANT SUPERGRAVITY MODELS LOW-ENERGY THEOREMS APPENDIX A: Notation and Spinor Algebra APPENDIX B: Results in Spinor Algebra APPENDIX C: Kahler Geometry APPENDIX D: Isometries and Kahler Geometry APPENDIX E: Nonlinear Realizations APPENDIX F: Nonlinear Realizations and Invariant Actions APPENDIX G: Gauge Invariant Supergravity Models 155 162 175 180 192 204 217 225 232 235 239 245 253 256 PREFACE TO THE SECOND EDITION Since these lectures were given, supersymmetric particle phenomenology has been the subject of extensive study. Many models have been proposed, including some that make essential use of the supergravity multiplet. A variety of experimental searches have been carried out, and more are planned for the future. Given this state of affairs, we felt that the second edition of this book should go substantially beyond the first. The second edition contains a total of six new chapters and five new appendixes. The new chapters are primarily devoted to deriving the component form of the most general supersymmetric gauge theory coupled to supergravity. The resulting Lagrangian, presented in Chapter XXV and Appendix G, is the starting point for all phenomenological studies of supergravity theories. Model- builders can use the Lagrangian without having to read the rest of the book. The new appendixes contain introductions to Kahler geometry, iso- metries, and nonlinear realizations of symmetries. The material is essen- tial for understanding the derivations in the book, but it is also of more general interest. In Chapter XXVI the techniques of nonlinear realizations are applied to supersymmetric gauge theories. The results pave the way for a model-independent approach to supersymmetry phenomenology, in the spirit of chiral dynamics. The new additions have broadened the scope of the book so that it should appeal to physicists of formal and phenomenological interests. In its present form, the book provides a theoretical basis for further phe- nomenological studies of supersymmetric theories. We would like to thank the Gottfried Wilhelm Leibnitz Program of the DFG and the Alfred P. Sloan Foundation for financial support during the preparation of the second edition. Juitus Wess JONATHAN BAGGER UNIVERSITY OF MUNICH JoHNS HOPKINS UNIVERSITY February 1991 PREFACE The strong interest with which these lectures on supersymmetry and supergravity were received at Princeton University encouraged me to make their contents accessible to a larger audience. They are not a sys- tematic review of the subject. Instead, they offer an introduction to the approach followed by Bruno Zumino and myself in our attempt to develop and understand the structure of supersymmetry and supergravity. This book consists of two parts. The first develops a formalism which allows us to construct supersymmetric gauge theories. The second part extends this formalism to local supersymmetry transformations. At the end of each chapter, two papers are cited which I recommend to the reader. I am aware that this selection does not do justice to many authors who have contributed to the subject. However, I would like to draw attention to the more complete lists of references found in P. Fayet and S. Ferrara, Supersymmetry, Physics Reports 32C, No. 5, 1977, and P. Van Nieuwenhuizen, Supergravity, Physics Reports 68C, No. 4, 1981. Throughout the text, important equations are numbered in boldface. They are collected at the end of each chapter. Exercises are also included along with each chapter; many of them contain information essential to a deeper understanding of the subject. This book was prepared in collaboration with Jonathan Bagger, without whom it would never have been written. Both Jon and I would like to thank Winnie Waring for her devoted assistance in the preparation of the manuscript. As a tribute to her high standards, we have tried our best to avoid errors in factors and signs. Many people have helped eliminate these errors. In particular, we would like to thank Martin Miiller for his assistance with the second half of the book. I wish to express my gratitude to the Federal Republic of Germany for the grant which made possible my stay at The Institute for Advanced Study as an Albert Einstein Visiting Professor, and Jon would like to express his appreciation to the U.S. National Science Foundation for his Graduate Fellowship at Princeton University. In conclusion, I would like to thank Stephen Adler and the Members of the Institute for Advanced Study, as well as David Gross and the Department of Physics at Princeton University, for their most encour- aging and critical interest in these lectures. JuLius Wess UNIVERSITY OF KARLSRUHE May, 1982 I. WHY SUPERSYMMETRY? Supersymmetry is a subject of considerable interest among physicists and mathematicians. Not only is it fascinating in its own right, but there is also a growing belief that it may play a fundamental role in particle physics. This belief is based on an important result of Haag, Sohnius, and Lopuszanski, who proved that the supersymmetry algebra is the only graded Lie algebra of symmetries of the S-matrix consistent with relativis- tic quantum field theory. In this chapter, we shall discuss their theorem and its proof. (Readers specifically interested in supersymmetric theories might prefer to start directly with Chapter II or III.) Before we begin, however, we first present the supersymmetry algebra: {0,4,.0pn} + = 204;"Pm On {0,4,057}. = {Qi4,Qpe} « =0 = @) [PmQx"]_ = [Pn Qia]- = 0 [PusP,J- = 0. The Greek indices (x, f,...,%,8,...) run from one to two and denote two-component Weyl spinors. The Latin indices (m,n,...) run from one to four and identify Lorentz four-vectors. The capital indices (A, B,...) refer to an internal space; they run from 1 to some number N > 1.Thealgebra with N = 1 is called the supersymmetry algebra, while those with N > 1 are called extended supersymmetry algebras. All the notation and conventions used throughout this book are summarized in Appendix A. We are now ready to consider the theorem. Of all the graded Lie algebras, only the supersymmetry algebras (together with their extensions to include central charges, which we shall discuss at the end of the chapter) generate symmetries of the S-matrix consistent with relativistic quantum field theory. The proof of this statement is based on the Coleman-Mandula theorem, the most precise and powerful in a series of no-go theorems about the possible symmetries of the S-matrix. 4 I. WHY SUPERSYMMETRY? The Coleman-Mandula theorem starts from the following assumptions: (1) the S-matrix is based on a local, relativistic quantum field theory in four-dimensional spacetime; (2) there are only a finite number of different particles associated with one-particle states of a given mass; and (3) there is an energy gap between the vacuum and the one particle states, The theorem concludes that the most general Lie algebra of symmetries of the S-matrix contains the energy-momentum operator P,,, the Lorentz rotation generator M,,,,, and a finite number of Lorentz scalar operators B,. The theorem further asserts that the B, must belong to the Lie algebra of a compact Lie group. Supersymmetries avoid the restrictions of the Coleman-Mandula theorem by relaxing one condition. They generalize the notion of a Lie algebra to include algebraic systems whose defining relations involve anticommutators as well as commutators. These new algebras are called superalgebras or graded Lie algebras. Schematically, they take the following form: {Q.0}.=X [XxX]-=x" [Ox]-=0" (12) Here Q, Q’, and Q" represent the odd (anticommuting) part of the algebra, and X, X’, and X” the even (commuting) part. The operators X are determined by the Coleman-Mandula theorem. They are either elements of the Poincaré algebra P = {P,,,M,,,} OF elements of a Lorentz-invariant compact Lie algebra «/. The algebra / is a direct sum of a semisimple algebra ./, and an Abelian algebra ,, A= A, ® Ay. The generators Q@ may be decomposed into a sum of representations irreducible under the homogeneous Lorentz group #: D = Ya. rain in: (1.3) The Q, a, °**G, and o, ---%. They belong to irreducible spin-+(a + b) repre- sentations of Y. Since the Q’s anticommute, the connection between spin and statistics tells us that a + b must be odd. We shall now invoke two additional assumptions to prove that a + b = 1. These assumptions are: - +a, are Symmetric with respect to the underlined indices 1 ot tasty (1) the operators Q act in a Hilbert space with positive definite metric; and (2) both Q and its hermitian conjugate Q belong to the algebra. I. WHY SUPERSYMMETRY? 5 We start by considering the anticommutator {Quy -+-aads +++ 4v Ops «afi =a) > (1.4) where all the indices are assigned the value 1. The product (1.5) a belongs to a spin-(a + b) representation of Y, so [Or niciGiins} (1.6) must close into an even element of the algebra with spin (a + b). From the Coleman-Mandula theorem, we know that this element is either zero or a component of P,,. For a + b > 1, it must be zero. The anticommutator (1.6) is a positive definite operator in a Hilbert space with a positive definite metric. This tells us that Q,...,,,...; =0 fora +b> 1. Since the Q,,...2,, é» are irreducible under %, they all must vanish for a + b > 1. From this we conclude that the odd part of the supersymmetry algebra is composed entirely of the spin-} operators Q," and Qin. The anticommutator of Q," and Q;y closes into P,s, {0.0m} = PaaC*us (1.7) where P,; = 6,;"Pm. In Exercise 1 we show that the finite-dimensional matrix C4, is hermitian. It may therefore be diagonalized by a unitary transformation. Since {Q,",0j,} is positive definite, the matrix C’,, has positive definite eigenvalues. This lets us choose a basis in the odd part of the algebra such that {Q.",Oim} = 2P2d*m- (1.8) We now turn our attention to the anticommutator of two odd elements, both with undotted indices. The right-hand side of this expression may be decomposed into symmetric and antisymmetric parts. The symmetric part has spin 1. From the Coleman-Mandula theorem, the only possible candidate is the Lorentz generator M4: {Q,",OpM} = bapXX% + Mag ¥™. (1.9) 6 I. WHY SUPERSYMMETRY? From the fact that P,, commutes with Q," (see Exercise 2), we find that the YE¥ must vanish. This lets us write the commutator (1.9) as follows: {0.'.05"} = e,,0°L4B,. (1.10) Here B, is a hermitian element of ./, ® »/, and a’& is antisymmetric in L and M. With this result, the supersymmetry algebra takes the fol- lowing form: {0.',Opm} = 2055"P mom ma") = [PuQja] = [PmQa"] = [Pm Qiu] = 0 {0.",05™} = bg SB, = E4pXE (Bin Din) = 660 a! = 09X" vy aay [0.°,B,] = S/*Q.™ [B..Oir] = S* Oi [Be.B] = icen*By- We shall now use the Jacobi identities to further restrict the coefficients aL! and S;*y in (1.11). The ordinary Jacobi identity may be easily extended to include anticommutators, as is done in Exercise 3: {A, {B,C]] + {B, {C,A]] + {C, {4,B]] = 0. (1.12) The bracket structure { , ] signifies either commutator or anticommutator, according to the even or odd character of A, B, and C. The signs are determined by the odd elements. If the odd elements are in a cyclic permutation of the first term, the sign is positive; if not, it is negative. By exploring the Jacobi identities in a certain order, we shall arrive at our results as quickly as possible. We first consider the identity [Be, {Q.". Qiu] + (2.", [QamsBel} — {Qiu [B..0,]} = 0. (1.13) The first term vanishes because B, and P,, commute. The second and third terms give — 10," Dax" ak + (DjnQe Sex = 0, (1.14) or 2PxilS* ye” — S/'y] = 0. (1.15) I. WHY SUPERSYMMETRY? 7 Equation (1.15) is true only if Sh = Sey, (1.16) so S/“ is hermitian. Next we use the identity [Be {Q."Op"}] + {Q,", [Qp4,Be]} — {Qp"[B,.Q."]} = 0 (1.17) to prove that the generators XL@ = a’ 4MB, form an invariant subalgebra of x, © ,. Evaluating (1.17) with the help of (1.11), we find byp{ LB XR] + SM EXE — S,',XX8} = 0. (1.18) This shows that the commutator of B, with X£¥ closes into the set of generators XL¥. The X&™ are linear combinations of the B;, SO we con- clude that the X£¥ form an invariant subalgebra of # = A, @ A. We now use the identity [Q.", {Qp™,Oix}] + [Op™, (jx.0n"}] + [Osx {Q2’.Op"}] = 0 (1.19) to show that the generators XL¥ commute with all the generators of . Combining (1.19) with (1.11), we find &plOjx, XX] = 0, (1.20) so [XEXL] = FPTO.K 0)", XB = 0. at) This implies that the X£¥ form an Abelian (invariant) subalgebra of «/. Since », is semisimple, the X£¥ are elements of ./, and commute with all the generators of .f: [X24,B,] = 0. (1.22) For this reason, they are called central charges. Inserting (1.22) into (1.18), SM pXLE — S$), XME = 0, (1.23) and substituting XMS = a@SB,, we find SM aht® — Sh eakM& = 0. (1.24) 8 I. WHY SUPERSYMMETRY? From the fact that 5“ is hermitian and a,¥* antisymmetric, we conclude SM ake = —GhtkS* 2, (1.25) In Exercise 4 we show that the S/“, form a representation of A, ® Ap. Equation (1.25) tells us that the matrices a, intertwine the representation S, with its complex conjugate —S,*. Central charges exist only if the algebra A, ® A, permits such intertwiners. A trivial example is given by SME = 0. Another is provided by orthogonal groups, where S, = —S,*. A third example is given in Exercise 5. No further restrictions follow from the other Jacobi identities, as may be proven by checking them all. We have therefore found the most general supersymmetry algebra: [PnoPu] = 0 [Pn.Q.!] = [Pm Qui] = 0 [PmsBe] = [Pm-X24] = 0 {Q.' Bayo} = 20." P Dh {Q4°,Qp™} = bap X O:1,Qpm = &5p. LM {Qin Qin} = &pX* [XY Dix] = [X40,5] = 0 [X2EX8] = [xXe4.B,] = 0 [Be.Bn] = icen'By [2.",.B,] = SQ." [QuB) = -S* "Qn XE = of LB, (1.26) This is the most general graded Lie algebra of symmetries of the S-matrix consistent with relativistic quantum field theory. If central charges exist, they must be of the form X&¥ = a £¥B,, where a’ intertwines the rep- resentations S, and —S*“. REFERENCES S. Coleman and J. Mandula, Phys. Rev. 159, 1251 (1967). R. Haag, J. Lopuszanski, and M. Sohnius, Nucl. Phys. B88, 257 (1975). I. WHY SUPERSYMMETRY? 9 EQUATIONS {0.4.0 pn} + = 2046"Pm dy {04,047} + = {QiaQpn} + =0 . * o [Pm:Q.“]- = [PmQza]- = 0 [PwP,J- = 0. (A, {B, CJ] + {B, {C, A] + {C, {4B]] = 0. (1.12) S¥ ye = Shy. (1.16) SMgah&h = —qhk gee L, (1.25) [PmsPr] = 0 [PmsQ2"] = [Pn Qin] = 0 [Pm-Be] = [Pm X24] = 0 {Q2',Qim} = 2044"Pmd* {Q,",OpM} = b_pXLF {Qin Om} = 6X * rm (1.26) [XE] = [X40] = 0 [X2E X82] = [X4B,] = 0 [BeBu] = icen By [Q.",Be] = S¢4Q.™ [Qin.BY] = -S*"Oim XE = a LEB, EXERCISES (1) Prove that Cy in (1.7) is hermitian by comparing the anticommutator (1.7) with its hermitian conjugate. (2) Show that [Q,,P,,] = 0. Start from the fact that there are no spin -3 generators. Deduce that [P,;,0,"] = Z"ye.,04", where the Z"y are some set of numbers. Use the Jacobi identity for [Pgg, [P2:.2,"]] to prove that all the Z“,, vanish. This shows that the Q,” are transla- tionally invariant. 10 I. WHY SUPERSYMMETRY? (3) Prove the Jacobi identity (1.12). In particular, verify [B,, [B2.B3]] + [B2,[B3,B,]] + [Bs, [B:,.B2]] = 0 (21, [B2-Bs]] + (Bz, [Bs.Qi]] + [Bs. [01,B2]] = 0 [B,, {Q2,03}] + {Q2,[Q3.B1]} — {Q3,[B1.Q2]} = 0 [Q1, {Q2,03}] + [Q2, {03,01}] + [Qs, (Q1,02}] = 0. (4) Use the identity [Be, [BmQz']] + [Bm [Q.",B/]] + [@,", [Be.Bm]] = 0 to prove [S,S2] = icne'S;- (The matrix S, has elements S,*,,.) Show that —S*, satisfies the same commutation relations. (5) The Pauli matrices 6 and their conjugates —o* both form representa- tions of SU(2). Show that « is an intertwiner between these rep- resentations. Verify that the commutator {Qn",OpM} = tape M(cyZ, + icrZ2) is consistent with the Jacobi identities if Z, and Z, are central charges. II. REPRESENTATIONS OF THE SUPERSYMMETRY ALGEBRA An exciting feature of the supersymmetry algebra is that there exist quantum field theories in which the supersymmetry generators Q, may be represented in terms of conserved currents J,”: 0, = fax, a, (2.1) ae J," =0. The currents J,” are local expressions of the field operators. The algebra (I) is satisfied because of the canonical equal-time commutation relations, and the Hilbert space spans a representation of the supersymmetry al- gebra. In this chapter we shall study the supersymmetry representations of one-particle states. The energy-momentum four-vector P,, commutes with the super- symmetry generators Q, and Q;. The mass operator P? is a Casimir operator, so irreducible representations of the supersymmetry algebra are of equal mass. We shall construct these irreducible representations by the method of induced representations, considering fixed time-like (P? < 0) and light-like (P? = 0) momenta. Before we do this, however, we shall first prove that every representa- tion of the supersymmetry algebra contains an equal number of bosonic and fermionic states. We begin by introducing a fermion number operator Np, such that (—)** has eigenvalue +1 on bosonic states that —1 on fermionic states. It follows immediately that (~)"Q, = —Q.(—)N". (2.2) For any finite-dimensional representation of the algebra (such that the trace is well-defined), we find Tr[(—)*7{0,4, Opn}] = TrE(—)""(Q.4O pn + in2.")] = Trl-O,4(-)"" je + Q.4(—)* Ope] =0. (2.3) 12 Il, THE SUPERSYMMETRY ALGEBRA Here we have used (2.2) and the cyclic property of the trace. Substituting {0,".Qpp} = 20,9"Pm Op (2.4) from the supersymmetry algebra (I), we conclude Tr[(—)"{O,4,05?}] = 200" 54y Tr[(—)"" Pa] =0. (2.5) For fixed non-zero momentum P,,, this reduces to Tr(—)*" = 0, (2.6) proving that supersymmetry representations contain equal numbers of bosonic and fermionic states. We are now ready to construct the representations of the supersymmetry algebra corresponding to massive, one-particle states, P? = —M?. We first boost to the rest frame, where P,, = (—M, 0, 0, 0). In this frame, the algebra (I) takes the following form: {Q,",Ojn} = 2M 5,048 {4.04%} = {Di4.Ojn} = 0. en The indices A and B run from 1 to N. The generators Q may be rescaled 1 a,4 = —— 9,4 2M 1 . =-—9@,; J2M aA (2.8) (a,4)* to show that (2.7) is isomorphic to the algebra of 2N fermionic creation and annihilation operators, (a,“)* and a,4: (2.9) A(q Byti = g BSA {a,4,(ap")"} = 6,° 545 Y j = {(a,)* (a,")*} = 0. The representations of this algebra are well known. They are constructed from a Clifford “vacuum” Q. The Clifford vacuum is defined through the Il. THE SUPERSYMMETRY ALGEBRA 13 condition a,/Q = 0, (2.10) where, in contrast to the usual case, P? Q = —M? Q. The states are built by applying the creation operators (a,4)* to Q: 1 QU ty = (a, AI)" (ay) °Q. (2.11) Because the (a,4)* anticommute, Q is antisymmetric under the exchange of two pairs of indices «,A;, %;A;. Each pair of indices takes 2N different values, so n must be less than or equal to 2N. For any given n, there are 2%) different states. Summing over all n gives the dimension of the repre- sentation (2.11): 2N d=y (*) = 2". (2.12) n=0 \ 7M If the vacuum Q is not degenerate, we call (2.11) the fundamental irreducible massive multiplet. It has dimension 27%, with 2?%~' bosonic and 2?"~! fermionic states. The state with the highest spin is obtained by symmetrizing in as many spinor indices as possible. Because we must simultaneously antisymmetrize in the second index, we may only symme- trize in N spinor indices. This leads to spin-}N. The highest spin in the fundamental multiplet is $N; it occurs exactly once. All other massive multiplets are based on vacuua Q which are not invariant under the stability group. Their representations are found by composing the representation of Q with that of the fundamental multiplet. We now list a few examples. In the case N = 1, the fundamental representation consists of the states Q (a,)*Q (2.13) (a,)* (ag)? Q = tL 6(a")*(a,)*Q. 22 It has two states of spin 0 and one of spin 3. When the vacuum Q, has spin j, with j > 0, it belongs to a (2j + 1)-dimensional representation of the stability group SU(2). This leads to a multiplet with spins (j,j + 4. j —4,j). These results are summarized in the following tables for 2p 14 Il. THE SUPERSYMMETRY ALGEBRA. N = 1, 2,3, and 4: N=1 N=2 N=3 N=4 The representation space (2.11) of the algebra (2.9) also spans a repre- sentation space of the invariance group of the algebra. It is obvious from Il. THE SUPERSYMMETRY ALGEBRA 15 (2.9) that SU(2) ® U(N) is a possible invariance group. However, SO(4N) is a larger invariance group of (2.9). It contains SU(2) ® U(N) and SU(2) @ USp(2N) as subgroups. To make the SO(4N) symmetry manifest it is convenient to write (2.9) as a Clifford algebra. To do this we define the operators [ar’ + (a3)"] (2.14) peyté = e244") p3vte — a [as’ — (a‘)*). The indices 1 and 2 refer to the SU(2) spinor indices and the index 7 runs from 1 to N. By definition, the 2N operators (2.14) are hermitian. In addition, they obey the following anticommutation relations: {CAP} = 3°, (2.15) where r,s = 1,..., 4N. This is a Clifford algebra with an SO(4N) invari- ance group. The 27% states of the fundamental representation span a spin- orial representation of SO(4N). This spinorial representation contains two irreducible representations, each of dimension 2?%~', corresponding to the bosonic and fermionic states. The algebra (2.9) may also be cast in a form which exhibits the SU(2) ® USp(2N) symmetry. This is done by defining a new set of operators qa = 4,6 2 (2.16) gas = Y eaplag’)”, pat where / = I,..., N. These operators transform as follows under hermi- tian conjugation: (q)* = (al)? = egg’ *? N+ey+ ° Ba 6 t (2.17) (a8) = ~ eal = — ea! 16 Il. THE SUPERSYMMETRY ALGEBRA Equation (2.17) may be written in a more compact form (q.’)* = eAT G's, (2.18) where r,t run from | to 2N and A is the following symplectic matrix: A= (A+): (2.19) The anticommutation relations of the operators q {a'sdp'} = —eagA™ (2.20) exhibit the SU(2) @ USp(2N) invariance. This invariance group is useful because states of a given spin transform irreducibly under USp(2N). We shall now analyze the massless case, P? = 0. We begin by boosting to a fixed light-like reference frame, where P,, = (—E,0,0,£). In this frame, the algebra (I) becomes ~ 2E 0 4,0jp} = 2 64 {Q.4.2in} = 2(5 0) 5 021) {0.4,0)°} = {QisQja} = 0. Rescaling the Q’s L a4 =——Q,4 ee ' (2.22) at, =——0)4 = (a’)*, angen we find that the algebra (2.21) consists of N creation and annihilation operators, a* , and a‘: {at,a* a} = O45 (2.23) {aa} = {a gas} = 0. The operators Q,4 and Q3, are totally anticommuting and must therefore be represented by zero. The operators a* 4 and a4 raise and lower the helicity of a state by 4. Consequently, a4 annihilates the state of lowest helicity, say, 2: =0. (2.24) Il, THE SUPERSYMMETRY ALGEBRA 17 The states Qs ean As dy OS at An” sa* 4,Q, (2.25) are built by applying the creation operators a* , on the Clifford vacuum Q,. The states OQ", 9, 4,... 4, have helicity 2 + 4n. They are antisymmetric in A, --- A, and (%)-times degenerate. The state with highest helicity in this representation has helicity 7 = 2 + 4N, so the representation (2.25) has dimension 2". From this we see that one massive representation splits into 2% massless representations. We summarize these results in tables for N = 1, 2, 3, and 4: N=1 4 hs | 72-72 EE OG 2 1 3 1 1 1 1 1 $ 1 1 0 1 1 —4 1 1 -1 1 1 -4 1 1 -2 1 N=2 d nea] 72 TF TE EO yo 1 1 2 1 2 1 Itt Nw SH Ne ORR Ne to N | 18 II. THE SUPERSYMMETRY ALGEBRA N=3 2 1 3 1 3 1 1 33 4 1 3003 41 0 1 3 3 0 4 -4 1 3 3 1 -1 3 3 1 3 3 1 2 1 N=4 2 1 3 1 4 1 1 4 6 4 1 4 6 64 0 1 4 6 4 1 -4 4 6 4 1 -1 6 4 1 -3 4 1 -2 1 In CPT-invariant theories, the number of states must in general be doubled, for CPT reverses the sign of the helicity. Note, however, that the N=2,4= -4;N=4,4=—-l; and N =8,4= —2 multiplets are automatically CPT complete. To conclude this chapter, we consider the supersymmetry algebra (1.26) with central charges. We assume that P? = —M? and study the algebra in the rest frame: {0,".(Q,")"} = 2M 4, 5*y {0,,0)!} = e592 (2.26) {(0.4)* (Qp")"} = Zw ZEM — ML Il. THE SUPERSYMMETRY ALGEBRA 19 The central charges Z’ commute with all the generators, so we may choose a basis in which the central charges are diagonal with eigenvalues Z'™. These eigenvalues form an antisymmetric N x N matrix. Any such matrix may be rotated into a standard form by a unitary transformation: ZM = ULLUMAZEN, (2.27) The standard form is given by Z=c@D (N even) ; (€@D 0 (2.28) Z= ( 0 0) (N odd) where D is diagonal with positive real eigenvalues Z,, and ¢ is the 2 x 2 antisymmetric matrix with e!? = 1. We shall study the case with N even. (The case with N odd is analogous.) We start by decomposing the indices L and M in accord with (2.28), L=(am), M = (bn), (2.29) where a,b = 1,2 and mn =1,...,4N. We then perform a unitary transformation on the Q,", 0,! = UO". (2.30) This allows us to write the algebra (2.26) in the following form: {0,0"(G,"")*} = 2M6,P 5%,6", {G25} = cage” OZ, 231) Ge") Gy") = Ean Sms The operators 0," and (6,2")* may all be expressed as linear combina- tions of (2.32) and their conjugates (a,")* and (b,")*. The operators a and b satisfy the 20 Il. THE SUPERSYMMETRY ALGEBRA following algebra: {a,",ap"} = {b,"by"} = {a,",by"} = 0 {a,\(ag")*} = 5,,6QM + Z,) (2.33) {b'lbp")"} = 5p 52M — Z,). From these relations we see that Z, < 2M for all n. If a set of Z; = 2M, withi = 1,...,r, the corresponding operators b; must vanish. This leaves us with a Clifford algebra of 2(N — r) creation and annihilation operators. The representations of this algebra have been studied before. REFERENCES W. Nahm, Nucl. Phys. B135, 149 (1978). S. Ferrara, C. A. Savoy, and B. Zumino, Phys. Lett. 100B, 393 (1981). EXERCISES (1) Show that there are equal numbers of bosonic and fermionic states in the representation (2.11). Assign the number +1 to each bosonic state and the number —1 to each fermionic state. Then compute the sum 2N y (-"ey. n=0 (2) Prove that the highest spin in the fundamental multiplet occurs exactly once. Construct the state with the highest spin in the z-direction. Verify that this state is unique. (3) Show that ¢,,(z,)* transforms like 7, under SU(2) transformations. This shows that complex conjugation raises and lowers SU(2) indices. (In particular, lower dotted indices of SL(2,C) transform as upper indices under the SU(2) rotation subgroup.) Ill. COMPONENT FIELDS To formulate a supersymmetric field theory we must first represent the supersymmetry algebra (I) in terms of fields not restricted by any mass- shell conditions. Anticommuting parameters ¢*,é, simplify the task: {e,0?} = {6,09} =~ = [Pn €*] = 0. (3.1) These parameters allow us to express the supersymmetry algebra entirely in terms of commutators: [60,60] = 2f0"EP,, [€0,62] = [60.20] =0 (3.2) [P™,éQ] =[P",€0] = 0. Here we use the summation convention outlined in Appendix A: 2= 80, = %,0". A component multiplet is a set of fields (A, W,...) on which we define the infinitesimal transformation 6;: 6A = (EQ + €Q) x A, 7 (33) =(0+ Q)xw The transformation 6, satisfies (6,5, — 6:6,)A = Ana" — Fo")PA = ~2i(go"E — Eo") yA (4) in accord with (3.2). This supersymmetry transformation maps tensor fields into spinor fields and vice versa. From the algebra (I) we see that Q has mass dimension 3. Therefore, fields of dimension / transform into fields of dimension / + 4 or into derivatives of fields of lower dimension. 22 III]. COMPONENT FIELDS Starting with the scalar field A, we define the spinor yw as the field into which A transforms: 5A = /2ey. (3.5) The field y transforms into a tensor field of higher dimension and into the derivative of A itself: dab = i20"@0,A + J2EF. G.6) The coefficient of 0,,4 is chosen to guarantee that the commutator of 5,0¢A = 2igo"T0,A + 2ENF (3.7) closes in the sense of (3.4). The same commutator acting on the field y yields (5,52 — bg5, yh = —2i(na"S — Cai) Oh —io"a" O,wlno"e — éo"7] + /2(E5,F — 1 5:F). (3.8) This closes if OF = if 2260, (3.9) It follows from (3.6) that the commutator on F closes as well. If we had been willing to use the field equations, —ia"0,y = my, Eq. (3.9) could have been satisfied by F = —m4A*. In this case we would have said that the transformations (3.5) and (3.6) close through the field equations. In extended supersymmetry we are sometimes forced to close the commutators through the field equations because we do not yet know the full multiplet structure of the theory. The component multiplet which we have constructed is called the chiral or scalar multiplet: eA = 28 bab = in/20"20,A + \/2EF (3.10) OF = i/2G"O a. These fields form a linear representation of the supersymmetry algebra (1). If A has dimension 1, then yy has dimension 3, while F has dimension 2 and must assume the role of auxiliary field. From Eq. (3.10) we see that F transforms into a space derivative under 6:. This will always be the case for the component of highest dimension in any given multiplet. III. COMPONENT FIELDS 23 To construct an invariant action it is sufficient to find combinations of fields which transform into space derivatives. Such combinations are given by Ly = id, ba" + AXOA + F*F (3.11) and Ly = AF + AtP* — Wy - 5. (3.12) From the complete Lagrangian L=L,+ mF, (3.13) we determine the field equations: ig" 0, + mp = 0 F + mA* =0 (3.14) OA + mF* = 0. They describe a Weyl spinor y and a complex scalar A, both of mass m. The Lagrangian (3.13) has the curious property L=:L3, (3.15) where : : denotes normal ordering. This simply reflects the fact that super- symmetric theories must contain an equal number of bosonic and fer- mionic degrees of freedom for a given mass. Equation (3.15) holds as long as supersymmetry remains unbroken. We may also expect that the vacuum expectation value of the energy-momentum tensor 7" vanishes in an unbroken supersymmetric theory. This may be seen by considering J,”, the local current of the supersymmetry charge Q,, Q, = Jaxse. (3.16) The supersymmetry algebra (I) yields the energy-momentum tensor T”" as an anticommutator {Oz,J™} = 2045"Tp" + S.T. (3.17) The additional Schwinger terms have zero vacuum expectation value. 24 III. COMPONENT FIELDS Therefore, = 0 as long as Q;|0> = 0. Bruno Zumino was the first to realize that this might account for a vanishing cosmological con- stant of the observable universe. REFERENCES J. Wess and B. Zumino, Nucl. Phys. B70, 39 (1974). B. Zumino, Nucl. Phys. B89, 535 (1975). EQUATIONS (5,5, — 5,65,)A = 2no"% — Ea") PA = —2i(no"S — E07) 0,4 (3.4) 5A = J28p Sah = i/20"%0,,A + /2EF (3.10) OF = if 2EG"A a Ly = id,We"y + A*OA + F*F. (11) Ly, = AF + A*F* — sw - 50. (3.12) EXERCISES (1) Show yz = wh, xo" = —Po"y, (xo")* = worZ, and (xo"a"y)* = Yo"o"Z. (2) Prove the Fierz rearrangement formula = 1 ms) m WOME = —5(Po"DWo");.- (3) Use (3.5) and (3.6) to calculate 5, 5eMy = —2ino"E Ana — iLo"a" AW],(no"e) + /26,5,F. (4) Eliminate the auxiliary field F from the Lagrangian (3.13) to obtain L = id,bors — Amie + WV) + AXA — m2 A*A, (5) Show that 6(AF — 4ynp) = i/2Eo"0,(AW). IV. SUPERFIELDS Superfields provide an elegant and compact description of supersymmetry representations. They simplify the addition and multiplication of rep- resentations and are very useful in the construction of interacting Lagrangians. We shall show that superfields may always be constructed from component representations. Component fields may always be recovered from superfields by power series expansion. We begin with the observation that the supersymmetry algebra may be viewed as a Lie algebra with anticommuting parameters [ Eq. (3.2)]. This motivates us to define a corresponding group element: G(x,0,0) = eft "Pm 0+ 001, (4.1) It is easy to multiply two group elements using Hausdorff’s formula ete® = e4*8+314.B1+--~ because all higher commutators vanish. We find GOEZG(x"0.) = Gx" + orE — Eo" O+ E048. (42) As usual, multiplication of group elements induces a motion in the parameter space, (EE): (x",0,0) > (x™ + iOo"E — i€o"H,0 + E,0+2%). (4.3) This motion may be generated by the differential operators Q and Q: +O =6 ( — i6,,,"0* an) +&, (a, — i6?o,j"eh* in): (44) Here we use the same letters Q,Q for the differential operators as for the group generators because the differential operators do indeed represent the infinitesimal group action on the parameter space: {,:03} = 2io.s"6, {0,06} = {Qi.0)} = 0. (45) 26 IV. SUPERFIELDS Note, however, the change in sign, P,, = —i0,,. This stems from the fact that the product of successive group elements corresponds to a motion with the order of multiplication reversed. For example, G(0,£,,€ ,)G(0,¢2,22) induces the motion g(€,,2,)g(£ 1,1). We could have studied right multiplication instead of left multiplication. We would then have found the induced motion generated by the differ- ential operators D and D, a. nmin D,= ae + 16,5"0* Om (46) D;, = = 1a" Om: 2 60* By their very definition, D and D satisfy the following anticommutation relations {DaDi} = —2i643! Om _ 8 (4.7) {D.Dg} = {D;Dj} = 0, while D and Q anticommute {D..Q5} = {DaQj} = {Di0,} = {Dz} = 0. (4.8) We are now ready to introduce superfields and superspace. Elements of superspace are labeled by z = (x, 0,6). Superfields are functions of superspace which should be understood in terms of their power series expansions in 6 and 0, F(x,0,0) = f(x) + O6(x) + OZ(x) + 00m(x) + DOn(x) + O0"Ov,,(x) + OO0A(x) + BOOW(x) + B60Bd(x). (4.9) All higher powers of 0,0 vanish. The transformation law for superfields is defined as follows: SeF(x,0,0) = Sef) + 05:G(x) + D5Z(x) + 00 5gn(x) + FO 5.n(x) + 00D b:r_(x) + 00D 5H(x) + 800 5.b(x) + 6685 5-d(x) = (60 + CO)F, (4.10) where Q and @ are the differential operators (4.4). The transformation laws for the component fields (f, , 7, ...) may be found from (4.10) by TV. SUPERFIELDS 27 matching appropriate powers of 6,0. The commutator of these trans- formations satisfies (3.4) as a consequence of (4.5). It is easy to verify that linear combinations of superfields are again superfields. Similarly, products of superfields are again superfields because Q and @Q are linear differential operators. Thus we see that superfields form linear representations of the super- symmetry algebra. In general, however, the representations are highly reducible. We may eliminate the extra component fields by imposing co- variant constraints, such as DF = 0 or F = F*. Superfields shift the problem of finding supersymmetry representations to that of finding appropriate constraints. Note that we must reduce superfields without restricting their x-dependence through differential equations in x-space. Superfields satisfying the condition D® = 0 are called chiral or scalar superfields. This constraint does not yield a differential equation in x-space. Extra conditions, however, often give differential equations. For example, DD® = D® = 0 yields massless field equations, while Db = D® = 0 implies ® = a = constant. Vector superfields are defined to satisfy V = V*. It is possible to construct all supersymmetric renormalizable Lagrangians in terms of vector and scalar superfields. We shall treat both vector and scalar superfields in great detail in the coming chapters. It is always possible to construct a superfield from a component multiplet. We start with any component of the multiplet, say A, and apply the operator exp(6Q + 8Q), whose action is defined through (3.3). This yields a function of x,0,0 which transforms like a superfield F(x,0,0) = e2*9) x 4= A+ 5A 4°°° (4.11) We define the function 5:F(x,0,0) to be the power series in 0,0 whose coefficients represent the transformed component fields, 5 ®; + a;. The new Lagrangian has parameters: Ae = Ae + Mia; + GijnjAq ij mig = my + 2WGijnAe (5.15) ik = Jin If the old potential had a minimum at ®; = —a;, the new potential has a minimum at the origin. The new potential belongs to a supersymmetric Lagrangian with parameters given by (5.15). The class of renormalizable Lagrangians may be restricted by R- invariance. R acts on chiral multiplets as follows: RO(O,x) = €2!"* (o> !90,x) - ~ (5.16) RO*(B,x) = 2 *(0G,x). Here n is called the R-character of the superfield. For the components, (5.16) implies R:A > eA ya eri Day, (6.17) F = e2i- Dap Mass terms or potentials are R-invariant only if the R-characters of their respective superfields add up to one. REFERENCES J. Wess and B. Zumino, Phys. Lett. 49B, 52 (1974). L. O’Raifeartaigh, Nucl. Phys. B96, 331 (1975). EQUATIONS D,® = 0. (5.1) ® = A(x) + i060 0,,A(x) + 5 0000 A(x) + /26y(x) — 90 Ono" + OOF(x). (5.3) Va 34 V. CHIRAL SUPERFIELDS @* = A*(x) — i006"06,,A*(x) + 0008 A*(x) ZOU(x) + 5 0060" 0,ib(x) + BOFx). (85) v OO, = A(V)ALy) + J20[i(V)A(y) + A owl + OLA()FQ) + A)FO) — WWQ)]- 6-7) O00, = Aly)A(y)Ay) + 26d + WjA,Ai + WAiAj] + OOLF;AjAy + FjAA; + FyAiAy — Wallin — Wied — Was]. (6.8) O70, = oe + a (x)AF(x) + (20;(x)A (x) + OOAF(X)F (x) + DOFH(X)A,(x) + oe ur Sat — 6nA¥A)) — Wiz ja) + ooo = 645" AF Ont? — On APY) — Vi | Nv | + Boer] 6 "U0, A, ~ alta) + J2M | v2 1 1 sfie ac Aj + 7 474; 1 i, - i 5 OWAE OA, + 5 Oni) ~ 5 Vi ent 69) 1 L = OF Oo@5 component + ((; m0, 1 + Gin POO, + 1) + he. (5.10) 3 100 component L = ido"); + AFOA; + FFF; +[m,(ae, - 50) + anlAAsPs — val + AF; +he} (5.11) RO(6,x) = e"*@(e°*0,x), — RO*(Gx) = e-2™H*(e*9,x). (5.16) V. CHIRAL SUPERFIELDS 35 R: A > ei" 4 Wo etieaey (5.17) Fo e2itn- Dap EXERCISES (1) Compute Q, and Q, in terms of y",0,0: é ap + 2i0%0,, ae (2) Compute D,, D;, Q,, and Q, in terms of y*",0,0. (3) Derive the transformation laws (3.10) using Q,0 and (5.3) expressed in terms of the variables y", 0, and 6. (4) Define the components of a chiral superfield (D,) = 0) as follows: A = Wpea-o Y= D,®o-5-0 F = DDV|y_9-- Express these components in terms of the component fields 4,y,F of (5.3). Compute the transformation laws for ./,‘¥, and ¥ using using Q and Q in the following form: Q, = D, — 2io,4"0* on 0, = Dz + 210044 (5) Show that © = DDU is chiral for any superfield U. Relate the com- ponents of U to those of ®. (6) Show that the mass term 4m®® for a single superfield ® is R-invariant if and only if ® has R-character 4. Note this condition excludes linear and trilinear terms from the Lagrangian. VI. VECTOR SUPERFIELDS Vector superfields satisfy the condition VevV*. (6.1) As usual, they should be understood in terms of their power series ex- pansion in 0 and 6: V(x,0,0) = ow) + iOz(x) — i8Z(x) +5 5 OL Mix) + iN(x)} — BLM(x) — iN(x)] = 00"6v,,(x) + a + 1 aur] - 00209 + so anna + § ana] D1) + ; C(x i (6.2) The component fields C,D,M,N, and v,, must all be real for (6.2) to satisfy (6.1). The vector field v,, lends its name to the entire multiplet. We have chosen very particular combinations of fields as coefficients of the 060, 660, and 6060 components of V. Our choice was dictated by the hermitian field © + ©*, where ® and * are chiral fields: © + ot = A+ A* + 20" + OY) + OOF + DOF* + 100" 6,(A — A*) + 0000" 8,7 2 W000" a, + 4 4 0000 (A + 4%. (63) This combination has the gradient i@,,(A — A*) as coefficient of 00", motivating us to define the following supersymmetric generalization of a VI. VECTOR SUPERFIELDS 37 gauge transformation: ViV4+O04O%, (6.4) Under this transformation, C>C+A+4 A* roe — iv M+iN—>M + iN — 2iF Um > Un — 1O,(A — A*) Aaah DD. (6.5) The choice of components in (6.2) renders 2 and D gauge invariant. From (6.5) we see that there is a special gauge* in which C, 7, M, and N are all zero. Fixing this gauge breaks supersymmetry but still allows the usual gauge transformations v,, > v,, + 6,,4. [tis very easy to compute powers of V in this gauge: V = —00"Dv,(x) + i0007(x) — MO0%(x) + 5 60080 (x) Vea 5 00000," (6.6) ve=0. Thus we may view the vector field V as the supersymmetric generaliza- tion of the Yang-Mills potential. To construct the corresponding super- symmetric field strength, we observe that 2, and 7, are the lowest- dimensional gauge invariant component fields in V. They are also the lowest-dimensional component fields in lao. W, = ~;DDD,V a (6.7) =, I = W, = —~ DDD,V. = —Z PDD, * In the literature this gauge is often called the Wess- Zumino or WZ gauge. 38 VI. VECTOR SUPERFIELDS These superfields are chiral and gauge invariant. Chirality follows immediately from (6.7), DjW, = 0 D,W, = 0 (6.8) while Db = D&* = 0 must be used to prove gauge invariance: 1 We —{DDbAV + + 0) = W,~ 1 D{D,D,}0 = W. (69) It is easy to compute the components of W, in the special gauge (6.6). The computation is further simplified by use of the variables y = x + i000 ory” =x — i000: V = —00"6v,(y) + i002(y) — i860A(y) + 5 90800( y) = 10,0") = —00"0v,( yt) — i00Ay*) + i00BA(y*) + 5 000810 yt) + 10,u™y*)]. (6.10) The result is W, = —idfy) + [200 - 5 (oat) r awrats0 | 9, + 00645" On AY), W, = Hy") + [eae + 5 tas") fut) — erat”) | = 64j080""* Ohl V")- (6.11) The superfields W,,W, contain only the gauge invariant fields D, 4,, and Unn = Oy — OV, Furthermore, they are chiral and satisfy the additional ‘mn men constraint equation DW, = D,W*. (6.12) For @ = @ = 0, this relation simply expresses the fact that the component field D is real. Equation (6.12) may be verified component-by-component from (6.11) or directly from the definition (6.7). It may be shown that (6.11) represents the most general solution to the chirality conditions (6.8) and the constraint (6.12). VI. VECTOR SUPERFIELDS. 39 The superfields W, and W, are examples of representations which have the component fields 1, and 7, as fields of lowest dimension. We could have constructed W, and W, from 4, and 7, by applying the operator exp(9Q + 00) x, as described in Chapter IV. We would then have found W, and W, to be related through Eq. (6.12). Since W, is chiral, the 6 component of W*W,, ~ i WW = —2i20" oy — 50" py + D+ 50S pras (613) transforms into a space derivative. Note that W*W, may also be written as 1a ww, = “4 DDW*D,V. (6.14) From (6.13) we see that 1 — L = GZ (WW hoo + WW") (6.15) is the supersymmetric gauge invariant generalization of the Lagrangian for a free vector field. After some partial integration, this reduces to 1 1 4. = fatydt p? — 1 vy — ea 7 fa xP = fa {5 D? = Fy — io ani (6.16) This Lagrangian may also be obtained as a 0006 component: Jatxe = fara! W°D,V + W,D*V} (6.17) 9088 Equation (6.17) is equivalent to (6.15) because of (6.14) and the fact that D and 0/66 differ only by an x-space derivative. We can always add the mass term m*V? to the Lagrangian (6.17). This term is not gauge invariant and cannot be computed in the WZ gauge. Starting from (6.2), we find 1 V7\ooa = 5 Ome" — XA — ZA + 5(M? + N*) (6.18) 40 VI. VECTOR SUPERFIELDS It is interesting to note that this term not only gives mass to the vector field v,, but also introduces the additional degrees of freedom C and x required for a massive multiplet. The Lagrangian (6.17) together with (6.18) describes one vector field, two spin-} fields, and one scalar field, all of equal mass. REFERENCES A. Salam and B. Strathdee, Phys. Rev. D1], 1521 (1975). J. Wess, Acta Physica Austriaca, Suppl. XV, 475 (1976). EQUATIONS veve. (6.1) V(x,0,0) = C(x) + i8y(x) — i8Z(x) + 5 e9f mis + inca | - 5 aa| ms - inca] — 00" Dvy(x) + ooo 2 + 3a" enn] - io} 0) + 5 o” a.n)| + 5 0000 ES + 5 new} (6.2) V3oVEO4O%, (6.4) C3+C+A+ A* 1x — iD M +iN—>M + iN — 2iF (65) Up Uy — 10,(A — A*) As D3D w, = —1DDD,Vv = —q DDD, (6.7) S I | a S S Si = x VI. VECTOR SUPERFIELDS 41 DjW, = 0 D,W, = 0. (6.8) We = idan + [A000 ~ Florareants) ~ eras | + 80043" iy) W, = ity") + [sow + 5 63(B"O") Oye") — exrasy |" 2 — 645000"! Oily"). (6.11) DW, = D,W*. (6.12) W*Wlog = —2i0"C yf. — a + D+ fom tne: (6.13) 1 oh Wy We L = {(W*Wiloy + WeW"lo0). (6.15) 4 p 4+. 1 2 1 mn, ama TF fatxe = fates; p G0 — HO yi (6.16) EXERCISES (1) Prove [D;, {D,.D,}] = 0, [D,,D;D"] = —4i0,;"0,D, and Ion g [D?,D7] -iD*s,;"D*d, — 20 iD*o,,"D*6, + 20. (2) Compute the 80 component of D,V in the general gauge (6.2) and the special gauge (6.6). (3) Compute e” in the WZ gauge. 42 VI. VECTOR SUPERFIELDS (4) Show that the constraints (6.8) and (6.12) yield the following equations for the component fields: D = D*, vm = penn, ante O04 = 0. (5) Compute (6) Use the definitions of W, and W, to verify (6.12). (7) Derive the Euler-Lagrange equations for the Lagrangian (6.17) + (6.18). (8) Use (6.13) to show fated Ww + Wy) = 4 fatxweWyo . 4 a| a 0: 2 2/00 (9) Find the supersymmetry transformations for the gauge invariant fields in the vector multiplet: = i[l€o" 0,2 + %5"a,A) — (no m)] iD + o™E0nn 5D = FG" yh — Eo" Oy. = & I m RS UI VII. GAUGE INVARIANT INTERACTIONS In this chapter we discuss the gauge invariant interactions of chiral and vector multiplets. We start with the U(1) case and later generalize our results to non-Abelian gauge groups. Chiral superfields ®, transform by a phase under global U(1) rotations, 0, = etd, (7.1) The t, are the U(1) charges appropriate to the ®,, and 4 is the rigid U(1) rotation angle. The t, and 4 are real constants. Constants are chiral superfields, satisfying the constraint equations D,A = D,j = 0. From (7.1), we see immediately that the ®/ are chiral superfields as well. It is easy to construct a Lagrangian invariant under (7.1) for constant parameters /: L= Leet Lrg, Lx, = 0} V,looae 1 1 Lope = E m0, + joao, || 06 + he. 0 Note that U(1) invariance requires m,; or gj, = 0 whenever t; + t; or t, +t; + | #0. In the literature, the term #p.g, is often called the superpotential. Equation (7.1) takes one chiral superfield into another when 4 is a constant chiral superfield. When / depends on x, the situation is slightly more complicated. In this case, 4 must be promoted to a full chiral multiplet: , = ec", D,A= oy = el" O,*, Dt = Only then do the ® remain chiral superfields. 44 VII. GAUGE INVARIANT INTERACTIONS The Lagrangian (7.2) is not invariant under such local transformations. In particular, Yp_,, remains invariant, but Zx , does not: Od, = 07 Oe", (7.4) It is easy to see that Lx, may be rendered invariant by introducing the vector superfield V with its transformation law (6.4): Vi=V+i(A— A‘), (75) With this addition, the full Lagrangian 1 — L= 4g (WWaleo + WW' la) + OF e',|oa95 1 1 + 6 mb; + 40) 00 becomes invariant under local U(1) gauge transformations. At first, (7.6) looks non-renormalizable. It may, however, be evaluated in the WZ gauge, where V> = 0: + ne | (76) * &”Olgoan = FF* + ADA* + 10,00" + (5 yao ee ) — = (Alp — A*AW) + (0 = soe) A*A. (7.7) V2 2 2 In this gauge, the Lagrangian contains no terms of dimension higher than four. The supersymmetric extension of electrodynamics is constructed in terms of two chiral superfields: =e O,, OF =e@. (7.8) In components, the Lagrangian 1 — - Loev = glow + WWhas) + Die", |aoa5 + Die” °'D _|poaa m(O , |g + DLO* fea) (79) VIL. GAUGE INVARIANT INTERACTIONS 45 becomes L =!p—!, v™" — iho" 6 QED = 5 4 Umm \ + F.F% + F_F* + AOA, + A*OA- + OD OW, + gh) + eo |b. 5 i... i inn i, +p AE OA — 5 OAGA, — 5 AE OWA + sata. | — “(4h — At 2 — ADD 4 A*W_A) V2 e 1 + 7 PLAT A+ — A*A_] —- ge eat Ay + A*A_) + m[A.F_ + ALF, — Wale — U.0_ + ASP* + A*P*]. (7.10) From (7.10) we see that the two Weyl spinors w,,\/- combine to form one massive Dirac spinor, the electron. It is straightforward to generalize the transformation law (7.1) to non-Abelian compact groups: DM =e 0, Mt = Ore, (7.11) In (7.11), A is a matrix: Ay = TiAa (7.12) The matrices T’ are the hermitian generators of the gauge group in the representation defined by the chiral field ®. In the adjoint representation, we normalize our generators as follows: Tr T*T? = kd”, kk > 0. (7.13) With this convention, the structure constants ¢“” (TT?) = it™T° (7.14) are completely antisymmetric. 46 VII. GAUGE INVARIANT INTERACTIONS. The Lagrangian (7.6) is invariant under non-Abelian gauge transfor- mations, provided we extend the transformation law (7.5): eM = eA eV Gid, (7.15) In (7.15), both A and V are matrices: A, = T2A, ‘a> Vy = THY. (7.16) With Hausdorff’s formula, we encounter only commutators of group generators in computing the product of exponentials in (7.15). Evaluating the commutators by the group commutation relations (7.14) allows us to express V’ in the following form: Vi=TV,. (7.17) This shows that the transformation law (7.15) is independent of any specific representation for the generators T*. Furthermore, the transfor- mation law starts with a term independent of V, Vi=VHEiA-Atyte, (7.18) so non-Abelian theories also allow a WZ gauge where V> = 0. Equation (7.15) may be evaluated for infinitesimal gauge transforma- tions with the following form of Hausdorff’s formula: e4eB = eAt Eas: Bt coth{as2) BLt (7.19) This expression contains all terms linear in B. The Lie derivative £4. -B is given by [4A,B]. The hyperbolic cotangent in (7.19) must be understood in terms of its power series expansion, where seoraeoftlsE-Le}-T 0m with n factors +A. Using (7.19) to evaluate (7.15) yields 6V = V' — V = ify [(A + At) + coth(£y)2) (A — A*)]. (7.21) VII. GAUGE INVARIANT INTERACTIONS 47 The supersymmetric field strength W* [Eq. (6.7)] may be readily generalized to the non-Abelian case: W, = —~ DDe~"D,e". (7.22) al In (7.22), the vector superfields V are matrices, as in (7.16), with the gen- erators in the adjoint representation of the gauge group. It is easy to verify that W, > Wi, = e-!wyei® (7.23) under non-Abelian gauge transformations. The proof is left to the reader as an exercise. We are now ready to write down the most general Lagrangian for the supersymmetric renormalizable interaction of scalar, spinor, and vector fields: g Tr(W?Walog + We Wap) + B* e¥ Ploooa oi ~ 16kg? 1 1 + [6 m,P@; + + 4x00.) + ne} (7.24) au Gauge invariance requires the mass matrix m,; and the coupling constants Gijx to be totally symmetric invariant tensors with respect to the internal symmetry group. The normalization of the gauge-field kinetic term is chosen to recover the canonical normalization for the component action after scaling V > 2gV (see Exercise 7). REFERENCES J. Wess and B. Zumino, Nucl. Phys. B78, 1 (1974). S. Ferrara and B. Zumino, Nucl. Phys. B79, 413 (1974). EQuaTIONS wo, = e *A@,, D,A =0 @,* = elt *@,*, D,A* =0. (73) Vi=V+i(A — A*). (7.5) 48 VII. GAUGE INVARIANT INTERACTIONS L = 7{(WWoo + WW laa) + OF eO-looae Bie 3 1 1 + IG 0,0; + $40.) + he} (7.6) 08 Ot’ Dgq99 = FF* + ADA* + id,o" 1 i i + w( Yow + ZAP OA - sata) oe Aap — A*A) + 5 (1 - sow") ata, (7.7) v2 1 wy Lor = g VW eo + WWlaa) + OLE’ , |ooag + DLE _|yoa9 + m(O,O_|g + OL H+ las). (7.9) 1 3D] emt ido" Oyd+ FFE +F_P*+AtOA, +42 0A. L orp = a ee ea i i i i +7 A4 GAs 5 ALA. 5 AE AA+5 ata.| ie > Tah FAM AD T+ AWA) e 1 +5 DIAtA, —AtA_]—3 Pout A, +t) +m[A,F_+A_F,—Wih_-V.W_+AS FE + A* FE]. (7.10) Tr T°T? = kd”, kk > 0. (7.13) (T*,T"] = it*T°. (7.14) oF = eA ogi, (7.15) Ay = TiAe Vy = THe (7.16) VII. GAUGE INVARIANT INTERACTIONS 49 Lio iy vo W, = —q Pde D,e". (7.22) W,> Wi=e Wei, (7.23) _ ol © 16kg? 1 1 + [6 m0; + 540) Tr(W*Waloo + WW lag) + D* e Popes + he} (7.24) 00 EXERCISES (1) Show that Yp z, of Eqs. (7.6) and (7.24) is also invariant under sym- metry transformations (7.11) with complex parameters. (2) Show that e~”D,e” = D,V — $[V,D,V] in the special gauge (6.6). (3) Demonstrate that W;, = e~'*W,e'*. Use the fact that D,A+ = 0 and D,A = 0: 1 __ iat + W, = 4 DDIe*eve'* D,ew* Xe] i | i =e weld — qe *D{DD,je" =e Ayes, (4) Construct an SO(3) invariant interaction using three chiral vector fields @4. (5) Use the multiplication properties of the Pauli e-matrices (n-6)(m-o)= n-m + i(n x m)-¢ to show that for infinitesimal values of b, ein -eeibm 7 — (cosa + io -nsina)(l + ibm-o) = cos[a + (n: m)b] + in-o sinfa + (n- mb], where p =n + b[(m — (n-m)n) cota — n x mj, and nm are unit vectors. (6) Use the result of Exercise (5) and [n o,m- o] = 2i(n x m)-o to verify Eq. (7.19) for the special case e’"’7e°"'°. Remember that bis infinitesimal. 50 VII. GAUGE INVARIANT INTERACTIONS (7) Expand 1 2 = Teg? Tr(W*W,loo + W,W* laa) + O* e¥ Olooas in the WZ gauge. Assume the gauge group is non-Abelian, and restore the coupling g by rescaling V > 2gV, L= - Uo — iH GA + Spb = GA* Dy — Wa" Db + FOF + if2g(At TOPs — TT Ap) + gDA + TA, where DyA = OA + ign@TA Dy = Opi + igh Ty Duh = Oy hk — gt vO Hoh = Oye? — Oy — guid (8) Compute the transformation laws for A, y, F, v®, 2 and D® in the WZ gauge. Use them to verify that the result of Exercise (7) is supersymmetric: 6A = /2Ep Oe = in/20"ED,,A + /2EF OF = i220", + i2gT ART 60 = —TGME + EoMA 5A = o™E® + iEDO 5D = —Eo"Q, J — 9, jqME, VIII. SPONTANEOUS SYMMETRY BREAKING If supersymmetric gauge theories are to find realistic application in high energy physics, both supersymmetry and gauge symmetry must be broken spontaneously. The spontaneous breaking of ordinary gauge symmetry is well understood, but supersymmetry imposes additional conditions which need further discussion. These restrictions rest on the property 1 = = = H = 7 (0:01 + 0:0; + 9.0, + 929;), (8.1) derived from the algebra (I), Equation (8.1) tells us that <‘P|H|¥> = 0 for every state |W). Furthermore, it tells us that states with vanishing energy density are supersymmetric ground states of the theory. Such states are ground states because the expectation value of H may never be negative; they are supersymmetric because = 0 implies Q|0> = Q|0> = 0. Ground states of zero energy preserve supersymmetry, while those of positive energy break it spontaneously. This situation is sketched in Figure 8.1. In this chapter we shall discuss three models which exhibit the general properties of spontaneous symmetry breaking in supersymmetric theories. We first consider a supersymmetric model, constructed from chiral superfields, in which the ground state breaks supersymmetry. We know from Eq. (5.14) that the potential energy in such models takes the form W = FxF,, where F, is given by Fy* = —(Ay + mA; + Gij-AjAj)- (8.2) Vacuum expectation values a, of A; for which F, = 0 signal super- symmetric minima of the potential. To break supersymmetry, we must choose special values for the parameters ,, mj, and gj, such that the equation OS A + Mya; + Gina; (8.3) 52 VIII. SPONTANEOUS SYMMETRY BREAKING tv ve >y > (a) (b) Ficure 8.1. The ground state of (a) preserves supersymmetry, while the ground state of (b) breaks it spontaneously. has no solution in a;. Such models have been constructed by O’Rai- feartaigh in the paper cited at the end of Chapter V. He found that three chiral superfields are required to break supersymmetry, the simplest model being given by Loy, = {LUM + m&,0, + gho,0,] + he}. (8.4) Fayet and Iliopoulos have shown how to spontaneously break super- symmetry in gauge theories with Abelian gauge groups. They observe that the 0000 component of the vector superfield is both supersymmetric and gauge invariant. They add this term to the Lagrangian (7.9) and find that it spontaneously breaks supersymmetry: 1 _ L= givin + W,W*) + OF e"@, + OF e %O, + m(O,O, + OF OF) + IV. (8.5) In this model, the potential is given by 1 Y= 5D + FFE + PFE, (8.6) VIII. SPONTANEOUS SYMMETRY BREAKING 53 where D, F,, and F, are solutions to the Euler equations: D+K+S(AtA — AEA.) = 0 F, + mAt=0 (8.7) F, + mA¢ =0. There is no solution to (8.7) which leaves ¥ = 0, so supersymmetry is broken spontaneously. Let us examine the potential (8.6) in more detail. Substituting for the auxiliary fields, the potential Y becomes 1 1 1 V= 5K + (m= + peat + (m - ys) AtAs 1 + geUAtA, — ATA. (8.8) We must distinguish between the two cases m? > ex and m? < }ex. When m? > ex, both A, and A, have real masses. The model describes two complex scalar fields, one of mass m,” = m? + 4ex, the other of mass m,” = m? — sex, as well as three spinor fields ,,W>,2, and one vector field v,,. The masses of the spinor and vector fields are unchanged by the symmetry breaking. In particular, the field ; retains its mass m, while 4 and v,, remain massless. Note that m,? + m,? = 2m?. The vector field v,, plays the role of gauge field for the unbroken U(1) symmetry group, and / is the Goldstone fermion arising from spontane- ously broken supersymmetry. From the transformation law for 4 (Exercise 6.9), 5A = iED + O™EVnn (8.9) we see that / transforms inhomogeneously as soon as D acquires a vacuum expectation value: Od = -ibK +00 (8.10) This identifies A as the Goldstone fermion. Non-zero vacuum expectation values of auxiliary fields induce the spontaneous breakdown of super- symmetry. 54 VIII. SPONTANEOUS SYMMETRY BREAKING When m? < 4ex, A, = A, = Ono longer minimizes the potential (8.8). To find the minimum, we must solve the equations av 1 2 aa = (w= + se) + Fut — AEA,)A, = 0 1 wy ' ; (8.11) a 2 Sat = (ne - 308)As - GTA, — A$A,)A, = 0. 3 This givesa minimum at A, = 0, A, = v, where ex? + (m? — Sex) = 0. By a gauge transformation, ~ may be chosen to be real. Expanding the potential around its minimum spontaneously breaks the U(1) symmetry. In terms of A = A,, A = A, — », the potential becomes: V= 2m? (ex — m?) + 2m?A*A e +4 (leee tu a) + alae” Wri + The constant os?) et +++ (8.12) Nie Nie 2 2 x (ex — m?) e is positive; both supersymmetry and gauge symmetry are broken spon- taneously. The vector field v,, acquires a mass by eating the Goldstone boson field (A — A*) )/./2, leaving the total number of degrees of freedom unchanged. The symmetry breaking also modifies the spinor mass terms: =m(hbs + Dis) + WE a2 = yo. (8.13) 2p With the following linear combinations, (8.14) VIII. SPONTANEOUS SYMMETRY BREAKING 55 the mass terms become diagonal: —_|m +5 (up + 0). (8.15) The Goldstone spinor 7 remains massless. Note that 7 transforms in- homogeneously, = 21 lek m+, (8.16) as expected for a Goldstone field. This model describes two spinor fields of mass \/m* + }e*v*, one vector field and one scalar field, each of mass ,/4e*, one complex scalar field of mass 2m’, and one massless Goldstone spinor. Note that the sum of the masses squared weighted by the number of degrees of freedom is identical for the bosonic and fermionic modes: 2 1 2,2 2 1 2,2 2: 2m +4 ee = 4m +e . (8.17) This is also true for the U(1) symmetric case described earlier. In fact, such relationships between bosonic and fermionic masses are common in supersymmetric theories. The situation for the Fayet-Iliopoulos model is sketched in Figure 8.2. Non-vanishing vacuum expectation values of auxiliary fields induce we a (a) (b) Ficure 8.2. (a) When m? > 4x, supersymmetry alone is broken. (b) When m? < 4ex, both gauge symmetry and supersymmetry are broken. 56 VIII. SPONTANEOUS SYMMETRY BREAKING supersymmetry breaking, while non-zero vacuum expectation values of dynamical scalar fields lead to the breaking of gauge symmetry. After having seen a model in which supersymmetry and gauge symmetry are broken spontaneously, one might wish to construct a model in which only the gauge symmetry is broken. We first discuss such models with chiral superfields. In this case we must find a solution a; to (8.3) which is not left invariant under the internal symmetry group. As a simple example, we consider the group U(1) with three chiral superfields: one neutral, one positive, and one negative. The Lagrangian 1 Ley = zmo + pO,0_ +140 + gOO,0_ + he (8.18) is U(1) invariant. The Eqs. (8.3) become A+ma+ga,a_=0 a_(u + ga) =0 (8.19) a,(u + ga) = 0. This set of equations has two solutions: (l) a, =a_ =0, a=-- (2) a,a_= (2 - me), a= af (8.20) The first does not break the U(1) symmetry, but the second does. In the second solution, only the product a,a_ is determined. This stems from the fact that “pis invariant not only under the U(1) group, but also under its complex extension. For any solution a,,a_ to (8.19), there exists an entire class of solutions, e*a,,e~*a_, for arbitrary complex 2. The ground state has a larger degeneracy than required by the initial symmetry group. If we gauge the Lagrangian (8.18), we must introduce the vector super- field V, coupling to ®, and @_ as in (8.5). This results in the following trilinear coupling between the scalar fields A, and the vector multiplet V: eV(A*A, — A*A_). (8.21) VIII. SPONTANEOUS SYMMETRY BREAKING 57 For the symmetry breaking solution, this contributes a piece to the D- term: vata, —a*ta_ + 2%) (8.22) Such a term would ordinarily break supersymmetry. Because of the degeneracy a, > e**a,, however, it is possible to transform away this term for any choice of x. In this model, D-terms do not induce the spon- taneous breakdown of supersymmetry. The mass term associated with (8.20) is given by 1 ze lata, + ata_)V?. (8.23) Tt cannot be transformed away. It gives a mass to the vector field v,,. Comparing (8.23) with (6.18), we see that spontaneous gauge symmetry breaking in supersymmetric theories gives rise to an entire massive vector multiplet. This is the supersymmetric extension of the Higgs-Kibble mechanism. These models are easily extended to non-Abelian symmetry groups. Supersymmetric solutions require Fe = -Ay — maa; — Ginetta; = 0. (8.24) The parameters A, m, and g are restricted by the internal symmetry group. In gauge theories, supersymmetric minima must also satisfy D! = af Tha, = 0. (8.25) The Fayet-Iliopoulos D-term is not gauge invariant and cannot appear in the non-Abelian sector of supersymmetric models. In the remainder of this chapter we shall show that (8.24) determines the supersymmetry breaking of non-Abelian theories. That is, if (8.24) has a solution a;, then it is always possible to find a solution a; which satisfies (8.25) as well. We shall demonstrate this for the case of a semi- simple gauge group G. To begin, let us suppose we have found a solution a; such that F(a) = 0. We may then compute d! = af T4a,. (8.26) 58 VIII. SPONTANEOUS SYMMETRY BREAKING The vector d’ specifies a certain direction in the regular representation. There is always a group element which transforms this vector into a linear combination of vectors in the Cartan subalgebra. Because (8.24) is invariant under G, this transformation rotates the a, into another solu- tion G. The vector d’ transforms into a vector d’ whose non-vanishing components lie in the Cartan subalgebra. We may now perform a linear transformation within the Cartan subalgebra such that the direction a defines a single generator with eigenvalues y;. In this basis, the only non- vanishing component of d° is d: d = a} ua,. (8.27) The equations (8.24) are also invariant under gauge transformations with complex group parameters. This is because the complex conjugate representations of the scalar fields never enter F*. We are free to perform such a transformation in the direction d: a, = exp(un)a;. (8.28) The parameters 4; solve (8.24) for all values of y. Taking 7 real, we find f uje?™";,. (8.29) We now distinguish two cases. In the first case, all the y; (for which G; # 0) are of the same sign, say positive. We then let 7 + —oo to find d = 0.In the second case, the 4; take both signs. We shall show that there is still a value of 7 where d = 0. In particular, we note that d= ante a. (8.30) Nie Considering @;' e?"""G, as a function of y, we see that it tends to +00 as n > +00. Therefore it has a minimum for some value of y. At this point the derivative vanishes and d = 0. This completes the proof. We have shown that spontaneous supersym- metry breaking in non-Abelian models is controlled by F-terms. Super- symmetry is spontaneously broken if and only if the equations F* = 0 have no solution. This is the O’Raifeartaigh mechanism for supersymmetry breaking. VIII. SPONTANEOUS SYMMETRY BREAKING 59 REFERENCES P. Fayet and J. Iiopoulos, Phys. Lett. 51B, 461 (1974). E. Witten, Nucl. Phys. B188, 513 (1981). EXERCISES (1) Show that A, = A, = 0 is a minimum of the O’Raifeartaigh model (8.4) when m? > 2dg. The value of the potential at the minimum is 22, independent of Ay. (2) Compute the boson and fermion masses in the O’Raifeartaigh model: Real scalar masses: 0, 0, m?, m?, m? + 2gd Spinor masses: 0, 2m. Massless scalars and spinors are general features of O’Raifeartaigh models with spontaneous supersymmetry breaking. Also compute OF¥/0A, in this model, and show that OFe det (7) =0 This is another general feature of O’Raifeartaigh models, (3) Consider three chiral superfields, ®), ®,, ®,, with R-characters No = 1,n, = 0,n, = 1. Construct the most general renormalizable, supersymmetric, R-invariant Lagrangian also invariant under the following discrete transformation: ®) > 0, > -, ®, > -,. Show that this determines the O’Raifeartaigh model. (4) Show that % of (8.14) does not shift under a supersymmetry transformation. (5) Add a D-term 2xV to the Lagrangian (8.18). Determine the values of a, and a_ at the minimum of the potential. 60 VIII. SPONTANEOUS SYMMETRY BREAKING (6) Consider three triplets of chiral superfields: ®,, i = 1,2,3. Find the minimum of the potential for the Lagrangian 1 Src, = MO, O, + 5 G&D; (D; x D,). (7) Show that the minimum of the potential of Exercise 6 is invariant under the rotation group with complex parameters. (8) Gauge the model of Exercise 6 and show that an arbitrary D-term may always be eliminated. Supersymmetry may never be broken by the Fayet-Iliopoulos mechanism in this model. IX. SUPERFIELD PROPAGATORS In previous chapters we have found superfields very useful for the con- struction of supersymmetry representations and invariant Lagrangians. In this chapter we shall see that they also simplify the calculation of radiative corrections in quantized supersymmetric theories. The Feynman tules for supersymmetric theories may be stated in terms of superfield vertices and propagators. Many component-field Feynman diagrams are contained in one superfield diagram, so many miraculous cancellations between component diagrams are manifest in one superfield diagram. For this reason alone one would like to find a superfield formulation of supersymmetric theories. To derive superfield propagators we must first introduce the concept of integration in superspace. An indefinite integral over a Grassmann variable 7 is defined as follows: fan = 0, nan =1. (9.1) Any function of 7 is polynomial, f(y) = c + An, so definition (9.1) ex- tends immediately to arbitrary functions of Grassmann variables: f(n) = ¢ + An fr (n)dn =A (9.2) fr (nn dn = c. Since é alm =0, partial integration is always possible. Note that integration and differ- entiation give the same result on functions of Grassmann variables. 62 IX. SUPERFIELD PROPAGATORS Delta functions are defined by the integral frensenan = f0). (93) From (9.2), it follows that on) =n, (9.4) so consequently, (1) d(n) = 0. (9.5) Defining volume elements in superspace, a0 = 1 d6* d0" =-q bag a0 = —1 dB, dB, (06) d*9 = d’0d70, we find 09470 =1, 00070 =1. (9.7) This allows us to write the Lagrangian (5.10) as an integral over superspace: g= f Jere, + m0 0) 1 ; 1 : +5 meds OF 5(0) + 3 Aix ® ©, 5(8) + FA? OF 0 woh areas, (9.8) Perturbation theory in superspace may be developed as a direct ex- tension of ordinary perturbation theory. In particular, one would like to calculate superfield Green’s functions, . (9.9) IX. SUPERFIELD PROPAGATORS 63 From these one recovers the component-field Green’s functions by power series expansion in 01,6! - -- 6",6". As with any field theory, we begin our analysis by evaluating the free- field two-point functions, the propagators. For chiral fields, these are derived from the free-field part of Lagrangian (9.8): “=f joro+s mo 3(0)+4 med? o>" wo} aroara = ADA +1 OQFO PE tm APA 3 04-500). (9.10) In components, we find: CO|T{A()A*(x')}]0> = iApox — x’) COT {A()F(X)} 0) = = —imA,(x — x’) OITEFOQF*(<)}[0> = 1 Ap(x — x’) (9.11) = 19,Fm Ap(x — x’) COLT T*C ph p()}[0> = 1d%pm A(x — x’) COLT (aD pl'}}|0> = 25" Op Ap(x — x’), where 1 A-(x) = Goa All other two-point functions vanish. We may use these component propagators to construct the superfield propagators. For example, = + 00<0|T{F(»)A(y)}|0> + 20°F OT {Wal YW pl ¥)}]0> =—im(0 — 0)? Ay(y — y). (9.12) 64 IX. SUPERFIELD PROPAGATORS From the definitions y = x + i0c0, y* = x — i0o6, we see that this propagator and the ®* ®* propagator have the following x,x’ dependence: = iexpli(dc") + 0’c"0' — 200"0')0,,*JAp(x — x’). (9.14) With these propagators we may evaluate the superfield Green’s func- tions (9.9) to any order in perturbation theory. We start by writing the n-th order contribution in terms of free superfields: (ofr fora.o-0 see * (x?* 1get get 1) ) fGeebepvater + f Salon's} = (rfo--0 1 < J 5 [0%1.8,.8,)50,) + gO" %G0.04.0,)80,)] d*x', d?0, a0, -- lo). (9.15) Using Wick’s theorem, we then reduce these expressions to the usual Feynman diagrams. As a sample calculation, let us consider the one-loop corrections to the superfield two-point functions. These are illustrated in Figure 9.1. Diagram 9.1(a) is proportional to 67(0 — 6’) = 5(0) = 0, while 9.1(b) goes as 6°(6 — 6’) = 5(0) = 0. This shows that all contributions to mass renormalization, both finite and infinite, cancel between the various com- IX. SUPERFIELD PROPAGATORS 65 Od rae , (a) ——_8/ \o ® \, bd Jf ® oot = oN. at / (ob) ———4 \o or \ get J o* oot (ec) ar \a’ @ \ a gt eat FiGuRe 9.1. One-loop corrections to the (a) ©®, (b) ®*@*, and (c) ®O* superfield propagators. ponent fields. The final diagram, Figure 9.1(c), is proportional to f d*xd*x' d20.d26' d20 d20' 58) 5(6') x ©(x,0,0) exp[i(00" + 0'o"0’ — 2008’) d,,*] x Ap(x — x’) exp[i(00"8 + 0'0"0’ — 200"0') 6,,*] Ap(x — x’) x O*(x',0,0) = fatxd*xd?0d70A, 7c — x')®(x,0,0) exp[ — 2100" 6," ]® * (x',0,8). (9.16) To obtain this result we have integrated over the 5-functions, replaced & by 6, and integrated by parts. The A,” in the expression above leads 66 IX. SUPERFIELD PROPAGATORS acts / an oP) g/ \ \ ] XN 4 Sole notNy + ~ 7 \ ® Or \ tat TT 1® ® \ / ‘LU? Fiure 9.2. All tadpole graphs vanish identically. to a logarithmic divergence which may be absorbed into a logarithmically divergent wave function renormalization. It is easy to see that all closed-loop diagrams vanish when they contain only ®® or ®** propagators. This follows immediately from the fact that they are proportional to 6(0) in 6,0 space. In particular, there are no non-vanishing tadpole graphs in this theory (Figure 9.2). Similarly, there are no finite nor infinite contributions to the coupling constant renormalization (Figure 9.3). The superfield propagators (9.13) and (9.14) may be obtained directly as superspace Green’s functions for the free-field equations. To see this, we write the free-field Lagrangian (9.10) in the following form: Ly = fiero 7 an(oo +o De )pataroera 1 © =J5 wo ya(e d*xd’6d"6, (9.17) where lm -45 PD 1 M= : 1 i" pp in IX. SUPERFIELD PROPAGATORS 67 o* + ot o* SONS 8} \ eter \ / otet Fiure 9.3, There are no non-vanishing corrections to the ? and *# couplings. This expression is valid because dO is equivalent to —4DD under an x-integration, and because 74(DDDD/D) is a projection operator on chiral fields: 1 DDDD 16 O A DDDD _ 16 ~ o©=0 ifDo=0 (9.18) D 0. 68 IX. SUPERFIELD PROPAGATORS If we wish to derive field equations from (9.17) by a variational principle, we must take into account the fact that the chiral fields ® and ®* are subject to constraints. We do this by varying ® and ®* in the y and y* bases: 6 re Sow 2H) = Hy = yO ~ 8). (9.19) In these bases the field variations automatically remain chiral. We may use this result to find the variations of © and ®* under full superspace integrations: BV F(x! 0°B) dx 426 20" wean a F(xc,0,8))d*x! d26 28 = eH D f O(y'O0F(y' — i0'o8, 0, Wy d*y' 26 a0" = [oly = yd — OV (y" — 16°08, 6,8) dy’ a76'd7O = fro — i608, 0,8) d20 = -4 DDF(x.09). (9.20) Here (x,0,0) = ®(y,0), where y = x + i0c8. Equation (9.20) may be summarized by a formal rule: 6 tag. pe a , SOG, GB) 29.8) = —ZDDHO — 050 — Bde — x). (9.21) This rule reproduces (9.20): socom 1 0B VF(x',00) d*x' d?0' a0 —q DSO — 050 — 8) d(x — x)P(0,8) d*x' a0 dO y DF(x,0,0). (9.22) ale Here we have integrated by parts to obtain the final result. IX. SUPERFIELD PROPAGATORS 69 The free-field Euler-Lagrange equations are found by varying (9.17) according to (9.21): 1/D? 0 oO -—F Y =0. 23) io pe)#(g-)=° (9.23 These equations may be simplified with the help of (9.18): 1 m® — 4 DDo* =0 1 (9.24) b* ——DD® = 0. m q D Here we recognize the field equations for a massive chiral multiplet, first encountered in Chapter IV, Exercise 7. We may always couple chiral superfields to classical external sources. For chiral sources, D,J = D,J* = 0, (9.25) we find the following Lagrangian 1D? 1 ® [- Go 8 yy G= é +), + 4. 929 72) Ji 5@o a (5. Hoe ) (7.) d*xd?0 a0 0-45 (9.26) and field equations: 1/D? 0 oO J Ho sdels)-(2) The superfield Green’s function is defined in analogy to (9.27): 50 — 060 — D) d(x — x’). (9.28) 70 IX. SUPERFIELD PROPAGATORS, The 5-functions are multiplied by the operators —}D? and —4D* because (9.27) has a solution only for chiral sources. The superfield Green’s function A gives ® and ®* in the presence of J and J*: 2 12 o\ (0) = — [Ac0:x'0'8) _ ( jas ee. 0-55 (9.29) In order to solve for the Green’s functions, we exploit the algebraic properties of the chiral projection operators 1 D?D? 1 D*D? =e a” 2= 76 (9.30) by introducing three additional operators: D? D* 1 P,=— ~=—, Pp=—-~=DD*D. 31 += P a T a DD?D. (9.31) After a short calculation, one may quickly confirm P,+P,+Pr=1, (9.32) as well as the following multiplication table: (9.33) With this multiplication table, we may readily express the differential operator of Eq. (9.28) in the following form: 1/D? 0 1/D? 0 P, 0 i pt - ilo (5 r) 634 Using (9.34), it is easy to show that A= (3 o )as — 656 — 8) d(x — x’) (9.35) 0 P, IX. SUPERFIELD PROPAGATORS. 71 is the Green’s function for (9.28). The additional projection operators insure that A is the propagator for chiral superfields. To find A, we must first invert .#. This is easiest if we expand it in terms of the P-operators: -A 0 0 0 M = PL + PL 0 0-4 01 + ( 0) + P, + Pr}. (9.36) The inverse of any operator of the type X = AP, + DP, + BP, + CP_ + EPy (9.37) is given by X~' = [A— BD"'C]"'P, + [D — CA“'B]'P, — A~'B[D — CA~'B]'P, — D-'C[A — BD“'C]"'P_ + E7'Py, (9.38) provided A,D,E are all invertible. This may be shown by direct multiplica- tion or by use of the P-operator representation given in Exercise 7. With this result, we find ° oO MU) = o\Petf Gam |p. Om? 1 0 0 0 + m 0 a m oFo= me ~{oo a Game t Pet Pr Gy 72 IX. SUPERFIELD PROPAGATORS, According to (9.35) and the multiplication table (9.33), the propagator A becomes d(x — x')6(0 — O68 — 0) = d(x — x')d(0 — 0')5(0 — 0). (9.40) To compare this result with the previous propagators (9.13) and (9.14), we must compute the spinor derivatives of the 5-functions: D,*(0, — 03) = —4 exp[ —i(0, — ,)0°0, 6,"] D,*O, — 92)? = —4 exp[i0,o°(0, — 83) 4,"] D,?D,7(0, — 03)°0, — 92)? = 16 exp[i(0,0°0, + 0,0°0, — 20,0°03)0,"] D,?D,?(0; — 92)°(0, — 92) = 16 exp[—i(0,0"0, + 050°0, — 20,0"0,)0,"]. (9.41) The proofs of these relations are left to the reader as exercises in straight differentiation. Substituting (9.41) into (9.40), we find a t a ( a) (x = x’), (9.42) where Ay, = —md(0 — 6’) exp[i(40"8 — 0’0"6')6,] Ay. = exp[i(0o"0 + 0's"6' — 200°6')d,] Ax, = exp[—i(40"0 + 0'c"6' — 20'o"D)é,] Az. = —md( — 0’) exp[ —i(00"0 — 6'o"8')0,]. (9.43) This result is identical to (9.13) and (9.14). In (9.42) we replaced 0 [0] by ' [8] whenever it was multiplied by 6(0 — 0’) [5(0 — 0’)]. IX. SUPERFIELD PROPAGATORS 73 Having gained some experience with superfield methods, we shall now compute the propagator for vector superfields. We start with the usual Lagrangian, 1 1. L= a W*Waloo + 4 WW'lag + m?V? lavas. (9.44) as outlined in Chapter VI. To this we add the gauge fixing term —30(D?V)(DV). This term yields a piece proportional to (0,v")? in the component Lagrangian. To find the propagator, we write the action as an integral over superspace: L= J {i WW,00) +5 W500) + meV? —§ wr yorr) hava 0479 1 8 =f; VDD*DV + mt vv — ViorD? + Dror hatnaraaD = [VL-OP +m P, +P, + Pr) EP, + Pa) a4 0470 = furvatxd?d70. (9.45) The Euler-Lagrange equations are found from a variational principle, NV =0, (9.46) and the superfield Green’s function is defined in the usual way: NA = d(x — x60 — 060 — 9). (9.47) Note that we choose to invert / and not 2.V as might be expected from the Lagrangian (9.45). This normalization of the superfield propagator leads to the usual normalizations for the Green’s functions of the com- ponent fields. Solving for A, A = N~'6(x — x)6(0 — 060 — 0), (9.48) inverting 1, 1 1 Nis Sam? + moe (Pi + P2), (9.49) 74 IX. SUPERFIELD PROPAGATORS and using (9.41), (P, + Pz) d(x — x’) 5(0; — 82) 50, — 92) “a5 (D,?D,? + D,?D,7) 6(x — x’) 6(0;, — 02) 56, — 6.) = 55 ExPLil6 0°, — 0,070.) Q1[4 +10, ~83)°0, ~9,)*] ox — x), P(x — x) (0; — 83) 5(0; — 92) =(1— Py — P2) d(x — x’) (8, — 82) 5; — 8) = a exp[i(,0"0; —9,0°02) O,] x [4-11; — 0,)°@; —82)7] d(x — x), (9.50) we find the propagator for the vector superfield: 1 7 A= 0 exp[i(0,0°0, — 0,0°8,)0,] 1 x jis ae [4 - Cate, - 0,)80, - 9,)] -¢ = oe [4 + D6@, — 0,)66, - aloes — x2). (9.51) REFERENCES K. Fujikawa and W. Lang, Nucl. Phys. B88, 61 (1975). S. Ferrara and P. Piguet, Nucl. Phys. B93, 261 (1975). EQUATIONS L = [foro + 5 moro 3) + pmvoror oc] a°0a79 = A*DA + 10,Ho"y + F*F + m( ar + A*F* — su - a) (9.10) IX, SUPERFIELD PROPAGATORS 75 COT{AQ)A*(x)}]0> = TAp(x — x’) COT {A(X)F(x')} 0) = COT LA*(x)F*(x')}]0> = —imA,(x — x’) = 15,?mAp(x — x’) (9.11) LOT Px) lx)}]0> = io*pm Ap(x — x’) COTY CIT G(x)}O) = op" Oy Ape — x’) Ap() = =——s. KO|T{O(x,0,8)0(x',0',0)}0> = —imd(0 — 0') exp[i(0o"0 — 0'0"0') d,,*] Ap(x — x’) = +imé@ — 8) exp[—i(00"O — 6'0"B) 0,7] Aplx — x’). (9.13) = iexp[i(o"0 + 0'o"0 — 2066’) d,*]Ap(x — x’). (9.14) Ly = [foro - in(oo +o 2? 2 o* harxaroare =f 5(O.0" jd (o.)atsaroara, -.= pp 1 M = lm 1 “ap? (9.17) 6 y 1, ") 5 PS , Sonn B) = —7 DDOO — 050 — B)d(x — x). 21) 1., 1 0 jo - i ° p) “a= - 1 | 88 = 8980 — Bya[x — x. 0-40 (9.28) 16 IX. SUPERFIELD PROPAGATORS 1 D*D 1 D*D? = =— 9.30 P=e » P= 0 (9.30) D? dD? 1, Ph=qa, P= aa Pr=—g5DD*D. 9.31) P, 0\ 4, 5 — BP , A= M-*5(0 — 050 — B)5(x — x’). (9.35) 0 P ™ pp ‘ gp? 1 4 A=— _ yn _ e509 7 Goml; om d(x — x')d(0 — 658 — 8). —D*D> — p? 16 4 (9.40) 1 ~ lou, - =f {i ww, 6O)+5 W,W* 5(0) +m? VED yoy) ded 00D =f (V[-OPrtm(P, +P, +Px)—E(P, + PIV) dx d204°0 = [vrvatedod9. (9.45) WA = (x — x')5(0 — 6')50 — 8). (9.47) A=— 4 exp[i(0.0°0, —0,0°8,) d,] 1 : = x “ae [4-19(0, -0,)66, ~8,)] “= am [4+ 150, - 0) 60, 1,1} 0 x2). (9.51) EXERCISES (1) Use definition (9.1) to show that 7' = an implies dy’ = a7! dy. (2) Check that J f(1) d(n — p)dn = f(p). IX, SUPERFIELD PROPAGATORS 77 (3) The bosonic part of Lagrangian (9.10) may be written as + A 7 m (A wya(s) where a = (0 ") Show that the 44*, AF, A*F*, and FF* propagators are given by the inverse of this operator. (4) Compute the ®** and ©@* propagators. (5) Verify that the kinetic part of the chiral Lagrangian may be written as follows: f (x,0,0) * (x,0,8) d*x d20d?20 = ! (x,0,0)e~ 29" mqp*(x,0,0) d4x d20.d20. (6) Show Jreca)a*xa2oa0 = [ (-4) DDF(x,0,0) d*xd20 = f ( -1) DDF(x,0,0) d*x d?0. (7) Confirm the multiplication table (9.33) and show that these operators have the following matrix representation: 100 00 0 P,={|0 0 0 P,=(0 1 0 000 000 0 1 0 00 0 P,={0 0 0} P_={1 00 000 0 0 0 0 0 0 P,={0 0 Oj. 001 (8) Prove (9.38) by direct multiplication, using (9.32) and (9.33), or by using the P-operator representation given in Exercise 7. 78 IX. SUPERFIELD PROPAGATORS x=(é > show that the inverse matrix X~* is given by ZY X= > Z=(A—- BDC)? (9) For any matrix X, where U = -D"'CZ V =(D— CAB)! Y = -A"'BY, provided A~! and D~! exist. Compute the inverse if only B~! and C7t exist. X. FEYNMAN RULES FOR SUPERGRAPHS In this lecture we shall derive the Feynman rules for the supersymmetric ®? model, = [eoroo" w+ {fae E mo? +j00'| + nol. (10.1) These rules may be applied to all chiral models and extended to super- symmetric gauge theories as well. We shall find that the effective action may be expressed in terms of one d*9 = d?0 70 integration of the follow- ing form: f d*0 f tx dx F (6.0.9) ++ Fy(%ps00)G(1, «65 %,)- (10.2) The function G(x,,...,X,) is translationally invariant and the F’s are products of superfields and their derivatives. No factors of ~' appear in the F’s, so for chiral operators the d*0 integration cannot be converted into a d?0 integration without introducing spacetime derivatives (see Exercise 2). This leads to the surprising result that mass and coupling terms of the d?0 form are not renormalized in supersymmetric theories. Furthermore, no higher-dimensional momentum-independent chiral operators are induced in the effective superpotential to any order in perturbation theory. Equation (10.2) also implies that all vacuum-to- vacuum diagrams vanish. This is because expressions of the type (10.2) without any superfields are immediately annihilated by the d*0 integration. Before deriving the Feynman rules we will give a short derivation of the generating functional ZLLJ*] = (9 olf exp i [a*odtx [10(-j2 Joa (10.3) 80 X. FEYNMAN RULES FOR SUPERGRAPHS. for superfield Green’s functions Gt. Mi cM tt Ny é 6 6 6 _ pn } ——— ON Te ere are (10.4) Here Z,[J,J*] is the generating functional for free superfield Green’s functions and z' = (x‘,6',6') is an element of superspace. Equation (10.3) may be verified explicitly in terms of component fields. We shall take another tack and derive it directly with superfields and superspace techniques. In the previous lecture we calculated the free-field two-point functions: i KO|THo(z)o(z’)]0> = Ay s(@2') (x I ® KOT (2)3(2)]0> = a Aa a(252') (x — x’) (10.5) COfTHo(z)G (z')|0> = Aj 2(z,2') 50¢ — x’) OTH 5 ()OY(2)]0> = = Aa la") Ce = »°). The right-hand side of the equation includes the matrix elements A;; of Eq. (9.43). These two-point functions may all be obtained from the free generating functional ZoldS*]= (9 ! T expi Jaoa'stuea(-j 2) es + rel -4 7) sia 0 1 . =exp — 5 fateatx ate d*x! 5 m \ \ / x 4 ((2),J*(2)) 1p? a7 0 (2 x =, 1D? 0 i (10.6) X. FEYNMAN RULES FOR SUPERGRAPHS 81 This functional generates all free-field Green’s functions as sums of products of two-point functions A. With the help of (9.40), (9.18), and a few integrations by parts, we find ZolJJ*] = exp — + sf d*x d*0 d*x’ d*6'(J(2),J*(2) )deaa2)( 74 & an (10.7) where Aggs is the propagator introduced by Grisaru, Roéek, and Siegel: /m D? des(e2) = [79 o(z - Z). (10.8) We may differentiate Z) with respect to J and J* using the rule (9.21): 5 1 5J@) 409 4409 1 An An / i]s |Z — fatx ato 5 +(e ni )ats x) \ar@)| 1 \( Ds(2) (NR) From this we find a functional equation for Zo 6 1/D? 0\ 1] 8G J(2) i ae ; = (j10) zo (10.10) Ie) Here we have used (9.28) and (9.18). We may generalize this equation to the case of interacting fields. For the ©? model, the field equations 1/D? 0 o oe J 10 Bea(S)-A)-() my 82 X. FEYNMAN RULES FOR SUPERGRAPHS lead to the following equation for Z: jo (P isa) 1/? 0\ 1/ J@\~ Je * 7 dJ(2), ilo Dt 3 |*7 (Sy) 8 pis) J * (2) ( ate) (10.12) Note that we have introduced projection operators in (10.12). We could have done this in (10.11), but there it is obvious that ® is chiral P,®=0, Pot =o, (10.13) The chirality of the functional derivative is less explicit, so we choose to keep P, and P, in (10.12). To solve for Z, we first compute the commutator 6 1 6a, , (gts) ver] =2(-!)(rogls) ae 1\,, 6). 7 (-3) {(r, uaa) ale — ah. (10.14) The last step is possible because DP, = 0. Integrating over d70d*x, } 6 \ [ Jeroats(, ue sia): 19] = 3 fatod*x ole — (P55) 5 \2 = a, aa) . (10.15) and using i ffavao(e, : ) ype eer tt) = J(z) + ol I : (10.16) X. FEYNMAN RULES FOR SUPERGRAPHS 83 we cast Eq. (10.12) in the following form: i 6 1/2? 0\ 1| 6 1 fates 4 ( 0 pt i) 6 Zne 8I*(2) Sxl (6 xo ls (10.17) This shows that 1 fatvtal, ZolJJ*] =e Z[J*], (10.18) since the right-hand side satisfies the free equation (10.10). No normaliza- tion factor is needed because of the fact that all vacuum-to-vacuum diagrams vanish. With (10.18) we have proven (10.3) and solved for the generating functional of an interacting chiral supersymmetric theory. Having found the generating functional, we shall now derive the Feynman rules. We begin by recalling the relation between the Green’s functions and the generating functional: GM (zt... Mp cM th Ny 2.8 8 8 OJ, Oy Ol yay OS *y Kay K! 56 x f dx Ping (F i) ZolIJ*] =(-i" (10.19) dast= The factors of 66 4K fa x! bu 7) generate vertices at z*. The derivatives in Y,,, act on previous derivatives and on Z, itself. Each derivative acting on Z, creates a new propagator at 2X. Each derivative acting on a previous derivative connects an existing propagator to 2X. In this way every new vertex is completely saturated with propagators. 84, X. FEYNMAN RULES FOR SUPERGRAPHS As an explicit example, we consider a term in ° theory in which two new propagators are created at the point z: . 5 i fated “oft 23 sia} 2 . 6 1 N 4.1 4407 = i fatsaroft P, aal{(44 >) fatx ate —— x G2 (2 — Z)J(2!) + (2 — 2ue)| x [(-3 >) fata —— x (3 DP se — Zule") + le — rte) |2o 4 / 1, 5 = ifarnatof P, ath x [ f dtx'd*0' — = (FE ae — 2)slz) ee 1 ngage a8 + (2 — ZW ch) |(-5>") fate a0" 5 x (FP ae - 22") + H(z — 2rees) [2 (10.20) Here we have used (10.7) and (10.8) for Z,. The last step (changing the a6 to a d*0) was possible because of the chirality property of each factor. Note that we also used the fact that i m D*D? —-mil6 O f dx! d40" (2 — z)s(z) = 0. (10.21) ate ) Such a piece corresponds to a closed ® tadpole in a Feynman diagram. The proof that (10.21) indeed vanishes is left to the reader as Exercise 7. The effective action is computed from the one particle irreducible (1PI) Green’s functions. In general, 1PI diagrams have at least two internal lines leaving every vertex. The external legs of the 1PI diagrams are amputated with inverse propagators. They are then multiplied by the superfield amplitudes ®(z) or *(z). This leads to the following Feynman X. FEYNMAN RULES FOR SUPERGRAPHS. 85 rules: (1) For each external line, write a chiral superfield ®(z), ®*(z). (2) At each © vertex with two [three] internal lines, include factors of —4D? acting on one [two] internal propagators. At each ®*? vertex, include similar factors of —4D?. (3) Write a factor of 4g for each vertex, and integrate {d*+xd*@ over each vertex. (4) Use Grisaru-Roéek-Siegel propagators for O®, O**, and O*® internal lines. These are given in (10.8). (5) Compute the usual combinatoric factors for an A? theory. This is most easily done directly from (10.3) and (10.4). Let us now use these rules to follow the 0-integrations around an arbitrary closed loop. The Feynman rules and the GRS propagators combine to give an expression of the following form: (D,?)(D,7)* 5(12)(D27)*(B7)* 6(23) - « (D,”)*(D,7)" (nl). (10.22) The exponents ¢;,k; are either zero or one, and 5(12) = 4(, — 8,)5@, — For a general loop, the D and D factors might appear in the opposite order. However, any higher powers of D? and D? may be reduced to the above form, up to powers of (): D?D?D? = 16D? _ 10.23 D?D?D? = 16[)D?. ( ) Of course, for the effective action, the above expression is multiplied by superfields for external legs and GRS propagators for adjoining closed loops. It is also integrated over d*x, d*0, --- d+x,d*6,. The final expres- sion is evaluated by removing the D and D derivatives from one 6- function after another by partial integration. This introduces new derivatives on the lines that leave the loop. It also introduces a certain number of derivatives on the last 5-function, say 5(n1). All but one of the 6-integrations may be performed with the aid of the 6-functions 6(12),..., 0([n — 1]n). This leaves a factor of f4*0,(07¥ D250, — 0,)50, - 9), _, (10.24) or fate D7(07y60, — 0,)60, - 9,)|,._, 01 =, (10.25) 86 X. FEYNMAN RULES FOR SUPERGRAPHS These expressions vanish unless k = ¢ = 1. In this case, we find fat0,D°D? 56, — 0,)50, - 0,)), _, a= 6, = fae, D*D? 50, — 0,)6@, — 8, , - 41 = 9 = 16 J d*6,, (10.26) as follows from (9.41). The whole loop in 6-space has shrunk to one d*0 integration. This process can now be carried over to the next loop and we finally arrive at the result (10.2). Note that (10.2) is true for each diagram, and as a consequence, for any particular sum of diagrams as well. REFERENCES T. Grisaru, M. Roéek, and W. Siegel, Nucl. Phys. B159, 429 (1979). A. Ovrut and J. Wess, Phys. Rev. D25, 409 (1982). EQUATIONS _ (pern0+ zl tg? a toe? ¢=[rorso os tfa af ame + jao'|+ nel. (10.1) 2" Z[JJ*] = (rex i f a*od'x| 19(—F2-)or) 1D? + ro(-j2)ore fo) Hh u( 2. =o (Gas despa". (10.3) G21, My Met 2M) =(-pvO 8 + =O Te aejarety Ie TT, (10.4) Zell" = exp — 5 fatea*0d's a°0102). 5°12) Benlee)( 7) (10.7) X. FEYNMAN RULES FOR SUPERGRAPHS. 87 Aars(2.2) EXERCISES (1) Show that (10.2) is supersymmetric. (2) Use (9.18) to show that Jersatoo(—j0") J = fax@oons for ® and J chiral. (3) Demonstrate that the generating functional (10.6) gives the two-point functions (10.5). (4) Compute ll? wr) | (5) Verify un (6) Compute ——— Zo. Use this to check (10.20). wi ) (7) Prove that (10.21) does indeed vanish. (8) Use the Feynman rules to calculate the diagrams of Figure 9.1. Show that (a) and (b) vanish and that (c) leads to a wave function renormalization. XI. NONLINEAR REALIZATIONS In this chapter we shall study a nonlinear realization of the supersym- metry algebra. This will introduce us to superspace differentials and provide a natural transition to differential forms. It will also demonstrate that supersymmetry may be realized entirely in terms of fermion fields. In fact, we shall construct a local supersymmetric Lagrangian from a single fermion field. We shall see that this nonlinear Lagrangian is highly non-renormalizable. It does not, therefore, change the pattern of Bose- Fermi symmetry in renormalizable supersymmetric field theories. Nonlinear transformations for fermion fields are reminiscent of the nonlinear transformations for the Goldstone spinors in Chapter VIII. We shall see that the nonlinear Lagrangian gives rise to spontaneous supersymmetry breaking and that the fermion field is indeed a Goldstone spinor. The nonlinear Lagrangian is quite useful for studying the super- symmetric Higgs effect in supergravity theory. The supersymmetric Higgs effect occurs when the spin-} Goldstone fermion combines with the spin-3 partner of the gravitational field to form one massive spin-? field. To derive the nonlinear transformation law, we first consider the supersymmetry transformation (4.3): x’ =x + i(0c& — Ea6) O=0+€ (11.1) O=0+8. This transformation induces a nonlinear realization on the spinors 0 and 6. We shall generalize this transformation to arbitrary spinor fields A(x) by drawing an analogy between 0 andi, 0 = ka: 2x!) = Ax) + te (11.2) ale od A(x!) = A(x) + XI. NONLINEAR REALIZATIONS 89 From (11.2) we may compute the changes of the fields at the same space- time point: 4 aH , 1 F T4 4 6% = 2%(x) — A(x) = xk & — ik(do™E — E02) 0,27 (11.3) ; > z l, = eomqya 7 Ohi = Hx) — Fy) = Fy — ie(ho™E — 602) Oh K After some algebra, which we leave to the reader as Exercise 1, we find (6,62 — 5¢5,)4% = —2ilno"E — Eo") Ay A*. (11.4) This verifies that (11.3) does indeed realize the supersymmetry algebra (I). Before constructing an invariant Lagrangian, we first examine the differentials dx, d0, and dO. These transform as follows under general coordinate transformations in superspace: x'™ = x'™(x, 0, 0) 0" = 6x, 0,8) 0, = Ox, 0, 8) ym n ox” Vv ex’ m é x dx'™ = dx + d0 i + ; —— WO, ge wo” 00" a" a0" = xt + AO ay + AO (11.5) a. = ax" oO , 8, @ 0", ae ce + Oe Here one should note the summation convention for the spinor indices and the placement of the differentials to the left of their coefficients. For (11.1), this becomes dx™ = dx" + id00"Z — iéo" dO do" = do" (11.6) a’, = dB ,. 90 XI. NONLINEAR REALIZATIONS It is easy to find a combination of differentials e” = dx" — id00%O + i00°dO et = de" (11.7) e;, = dd; which is invariant under (11.1) and (11.6): eo = dx — id0’o"O' + i0'o" dd" = dx" + id0o%E — iéo*d@ — id0o°O — idOo%E + i0o'dd + ito°db =e. (11.8) Substituting 0 = «/ and dO = «(0A/@x") dx” into (11.7), we find e* > dx"[6,,° — ik? 0,A0% + ik?20% 0,2] = dx™A,". (11.9) A short calculation shows bAn’ = ix(E0" 0,4 — 6,A072)A," — ix(Ao"E — E0"/)0,A,°. (11.10) With this transformation, the Lagrangian 1 = —~ > de 111 gL Tye det A ( ) yields an invariant action 6, det A = det ATrd,AA™! —iké,,[(ao"~ — Eo") det A]. (11.12) From (11.9), we see that Y describes one massless spinor /: Loi - . £=-55- 5 Aa" oy — @,402) + [interaction terms]. (11.13) The constant « spontaneously breaks the supersymmetry. It also leads to a non-vanishing vacuum expectation value for the Lagrangian. This gives rise to a cosmological constant when (11.11) interacts with a gravi- tational field. XI. NONLINEAR REALIZATIONS 91 REFERENCES D. V. Volkov and V. P. Akulov, JETP Lett. 16, 438 (1972). S. Deser and B. Zumino, Phys. Rev. Lett. 38, 1433 (1977). EXERCISES (1) Use (11.3) to compute 5,524 = —i(no™— — Eo™H) Oy A — K7{(Ao"E — Eo™A)(Ao"H — No"2) 00,4 + (ho"E — Eo" )\(Ando" — 10" OA) Oy + (Ao" — n0"A)(d,A0"E — Eo" 0,2) dpa. Note that the terms in (11.3) quadratic in 2 come from a shift in the argument x. Verify the closure relation (11.4). (2) General transformations in superspace x™ = x(x’, 0, 0’) Oo” = 0x’, 0, 0") 8, = 0,(x', &, 8) induce the following transformations on the partial derivatives: a _ ox" 2 oO 8 wD, 0 dx” — Ox™ Ox" * Oe OO" * Ox" OD, ee 30" — BO ax" * GO" ao * G0" TU, oe OO OO Ox" 60", 00" 60’, 00, Show that this, together with (11.5), gives 6 a é é é m a —d™ a dx' Gen + 6! aot d,, 30, 7 dx’ aem + do a0" _ @ + d,—. i “O;, 92 XI. NONLINEAR REALIZATIONS (3) Write e*,e%,e, of Eq. (11.7) in the following form: e = dx™e,” + d0"e," + d0,0" & = dx"e,* + d0"e,* + db,e! 5 = AX, + AOC, + dO,e%;. Show that Cn’ = On’, = — i920", el’ = - 160, ent = 0, et = 6,2, el = 0 Cm = V5 eu, = 9, ey = Oy. (4) Compute 5A," = ix(Eo" 0,1 — 0,A0°%) + «3(Ao™E — Eo) 8,,(1070,2 — 0,40°%) + 1(0,40"E — Eo A,2)(Ao" Og — Onda“). Use this to prove (11.10). (5) Use 0/éx™ det A = det A Tr 6,,4A7! to verify (11.12). XII. DIFFERENTIAL FORMS IN SUPERSPACE Supergravity theories have been successfully formulated in terms of dif- ferential forms in superspace. This is not surprising, for supersymmetry transformations are among the general coordinate transformations of superspace. It is natural, therefore, to introduce supergravity in a way which is manifestly covariant under such coordinate transformations. This leads us to extend the concept of differential forms to superspace. The elements of superspace are denoted by 2M ~ (x"0"0,). (12.1) The capital letter M represents the four-vector index m as well as the spinor indices u and ji. M, m, and yw are all upper indices, while fi is a lower index. Elements of superspace obey the following multiplication law: MN = (—yim2N pM (12.2) Here n is a function of N and m is a function of M. These functions take the values zero or one, depending on whether N and M are vector or spinor indices. Exterior products in superspace are defined in complete analogy to ordinary space: dM ~ dzN = —(—y'™dz® ~ dM dzMzN = (—ym2N dM, (12.3) With this definition, differential forms have an obvious extension to superspace: Q = de nv a deMWy,..(2)- (12.4) The differentials are written to the left of the coefficient function and the indices are labeled in such a way that there is always an even number of indices between those being summed. From now on we shall drop the 94 XII. DIFFERENTIAL FORMS IN SUPERSPACE symbol ~ for exterior multiplication. This is not ambiguous because we know no other way to multiply forms. Functions of the superspace variable z are called zero-forms: F(z). (12.5) One-forms are written as A = dzMWy(z) = dx"W,(z) + d0"W,(z) + d0,W*(z), (12.6) while Q in Eq. (12.4) is a p-form. Note that the definition (12.3) leads to coefficient functions of mixed symmetry. Thus, in contrast to the usual case, there is no value of p above which all forms vanish. We shall always assume that coefficient functions with an odd number of spinorial indices are fermionic in character, and that those with an even number of spinorial indices are bosonic. These assignments repro- duce the familiar rules for the multiplication of forms: (c:Ay + C2A,)Q = cyA,Q + c,A,Q AQ = (—)PIQA (12.7) A(QE) = (AQ)E. Here we have assumed that A is a p-form and Q a q-form. Having defined superspace forms, we must also introduce exterior derivatives. Exterior derivatives map zero-forms into one-forms, dF = at oF = do 0yF, (12.8) and p-forms into (p + 1)-forms, Q = d+ dW, u,(2) (12.9) dQ. = dzMs-- + dzMv d2N ar Wu, ---m(2)- In general, exterior derivatives have the following properties: d(Q + XZ) = dQ + dd d(QE) = QAE + (—)dOE (12.10) dd =0, XII. DIFFERENTIAL FORMS IN SUPERSPACE 95 where & is a q-form. Equations (12.8) and (12.10) follow immediately from (12.3), (12.4), and (12.9). Alternatively, it is possible to define exterior derivatives through (12.8) and (12.10). This is done in Exercise 5. Equations written in terms of differential forms and exterior derivatives are covariant under coordinate changes. To see this, let us assume that y and z represent two sets of superspace coordinates: yM = yM(z). (12.11) Functions of y have a natural mapping into functions of z: F(y) = F(y(2)) = $*F(2). (12.12) If we maintain that y and z label the same point in superspace, the defini- tion of @*F in (12.12) guarantees that a certain quantity takes the same value at the same point, independent of labeling scheme. In a similar fashion, @* induces a natural mapping between p-forms in the two coordinate systems Q(y) = dy dyM Wag, aes(¥) ,, Oy , Oye = (e J vd FO) aya) Oz" Oz = dz... dzXvp* Wy, -.. (2) = p*A(2). (12.13) The map $* enjoys the following properties: (1) $*Q + Z) = PQ + PFE (2) @*(QZ) = ($*Q)($*Z) (3) d(@*Q) = b*(dQ). (12.14) The proofs of (1) and (2) are straightforward. The proof of (3) is left as Exercise 10. These properties make a formalism based on differential forms and exterior derivatives automatically covariant under coordinate changes. The mappings (12.12) and (12.13) simplify for infinitesimal coordinate transformations: zM = yM 4 EM, (12.15) 96 XII. DIFFERENTIAL FORMS IN SUPERSPACE In particular, we find OF(z) = $*F(2) — F(z) 12.16) = —6 dF) rg for zero-forms and OWr(2) = 0* Walz) — Wal) 1: age = ~ EO, Wule) ~ Sn Wale) (12.17) for one-forms. These expressions may be easily generalized for arbitrary p-forms Q. Gauge theories are not only covariant under general coordinate trans- formations. They are also covariant under a local structure group. This is a compact Lie group for Yang-Mills theories and the Lorentz group for gravity theories. In general, differential forms span a representation of this group: Q" = 2X, "(2) (12.18) Q = Ox. The index a runs from 1 to L, where L is the dimension of the representa- tion X of the group. Objects which transform linearly under a representation of the struc- ture group are called tensors. Note that exterior derivatives do not map tensors into tensors: dQ’ = QdX + dQX. (12.19) A connection must be introduced to compensate for the inhomogeneous term QdX. Connections are Lie algebra valued one-forms @ = dMoy'(2)iT" (12.20) with the following transformation law: gb = X-1pX — X-1ax. (12.21) In (12.20), the matrices T are the hermitian generators of the structure group, and r runs over the dimension of the algebra. XII, DIFFERENTIAL FORMS IN SUPERSPACE 97 Connections allow us to define covariant derivatives, QQ = dQ +O, (12.22) or, more explicitly, DQ = dXDyQ = dz... dzMv dz é W, 2 aie zie dS x Way ms(2) + dz. ++ dzMed2\oy' Way, ...m,(QiT" (12.23) for Q a p-form. Covariant derivatives map p-forms into (p + 1)-forms and tensors into tensors: BQ! = dQ’ + Q'¢' = OdX + dQX + QX(X~'HX — X~'dX) = (dQ + QG)X = (GQ)X. (12.24) There is one tensor which can be constructed from the connection and its derivatives. It is called the curvature tensor: F = db + $9. (12.25) The curvature tensor is a Lie algebra valued two-form: 1 F= 3 dM d2® Fy y(z) (12.26) Fyu(2) = Fya'(2)iT". Its transformation law is computed in Exercise 8: F = X"'FX, (12.27) The curvature form and the covariant derivative of a tensor are, in general, the only tensorial quantities which may be constructed by taking derivatives. Higher derivatives lead to identities (and not to new tensors) because of the fact that dd = 0, These identities are called Bianchi identities. 98 XII. DIFFERENTIAL FORMS IN SUPERSPACE Bianchi identities of the first type are found from the covariant derivative: dDQ = Qdh — dQ = QF — $$) — (GQ — QO)p = OF — 9Q¢. (12.28) These may be written as follows: DDQ = OF dz d2XDyDyQ = 5 dz™ dz FyyQiT’. (12.29) Bianchi identities of the second type are found from the curvature form (12.25): dF = pdb — dog = GF — b9) — (F — bp) = oF — Fo. (12.30) These tell us DF =0, (12.31) or, in terms of the coefficient functions, dz dz' dz" DF yy = 0. (12.32) Summing over all permutations of the indices, and using the fact that Fru = —(—)""Fuw, we find DpFyy + (—POOMI Fags + (MOT yy = 0. (12.33) This is the superspace generalization of the usual cyclic identity on the curvature. REFERENCES H. Flanders, Differential Forms, New York, Academic Press (1963). F. A. Berezin, Sov. J. Nucl. Phys. 30, 605 (1979). XII, DIFFERENTIAL FORMS IN SUPERSPACE EQUATIONS Mw (x"045,). 2M2N = (—yim2N2M, dzM , dz’ = —(—)"™dz® a dzM aM = (=ym A aM, Q = dz n+ ++ dzM?Way,..- (2) (cyAy + C2A3)Q = cyAyQ + c2,A,Q AQ = (—)PIQA A(QE) = (AQ)E. a dQ = deMt ++ deo de Wy, (2) d(Q +3) =dQ+ dB d(QE) = QA + (—)"dOE dd =0. Q' = Ox. g = X~'oX — X~'dX. DQ = dQ2+ AG. BA = (BX. F=do t+ oo. F' = X"'PFX,. QDQ = QF 1 aM dS DyDyQ = 5 dM d2XF yy QT". 99 (12.1) (12.2) (12.3) (12.4) (12.7) (12.9) (12.10) (12.18) (12.21) (12.22) (12.24) (12.25) (12.27) (12.29) 100 XII. DIFFERENTIAL FORMS IN SUPERSPACE QF =0, (12.31) dM dz dz'D,Fyy = 0. (12.32) EXERCISES (1) Show that the p-forms on an ordinary n dimensional manifold span an (5) dimensional linear space. (2) In three dimensions, show that dd = 0 implies V x V- = 0 and vV-Vx =0. (3) Verify é aa 2 = (yim om oO zM zN 62% @zM" (4) Use Exercise 3 to show that dd = 0 holds for forms (12.9) in super- space. (5) Demonstrate that (12.8) and (12.10) define dQ as in (12.9). (6) Check that the connection ¢ remains Lie algebra valued under the transformation (12.21). (7) Show that if @ is Lie algebra valued, (12.25) implies that F is Lie algebra valued as well. (8) Prove that the curvature F transforms like a tensor (12.27) under the structure group. (9) Show that Z9Q = OF gives (DrIu — (—Y"IuPn)W(2) = WEym for W a zero-form. (10) Compute #*Q, d(*Q), dQ, and *(dQ) for an arbitrary p-form Q. Verify that d($*) = @*(dQ). XIII. GAUGE THEORIES REVISITED Before beginning our study of supergravity, we shall examine supersym- metric gauge theories in the language of differential forms. We will repro- duce our previous results and gain confidence in geometrical methods. Whenever possible, we will follow the steps we later take in formulating supergravity theories. In this way we will treat supersymmetric gauge theories as a model for supergravity. In the previous chapter we introduced differential forms and exterior derivatives in superspace. We used the superspace differentials dz™ as a natural basis. We could, however, have chosen any other basis, dM Ey “(z). (13.1) Here E,,“(z) is an arbitrary invertible function of superspace, Ex“(2EAM(2) = dx" - - (13.2) E,N(2)EN*(2) = 54°, where 5," 0 0 byY=|0 46,” 0 (13.3) 0 oO 6" In (13.3) it is important to note the position of the dotted indices. The dz basis is not particularly useful for supersymmetry because the exterior derivative é M dzM (13.4) > does not map superfields into superfields. This is because the differential operator 0/dz does not commute with the supersymmetry generators (4.4). 102 XIII. GAUGE THEORIES REVISITED A more natural basis is defined by the supersymmetry covariant derivatives (13.5) These differential operators commute with the supersymmetry generators {D,,0,} = {D,,0"} = {D*.Q)} = {D*.0"} = 0 and map superfields into superfields. (13.6) The exterior derivative may be written in terms of the differential operators (13.6) if we introduce a new basis e4(z) = dzMey,4(z) such that A — M Ag N 6 = e*D, = dzMey*e, awe where , O Dy = eg’ =. 4 = Oa BN The matrix e,™ follows directly from (13.5): em = 6" eK = 0 ey = 0 eg! = |e," = io,z"0* , em = i0*o nel Its inverse is given by Cnt = Spf Cpt = 0 eng = 0 ew’ =e," = —i,;°0" ef = 6,7 ey =0 I ef = —i0%o,:4e% el = 0 elt; jl (13.7) (13.8) (13.9) (13.10) (13.11) XIII. GAUGE THEORIES REVISITED 103 These matrices define supersymmetric flat space. Note, however, that the exterior derivatives of the basis forms do not vanish: det = dz dz® oy en “(z) det = — ie ,8e% (13.12) de* = 0 de, = 0. This is the price that must be paid in the flat space basis. To discuss gauge theories, we must introduce a connection ¢. As usual, the connection is a Lie algebra valued one-form: b = de bu = eo a ‘ (13.13) b4 = b4iT". We shall make contact with ordinary gauge theories by demanding m'|o=0=0 = Ym (13.14) The field v,, is the familiar Yang-Mills vector potential. The curvature two-form is defined as in (12.25): F = do + o¢ 1 , =5 dM dz®Fyxy = 5 oF ee (13.15) In the flat space basis, this becomes: F = e4e®Dgh, + deth,s + eth yerde 1 detha + 5 e*e"[Daba — (—)"Dabs — baba + (-)" babs]. (13.16) 104 XIII. GAUGE THEORIES REVISITED The coefficient function Fg, may be decomposed into its Lorentz- covariant components: Poa Oya — Gaby — [hoa] Fry, = Onb, — Dabo — [bnG.] Fo = Ons — Dido — [ba] (13.17) Faz = Dybs + Diaby — (Ppa) Fig = Diba + Dihs — 1p.) Faz = Dabs + Diby — (Ppha} + io pa"Pa- Note that Fyalo=a=0 = Usa iT". (13.18) The Bianchi identities may also be decomposed into their Lorentz- covariant components. In particular, DF =0 ; 1 1 1 (349) GF = 3 EDF 54 + 38 de® F 4 — 5 dete*F x4 gives () DF ya + PyFuc + Ge» = 9 (2) DF ye + DyFeg + DF = 0 (3) Fy + DF + DFin = 0 (4) DF pa + -~ DF =0 (5) DF ig + - G:F 4 = 0 (6) oF, + Dike - oF, + 2ioyi"Fac = 0 (1829) (1) DF oy + DpFay + DaF yp = 9 (8) D;F py, + DpFay + GaP ig + io g"Fag + 2i6,;"Fag = 0 9) DF ig + DpFiy + Bak yp + io f"F yy + 2ioy<'F yj = 0 (10) DF yy + DoF, + G:F i = 0. In (13.20), the derivatives Z are the full gauge-covariant derivatives. XIII. GAUGE THEORIES REVISITED 105 Each tensor component of F represents a full superfield multiplet. These multiplets contain a large number of component fields. Most of the component fields are superfluous and must be eliminated through con- straint equations. The constraint equations must be gauge covariant, Lorentz covariant, and supersymmetric. In addition, they should not restrict the x-dependence of the component fields. Finding the proper set of constraints is not easy. It turns out that Fig = Fig = Fp = 0 (13.21) gives the right results. We shall solve the Bianchi identities subject to these constraints. Without them, we would have found (13.17) as the most general solution. Identitites (7) and (10) in (13.20) are automatically satisfied because of the constraints. Identity (8), however, yields a further restriction on F: Gaf"Fap + Opi" Fax = O- (13.22) The vector-spinor F,,, has spin-3 and spin-} components. Equation (13.22) tells us that the spin-} component vanishes: Fy, = id a4pW! -_ i (13.23) Ww = 5 FF, Identity (9) gives a similar result, Fay = —iW ops i (13.24) wee 5 F,,6%, while identity (6) allows us to express F,, in terms of W and W: Fun = —y OPT pF an + Dok i) (13.25) 1 = {Pe.0W — 90,5,W). Exploiting the antisymmetry of F,,, we find gW -9w =0, (13.26) 106 XIII. GAUGE THEORIES REVISITED so (DEW — DoW). (13.27) Identity (5) leads to another restriction on W: (o4£Dj + o5j°D,)W* = 0. (13.28) aco Contracting with o°°’ and using (A.12), we have (56°F) + 5j°D,)W7 = 0. (13.29) Summing over & and 4 yields GW, =0. (13.30) An analogous result follows from (4): DW, = 0. (13.31) Identities (1), (2), and (3) do not lead to any new results. Identities (1), (2), and (3) are consequences of the other identities even without the constraints (13.21). To show this would require some tedious work which we shall omit here. Features like this are quite common in supersymmetric geometries. In general, part of the covariant curvature tensor may be expressed in terms of the other parts, and not all the Bianchi identities are independent. The technical reason for this stems from the fact that the derivatives of the basis forms E“ always contain a piece proportional to o,;“. We shall encounter this again (albeit in a much more complex form) in supergravity theories. To conclude this chapter, we shall summarize our solution to the Bianchi identities, subject to the constraints (13.21). We discovered that the Bianchi identities are satisfied by two superfields, W, and W*. These superfields obey the following constraint equations: GW, =0 IW, =0 (13.32) PW, — G,W* = 0. In the Abelian case, we recognize the conditions (6.8) and (6.12). These equations have (6.11) as their most general solution. In the non-Abelian XIII. GAUGE THEORIES REVISITED 107 case, it may be shown that Eqs. (13.32) have (7.22) as their most general solution. REFERENCES J. Wess, in Topics in Quantum Field Theory, J. A. de Azcarraga, ed., Salamanca (1977); Lecture Notes in Physics 77, New York, Springer- Verlag (1978). R. Grimm, M. Sohnius, and J. Wess, Nucl. Phys. B133, 275 (1978). EXERCISES (1) Show that (13.11) is the inverse of (13.10). (2) Compute e’, e%, and e, explicitly. Compare the result to (11.7). (3) Decompose F,,, into its spin-3 and spin-} parts: Fa, > 95" F ax = F pap 1 1 Pont = 5 Pani + 5 Pons 5 Fans = 5 Fd + Fai] (spin-3) 1 1 : 7 Feat = 5 [Foxit — Fepil — (spin-2). Show that Fag = — 3 Fens Fog if the spin-3 part of F,, vanishes. (4) Verify that (13.23) satisfies (13.22). (5) Derive the explicit form for 9.F,, and D,F 5. (6) Extract the Yang-Mills field v,, from the superfield W, in (6.11). Compare the result to (13.25). (7) Show that DW — DW = 0 implies Gave + Op0cq + OVay = 0 in the Abelian case. 108 XIII. GAUGE THEORIES REVISITED (8) Demonstrate, again in the Abelian case, that identity (1) is a con- sequence of the other identities and the constraints (13.21). Use i Fae = — GF" (DpFc + DAF jc) and oF, = —i¢l4D,0 OF ac = — 4 Fal (Dj OyP ac + Dz Op fc) i ~ es > = GF [Di(DF oe + OF ap) + D{DiF ne + OF p)].- (9) Verify that dy = —e”D,e~" is a solution to Fg = 0. (10) Compute the coefficient functions of the identity (12.29), PZQ = OF, in the e4 basis. XIV. VIELBEIN, TORSION, AND CURVATURE In previous lectures we considered theories invariant under rigid super- symmetry transformations. We now wish to gauge these transformations. In particular, we would like to construct theories invariant under x- dependent supersymmetry transformations. As in Chapter TV, such transformations induce motions in superspace: x" > x" + i(O0"E(x) — E(x)o"D) OY > 0" + EM(x) (14.1) 0, > 8, + E(x). These motions generate certain coordinate transformations: zM 5 7'M — 7M _ EM(z), (14.2) Thus it is natural to express our theories in the language of differential forms. This formalism is automatically covariant under coordinate transformations, as was shown in Chapter XII. Our basic dynamic variables shall be the vielbein and the connection. These superfields contain a large number of component fields. Some will be eliminated through covariant constraint conditions. Others will be gauged away with (14.2). In this way we shall arrive at a theory with the minimum number of component fields. The vielbein forms E4(z) define a local reference frame: EA = dE A(z). (14.3) They are manifestly coordinate independent. The vielbein fields E,,* are the coefficient functions of the vielbein forms. The vielbein fields change with the coordinates: a N OP Ey") = Bye) = 59 Ev". (14.4) 110 XIV. VIELBEIN, TORSION, AND CURVATURE In the infinitesimal case, this becomes 2M = M EM(z) Ey! = E’y(z) — Exy“(2) (14.5) = —C0,Eu* — (OyS)E,4 in accord with (12.17). Note that only the lower index M enters the above transformation. It is an Einstein index. Einstein indices take part in coordinate transformations. They will be denoted by letters from the middle of the alphabet. The upper index A is reserved for the structure group. We shall take the Lorentz group as our structure group. This is because we would like to recover supersymmetric flat space (13.11) as a solution to our dynamical theory. With this choice, the reference frame defined by the vielbein is locally Lorentz covariant: bE4 = EPL, A(2) - (146) Ey! = Ey®Lp4(2). In general, indices transforming under the structure group will be taken from the beginning of the alphabet. They will be called Lorentz indices. Note that the Lorentz generators L,* have three irreducible components: Lye Lg LP ;. (14.7) These components are related through the o-matrices, 6440 pj Lan = — yplag + 2esplaps (14.8) as may be seen from (A.13). The vielbein and its inverse EytE,’ = dy" uA” (14.9) EMEy® = 64° connect the two types of indices: Vay = Eny’V4 (14.10) Va = EaVy. XIV. VIELBEIN, TORSION, AND CURVATURE 11 Wherever possible, we shall write physical quantities in terms of Lorentz indices. They then have simple transformation properties. In addition, they may be fully decomposed into components irreducible under the Lorentz group. As an example, the vielbein forms E* = dz“E,y*, E* = dMEy*, and E, = dz“Ey, are coordinate-independent irreducible Lorentz tensors. To formulate covariant derivatives we must introduce a connection form &=dMby, bm = bua” (4.11) transforming as follows under the structure group: 5 = OL — Lo — aL. (14.12) The connection is the second dynamical variable in our theory. Note that gma? is Lie algebra valued in its two Lorentz indices: Oman = —(-)”buea- (14.13) Its third index M is an Einstein index. The covariant derivative of the vielbein is called torsion: TA = dE4 + E®p,4 =3 dz d2\Tyy4* 1 C PB A = 3F EPT gc*. (14.14) Explicitly, this becomes Tyo’ = OyEu* — (—)" Oy En* + (—)" ME Bobyyt — (—MEyB bun’ (14.15) The Lorentz tensor Tc is obtained from Tyy“ with the help of the inverse vielbein: Tyc4 = (—P" OE MER Tuy. (14.16) 112 XIV. VIELBEIN, TORSION, AND CURVATURE The sign factor may be derived from (14.14) and the definition of E4. It simply expresses the fact that the summation over M is carried through the index B. In flat space it is possible to transform the vielbein into the global reference frame (13.11): EA =e, (14.17) It is defined up to rigid Lorentz transformations. In this frame the connection vanishes: ¢=0. (14.18) The torsion, however, is non-zero: Tajo = Tjaé = 2ior gg’. (14.19) All other torsion components vanish. The curvature tensor is defined in terms of the connection: R= db + oo. (14.20) As usual, it is a Lie algebra valued two-form: B 1 M qoN B Ry = 342 dz’ Runa 1 'C ED B = 5 EP EPRoca = do™ dz’ Onda? + AM brea’ de® wee (14.21) From (14.21) we may read off the coefficient function Ryyy 42: Ryma® = Oxdua® — (=) Oudna + (HS rg yc? = ("by abuc?- (14.22) Since R is a two-form, we have Ruma? = —(—Y"Rawa” (14.23) XIV. VIELBEIN, TORSION, AND CURVATURE 113 and since it is Lie algebra valued, we find Rymav = — Rava Rymap = Rympa ab is Ryu? = Ryn! (14.24) Oxi O55” Ruma = —26apRwmag + 263) Rwmap- The last relation follows from (14.8). All other components of R vanish. The torsion and the curvature are the only covariant tensors which may be constructed from the vielbein and the connection. We must now find constraints in terms of these covariant quantities which reduce the number of component fields as much as possible. There are, unfortunately, no general recipes to indicate the proper constraints. Instead, one must examine the consequences of various choices. For example, it is impossible to set all torsion components to zero, for that would exclude super- symmetric flat space as a solution to our theory. Similarly, Eq. (14.19) allows only supersymmetric flat space as its solution. It turns out that T. ap = 0 Typ! = Tip’ = 0 Tj = Tyaf = 2io ng Ty = Tyg = 0 Tas = 0 (14.25) are the proper constraints. Here « denotes either « or @. In the next chapter we shall solve the Bianchi identities subject to these constraints. As with gauge theories, we will find that they con- siderably restrict the number of independent superfields. In fact, we will find that (14.25) yields the minimum number of independent component fields. These are the graviton, e,,“(x), the gravitino, Wn7(x), Wms(x), and the auxiliary fields, M(x) and b,(x) = b,*(x). These fields are not re- stricted by any differential equations in x-space. The spin-2 graviton couples to the energy-momentum tensor, while the spin-} gravitino couples to the spin-3 supercurrent. The auxiliary fields are just enough to equalize the number of bosonic and fermionic degrees of freedom off mass shell. REFERENCES V. P. Akulov, D. V. Volkov, and V. A. Soroka, JETP Lett. 22, 187 (1975). J. Wess and B. Zumino, Phys. Lett. 66B, 361 (1977). 114 XIV. VIELBEIN, TORSION, AND CURVATURE EQUATIONS Oi pf? Lay = —2aplap + 26apLap- 6 = pL — Lo — dL. TA = dE* + EPpy" 1 =5 dM d2% Tuy. Tym* = OyEy* — (—)"™ Oy En“ + (-)*ME, Fbyet — (—Y™E Pobre R= dp + o9. 1 R= 3 dzM dz§ Ryya® = dM dz’ dybya® + d2Mbya dX onc®. Ryma® = Onda” — (YOu bna® + (HD sb yc® 7 (- ym by yc. Tag! =0 Tig = Tif =0 Taji = Tjef = 2io Ty = Tyg = 0 Tw» = 0. EXERCISES (145) (14.6) (148) (14.12) (14.14) (14.15) (14.20) (14.21) (14.22) (14.25) (1) Compute dx™ and 6/éx"™ under the transformation (14.2). Show XIV. VIELBEIN, TORSION, AND CURVATURE 115 (2) Compute dQ, = Q'yn(z) — Qyy(z) for Q a two-form. (3) Show that ¢,,” transforms as a four-vector under (14.8). (4) Explicitly evaluate the covariant derivatives of the covariant and contravariant Lorentz vectors X 4 and X“: i DyX4 = byX4* + (—"X Oy 54 DyX4 = OyX4- bua’Xp DyX4 = Ex GyX4 DyX 4 = Eg DyX 4. (5) Use the covariant derivative of a Lorentz vector to define the co- variant derivative of an Einstein vector: VyXuy = (OME yABy?D pX 4 = OyXy + Uy Xp Py’ = (—MME yA GyE a"). (6) Show that V in Exercise 5 reduces to the usual symmetric connection in torsion-free ordinary space. (7) Decompose Ty¢4 into its Lorentz-irreducible tensors. (8) Linearize Tg-‘ about supersymmetric flat space: Ey4 = ey’ + Key®Hs’. (9) Assume that E,,“ = 6,,7 + *+* has mass dimension zero. Give the dimensions of E,, E,,7, and E,*, as well as Ta,°, Tag’, Tas’, and Rage’, R 6. aby * kJ=-1 0 [ol=-5 [Fl=0 [B"}= -5 [Ewl=5 [&"]=0 116 XIV. VIELBEIN, TORSION, AND CURVATURE [Twl=t [Tul = 5 [T’] = [Rasc!] = No NI [Ran°] = 2 XV. BIANCHI IDENTITIES We are now ready to solve the supergravity Bianchi identities subject to the constraints (14.25). We will proceed in analogy to Chapter XIII, where we solved the Bianchi identities for supersymmetric gauge theories. We will find that the Bianchi identities reduce the number of independent superfields contained in Ey“(z) and @y,4°(z) to one complex chiral super- field R, one hermitian vector superfield G,;, and one chiral superfield W,g,, totally symmetric in its indices. The torsion and the curvature may both be expressed in terms of these three superfields. We shall summarize our results at the end of this lecture. These formulae will be used frequently in the coming chapters. It is not necessary, however, to work through the details presented here to understand the rest of the book. We begin by stating the Bianchi identities for the torsion and curvature. These follow directly from (12.29) and (14.14): DDE" = EPR,A DT4 = EPR,4. 5.) Here R,4 denotes the superspace curvature and T4 the torsion. We wish to break this equation into its Lorentz-irreducible components, so we compute YT“ in the basis defined by the vielbein forms: 1 - OT = 5 HEE T *) 1 BRC, A 1 Bre 1 BRC A (52) = 5 PPEDT eg" + 5 ETT ew — 5 TPE Ten. Substituting this in (15.1), we find EPECE Dy Tex" — Roce’ + Tye Tre) = 0. (15.3) This identity contains thirty Lorentz-covariant components. Some of them, however, are related by complex conjugation, and others are automatically satisfied because of the constraints (14.25). In all, there 118 XV. BIANCHI IDENTITIES are thirteen independent components: (1) Resye + Roypa + Rypsa = 9 (2) Reign + Ripia = —2i65;!T py — io gs! Toya (3) Rosia = — its! Tpsa — 2io gs! Taya (4) Rovyx + Rybia = —D,Tirx — YoT you (5) Raja = DoT jn + DjTovs + ios; T yrs (©) DsT inn + DiT jon = 9 (1) Rééca = — 2G agsT pe? — Lio agp Tic? (8) Rgica = —2itagsT pe? + iG agp Tic? (9) Roasi = DoTaja + BaT na + Dj Tan + Tas? T oo + Tye? T gai (10) DyT aja + BaT jn + DT yaa + Ta5?T ona + Tip? T gaa =0 (11) Rgaca + Repaa + 2idagpTa® = 0 (12) DT aca + BaT con + DeT ran + Toa®T gen + Tac®T gra + Tet? Ton = 0 (13) Roaca + Recra + Revda = 0- (15.4) The underlined index ¢ is summed over both ¢ and ¢. We shall first solve the identities which are linear and without deriva- tives. These are Egs. (1), (2), (3), (7), (8), (LL), and (13). We start by con- verting (7) Réica = — iG gsT je? — iG agp Tic? (15.5) to spinor notation: Rhdyjat = Fy5Cxi Riiica © 15.6) Tiyig = %y5° Tig (56) Since R is Lie algebra valued, we have Rginiaa = — 28 aRpaja + 28jaR pire» (15.7) where Rjs,, and Rgjjq are symmetric in ya and yo respectively. Since R is a two-form, Rj, and Rgj;, are also symmetric in Bd. With these ex- pressions for R and T, Eq. (15.5) becomes bi aRpinn — SyaR joie = Esp T iyin + C45 T fyja)- (15.8) XV. BIANCHI IDENTITIES 119 The tensor T;,;, may be decomposed into components with definite symmetry properties: Tiyin = P4j8yaT + 04jT ye + PT ig + Trad: (15.9) In (15.9) and in what follows, tensors are symmetric with respect to underlined indices. Equation (15.8) now splits into several symmetry classes. We first consider the part which is antisymmetric in both pa and ja. This may be projected out with e” and e*: Tifa + Tin = 0. (15.10) Note that the curvature R drops out of this expression. Substituting (15.9) into (15.10), we discover Ts = 0. (15.11) We next consider the part of (15.8) which is symmetric in ya: b5aR pine = i(Cspls; + Fx5eHj)T ya t 2i(65jT yx53 + €45T yx) (15.12) If we multiply this by e, we obtain Ring = Si8. 2 ~ OT yi. (15.13) dye However, R;j,. is symmetric in 70, so Ty, = 0 (15.14) and Ring = — GIT 9 55- (15.15) Tf we multiply (15.12) by <*’, we find Rijinn = —2iT ya fi- (15.16) Equations (15.15) and (15.16) are consistent if and only if Taps = 0. 15.17, Rix = 0. ( ) 120 XV. BIANCHI IDENTITIES Only one term remains in the decomposition (15.9); we call it R, where T= —2iR: Tinig = —2itjl,aR- (15.18) yO ye From (15.8), we see immediately that Ryaia = Mepis; + bssepi)R. (15.19) With these results, we have found the most general solution to identity (7). We have also learned that Rgj,, and T;.4 may be expressed in terms of a single superfield R. Similarly, we may write Rys.q and T,., in terms of R*. The above expressions satisfy identities (1) and (3) as well. We now consider identities (2) and (8). The computation is quite similar to what we have done, so we merely list the results: Reina = €,pGas + FxpG5 ~ ep, Ggy — 3epoGyy) (15.20) We leave the details of this calculation to the reader as Exercise 1. Identity (11) gives Rgq.q in terms of the torsion. Since Rgg.q is anti- symmetric with respect to c and a, we find Ryde = iG apsTea® ~ FopiT acd ~ Fp Tad) (15.21) We have now solved all the derivative-free linear identities except (13). Identity (13), however, is just the usual cyclic identity on the curvature in four-dimensional space. It is familiar from ordinary gravity theory and its consequences are well known. In spinor notation, the symmetry properties of eos. ce db, Ry sip hai = Fy} O38 Op Oxi Reaa (15.22) lead to the following decomposition: Ry ssippae = 46. 56 paX joa — 4666 aF 35 pe — 4€;8 pa? yo ps + 4855 p2Xyop0- (15.23) XV. BIANCHI IDENTITIES 121 Identity (13) is satisfied if and only if Frais = Pains Xapy! = bays (15.24) where A is real. We shall now proceed to solve the identities which contain derivatives but remain linear. These are identities (4), (5), and (6). We begin by insert- ing (15.18) into identity (6). This yields £5j6pxD sR + £;p6p,D5R = 0. (15.25) Contracting with 28, we find DR = 0, (15.26) so the superfields R and R* G,R*=0 (15.27) Y are chiral. Evaluating identity (4) is tedious. We must make use of the fact that Rica is Lie algebra valued: Ro yiaip = Fj Fa" O pf Roca —2egpRiniap + 264pRoysap- (15.28) The component Rj, is related to the torsion through (15.21), where Tyg has the following decomposition: Toiyié = —2eay(Wijg + £53W5 + 2gsW3) + 2e5;Wo,g- (15.29) Combining (15.21), (15.28), (15.29), and identity (4), we find Tsiysa = — 285, Wie 1 ~ 5 ayl0ayD°Gps + 235D°Gy) 1 + 581 PG ye + DG) (15.30) 122 XV. BIANCHI IDENTITIES and 1 Rosina = 5 ie gsD, + bpy,Ps)Gas 1 i G Gg + 5 Hep. + n.DIG,j + iExpOs, + yp 5.)D°G,5 g . 1. Rysiga = ie gs Wyas + 3 il Goi + &54P pGo;)- (15.31) The symmetric tensor W,,, drops out of the identity, so it remains un- determined. The tensors W; and W,,4 are related to G. Identity (5) gives another relation between the same curvature and torsion components. It yields DC, = GR* (15.32) as a consequence of (15.30) and (15.31). We have now solved all the linear identities. The nonlinear identities either define components of the curvature and torsion as nonlinear ex- pressions in G and R or they may be reduced to linear equations through the commutation relations of the covariant derivatives. For example, identity (9) expresses R,,;, and therefore X,;,, and V,,,, in terms of torsion components. These, in turn, may be expressed in terms of W, G, and R: > 1 _ z Xpin=] [- Dp Tj! se—DpsT p+ DT pi sa Ts Toppa Tit Tes (15.33) Because of the symmetry properties of Xa ja and because of the relation DPC jg + DjiDG; (15.34) we find = 1 P Wags + ZO p's + DysG*j) = 0 (15.35) from (15.33). Finally, we examine identity (10). The torsion terms may be expressed in terms of W, G, and R. All but one contain an ¢-tensor, so symmetrization XV. BIANCHI IDENTITIES 123 in all indices yields DWyi5 = 0. (15.36) The symmetric tensor Wj,; is a chiral superfield. We have now solved the Bianchi identities (15.4), subject to the con- straints (14.25). We have learned that all the components of the torsion and the curvature may be expressed in terms of the superfields R, G,z, and W,,,. These superfields are subject to the following conditions: (1) ,R =0 2) Gg =DjR* D'G,;=F,R (3) DzWps = 0 D,Wp;5 = 0 wm, +! (4) PWgs +5 i(DpjpG"5 + DysG";) = 0 1 _ . DWeys + 5 i(DggGy + DgG,?) = 0 (5) (Gaa)” = Goa (6) (Wag,)* = Wags- (15.37) The superfield W,,, is completely symmetric in its indices. For future reference, we collect our results below. Torsion: (1) Ty = Ta? = 2io,¢ Q) Tat = Tal = RPT Ticig = —2i€5;6,,R (3) Tse® = —Tes' = -} ET et Tess = — 2ies8igR* 1_, (4) Tse! = — Tost = 5 8 T Tix = G CaGas — 3en.Gea — 3er.Gas) 124 XV. BIANCHI IDENTITIES i 1 (8) Tid = Tes? = — 5 FT i Tisée = g (aCe — 365sGee — 385:Gea) (6) Tac’ = — Tea” = GEHT,. 7 GO." T 55,5 Ale Tinian = — 265; Woy 1 = = - 5 %6ilC xD GG,” + £,D G59) 1 oo = + 5 %(BiG.j + GjG,5) ae a_i gbiginp,. 4 (7) Tac’ = — Tea’ = 7 54°F" T55y5 AL Tinie = — 285, Weise 1 = 5 PaileieDpG*; + &54DoG"s) 1 + 5 8i(PsG + FG) (15.38) Curvature: (1) Royex = Feat: Rijia = Meee n + G85) R (2) Rois = Rijex = 0 + bE,)R* (3) Roier = Ryoer = —(EseGaj + €4xG) Roiix = Riots = —(€Gox + 852Goe) 1_, (A) Recda = ~ Reais = 7 Fe Revi i Reyisa = He ca8 yx + bexbys)D GG" + 3 (iyBs + &esP Guy i + 3 (eyFe + bg,D G55 XV. BIANCHI IDENTITIES 125 1, (5) Recéis = —Recés = 7 FR er ée oy I Reyiba = Mlb, Wiis + 5 (CPGya + bePG,s) low (6) Ricsa = —Reéon = 3 Fe Revie . i = - Riyjon = Fibs; Woya + 3 €¢P Ga + b,D:G5;) 1_ (7) Resa = — Reese = 35 G Revie Ranji = Uess8ja + F:xt53)D 5G," i = i = = + 5 iis + 3D NG + 3 (ij Bs + &4Ds)G,5 1_. es (8) Rene = 5 Fea” Reisiya 1 sez io Reagi = 4 GG R sasiia Ressija = — WeoXssyn + 2s V cov Resoiya = 2:5 X copa — Essie 1 Xica = 4 (Pi Woce + BsWeay + DWays + G,W,s0) ; +4 pb + (yates + €5afe,)) —2RR* + 3 Gop 1 + 7g (FDR + 29,0} Peas = ¥. jdex, 1 = [(GaiGs; + GaGa) +g PuiGes + DGus + DiGey + PiGes) 1. - - + G(FiIGes + FjDG5 + DsP.Gu + FxPGu). (15:39) All other components vanish. 126 XV. BIANCHI IDENTITIES It is quite remarkable that these relations also solve the Bianchi iden- tities arising from the curvature (12.31): PR=0 ECE-E"[DpRyca® + Teo’ Reca®] = 0 (1540) [9eRoca® + Teo" Reca”] : REFERENCES R. Grimm, J. Wess, and B. Zumino, Nucl. Phys. B152, 255 (1979). N. Dragon, Z. Phys. C2, 29 (1979). EXERCISES (1) Show that (15.20) is the solution to identities (2) and (8). (2) Derive the conditions (15.24) from identity (13). (3) Show that identity (5) implies (15.32). (4) Verify that (15.38) and (15.39) satisfy (15.40). XVI. SUPERGAUGE TRANSFORMATIONS In the past few chapters we have considered the general coordinate transformations of superspace Mm M4 EM(z), (16.1) We have also introduced a structure group and explored its transforma- tion laws (12.18). In this chapter we shall define supergauge transforma- tions. Supergauge transformations are constructed from the general coordinate and structure group transformations of superspace. They amount to a convenient reparametrization of these transformations. Supergauge transformations map Lorentz tensors into Lorentz tensors and reduce to supersymmetry transformations in the limit of flat space. The parameter € characterizes infinitesimal changes in coordinates. It may be written with either an Einstein or a Lorentz index: &4 = EME, A. (16.2) Note that either €4 or €M may be chosen as the field-independent trans- formation parameter. Its companion then depends on the fields through the vielbein. Since we would like Lorentz tensors to transform into Lorentz tensors, we shall choose €4 to be field-independent. We must now write the transformation properties of tensor superfields 6V4 = —EMAyV4 + VPLS (16.3) in terms of the parameter &4. In (16.3), V4 represents a general tensor field, and the representation L of the Lorentz group corresponds to the tensor structure of V. For scalar fields, we have 5V = —EMOyV = —E4EWMOyV = —U4D,V, (16.4) while for tensor fields, we find V4 = ~EFEgM yV4 + VPLy". (16.5) 128 XVI. SUPERGAUGE TRANSFORMATIONS As it stands, Eq. (16.5) is not covariant under Lorentz transformations. The derivative in (16.5) must be replaced by a covariant derivative: DygV4 = OyV4 + (PVE Dae! ; ; (16.6) DyV4* = EgOyV4. Substituting (16.6) into (16.5), we obtain V4 = —BD VA + VEE dept + VOLS. (16.7) The connection ¢¢," is Lie algebra valued, so €“#c,* acts like a field- dependent Lorentz transformation on V*. If we set Ly = —E des", (16.8) we find Il 6V4 = -EDV4A (16.9) for any tensor superfield V4. Equation (16.9) is manifestly covariant under Lorentz transformations. The condition (16.8) defines supergauge transformations. Supergauge transformations consist of a general coordinate transformation with field-independent parameter 4 followed by a structure group Lorentz transformation with field-dependent parameter Lg* = —€°¢¢,*. It is among this restricted class of transformations that we shall find the gauged supersymmetry transformations. Let us now compute the commutator of two supergauge transforma- tions. Since £4 is field-independent, we have 5,6V4 = —€°5,9oV4 = “PD yDV4, (16.10) so (6,5: — 5:5,)V4 = END Be — (—MPDpV4.— (16.11) This expression is easily evaluated with the help of the Bianchi identities (12.29): DOV4 = VR». (16.12) Here Rg“ is the Lie algebra valued curvature two-form and V4 is a tensor XVI. SUPERGAUGE TRANSFORMATIONS 129 zero-form. As in (15.2), we write DOV* = GE" D,V4) = E®(D5V*) + (GED,V4 = EXECDDgV4 + T?D V4. (16.13) Substituting (16.12), we find EPECDDgV4 = VR, — T?DpyV4, (16.14) or, for the coefficient functions, (DeDpy — (—PDyD V4 = (—Y"VPRegy’ — Tep”DpV4. (16.15) This tells us that the commutator (16.11) closes into a field-dependent Lorentz transformation and a field-dependent transformation of the type (16.9): (5,5: — 5,5,)V4 = VPENPRecv’ — EnPT xe? DpV4. (16.16) In flat superspace, where the curvature vanishes and the torsion is pro- portional to the o-matrices, Eq. (16.16) reduces to a familiar form: (5, 5¢ — 6:5,)V4 = —2i(qo"E — Eo"7) 0,04. (16.17) The 0 = 8 = 0 components of € and give the commutator of two supersymmetry transformations (3.4), so (16.9) indeed includes gauged supersymmetry transformations. We conclude this chapter by computing the changes in the vielbein and the connection under supergauge transformations. In general, the trans- formation properties of the vielbein are given by (14.5) and (14.6): OEy* = ~ ECL En! = OyS" Ey! + Ey? Ly* = EM OLEM* — (—S" OuEr*) — O44 + Eu? Ly4 = —dub4 — (Tia — bum’ + (—)™ ban’) + Ex? Lp". (16.18) Here we have used the definition of the torsion (14.15). The connection mx‘ combines with 6,64 to make a covariant derivative. Substituting 130 XVI. SUPERGAUGE TRANSFORMATIONS the special Lorentz transformation (16.8), we find the following super- gauge transformation law: bEyt = —Gyk4 — EFT gy4. (16.19) We proceed similarly for the connection: Sbya® = —o" Orbea” — (One )bLa? + dual? _ (—n@ OL bye = Oy La®. (16.20) For a supergauge transformation, this becomes Seba” = —ERema?- (16.21) The proof of this relation is left to the reader as Exercise 3. The transformation laws (16.9), (16.19), and (16.21) allow us to compute the transformation properties of all the independent supergravity com- ponent fields. This we shall do in the following lectures. REFERENCES J. Wess and B. Zumino, Phys. Lett. 79B, 394 (1978). J. Wess, in Quantum Flavordynamics, Quantum Chromodynamics, and Unified Theories, K. T. Mahanthappa and J. Randa, eds., New York, Plenum (1980). EQUATIONS 5V4 = —EBEM OyV4 + VELS4. (16.5) 5V4 = —E8D V4 + VEE egt + VELg4. (16.7) 6:4 = —EDVA. (16.9) (DBy — (-"DpBNW4 = (—)!" VW Rego — Tex? DyV4. (16.15) (5, 5¢ — 625, V4 = VPENPRecp* — ENT gc?DpV4. (16.16) bgEy" = —Dyb4 — PT gy. (16.19) Sbua” = —€Rewa®. (16.21) XVI. SUPERGAUGE TRANSFORMATIONS 131 EXERCISES (1) Show that (16.15) may be written (QeDp—(-/DyDIV4 = —(—)*Reg*yV? — Ten? DyV* for contravariant vectors V4. (2) Use the definition of the covariant derivative of a covariant vector DyVa = OnVa — bmaVe to derive the analog of Exercise 1 for covariant vectors V;. (3) Prove (16.21) using (14.22). XVII. THE @ = 6 = 0 COMPONENTS OF THE VIELBEIN, CONNECTION, TORSION, AND CURVATURE In Chapter XIV we defined the torsion and the curvature in terms of the vielbein and the connection, the dynamical variables of supergravity. By construction, all are superfields, whose expansion coefficients are x- dependent component fields. In this chapter we will see that the compo- nents of the torsion and the curvature can be expressed in terms of the lowest components of R, G, and the vielbein. The same holds true for the vielbein and the connection. This implies that the lowest components of R, G, and E are the physical supergravity degrees of freedom. The re- maining degrees of freedom are pure gauge, and can be transformed away. The transformation parameters £4 and L,, are functions of superspace. Their lowest components characterize general coordinate transformations in four-dimensional x-space [é4(x)], gauged supersymmetry transforma- tions [€(x),£,(x)], and local Lorentz transformations [L,,(x)]. We will use their higher components to transform away certain 0 = 9 = 0 com- ponents of the vielbein and the connection. We first consider the vielbein. Its transformation law (16.18) may be written as a supergauge transformation (16.19) together with an addi- tional Lorentz transformation L,4: OEy*t = —Dyt4 — EBT gy’ + EyPLy'. (17.1) The lowest component of this equation gives the transformation property of Ey“|e-7=0- Higher components of ¢4 enter 6Ey“| through the co- variant derivatives 9,¢4 and J*é4, We may use these higher components to transform Ey,4| to the following form (see Exercise 1): eal) 5 Ua) 5 Dna) Ewialoao=| 9 5 gf TD in 0 0 oF, The fields e,,", V7, and W,,; cannot be gauged away. They describe the spin-2 graviton and the spin-} gravitino, The inverse vielbein E,™| has a XVII. 6 = 0 COMPONENTS 133 similar structure, 1 1_ ea) 5 Vax) 5 Vai) Ex" (2)|o-0=0 = 0 a » ff as 0 0 5, with ese? = 6,° Val = eg Wy?5y! (17.4) Vai = Ca Di i» We now consider the connection. Its transformation law (16.20) may also be written as a combined supergauge and Lorentz transformation: é B_ _¢CR By crs dua MA outa b c . B (17.5) = (HL buc® — Ola? We may use the higher components of L,* to transform away ,,4°| and $" ,®|. This is possible because 4” is Lie algebra valued: Oma (Dlo=0= bya lo=a=0 = b44%(2)|o-a-0 = 0. No further components of E| and ¢| may be gauged away. In ordinary relativity it is possible to express the connection in terms of the vierbein. This follows from the fact that the torsion is constrained to vanish (see Exercise 2). In supergravity we also have constraints on the torsion (14.25). These constraints allow us to express the connection in terms of e and w. To proceed systematically, we start from Eq. (14.15). The T,,,4 com- ponents of this equation contain no 0 or 0 derivatives. They relate T,,,4 to the lowest components of E and ¢: = Oma") (17.6) a ‘mn Tm’ | = x2 — Omen + On — © T, | 1 1 5 Cala? — Ou’) + 5 Pap” ~ Vib ong) =!Gy2-Gyy—ty4 = SB — Dy’) = 5 Von) 1 Tamil = 3 Pm) (17.7) 134 XVIL. 0 = 0 COMPONENTS Here we have used the following definitions: Om" = Cm’ Onn! Didnt = On? + VP Ong (17.8) Want = Dan? — Daa A To apply the constraints, we must relate Tyy“ to Tcg* through the vielbein Tya’ = EyP Ey Tcx(— yr". (17.9) Taking the @ = 0 = 0 component of (17.9) and applying the constraints, we find Tam’ | = Ey PE ys Tg’) + EmpEn TP -5 nD — Vu Vn) (17.10) and E,PEfT | + En’ En'T "| + E, PET 43] + Eq? Es T's'| + EmpE, TE. (17.11) | mm Combining (17.10) and (17.7) gives the connection in terms of e, , and 1 i _ i - _ Ome = 5 \-3 Can" — Von) — 5 male — Wey) + 5 ale — Vu) ~ eel Oaent ~ One ema(C cent — Gy) + enn = arnt (17.12) We must now evaluate Eq. (17.11). The torsion components in this equation were computed in Chapter XV. Equation (15.38.6) relates T,,” to W,,, and G,Gp;. Equations (15.38.2) and (15.38.4) relate T;,” and T,, to Rand G,j. We may use these expressions, along with (17.7) and (17.11), to compute W,,,] and Z,G,,| in terms of e,", Wn", Uni, and the lowest com- ponents of R and G,y. This is done in Exercises 5 and 7. XVII. @ = 8 = 0 COMPONENTS 135 The lowest components of R and G,; cannot be expressed in terms of Cn’. Wm, and Winge Nor may they be gauged away: bR = —EDR / (17.13) 6G, = —€ PG, + GL, This forces us to introduce two new component fields: 1 R(2)\o-0-0 = ~5 Me) ' (17.14) Gs2o-n-0 = —5 Pal These fields equalize the number of bosonic and fermionic degrees of freedom within the supergravity multiplet M, b, y, and e. We shall see that the supergravity multiplet forms a complete set of dynamical fields. To conclude this chapter, we follow the same procedure with Ruma®, the only tensor we have not yet discussed. Taking the 0 = 0 = 0 com- ponent of (14.22), we find b_o b cw? cw? O,Oma — Om@ng + Oma One — Ona’ Ome , (17.15) This equation defines the Riemann curvature yma” in terms of the connection ,,,°. In analogy with (17.11), we relate Rymq” to the Lorentz- covariant tensor Rep,”: Ruma’ = EyEm? Reva)“ = Ey Em'Reag’ + En?Em'Ryaa” + E,SEn? Resa’ — (17.16) The underlined spinor indices are summed over dotted and undotted indices. Comparing with the solutions to the Bianchi identities (15.39.8), we see that R,4,° is related to the second derivatives of R and G and the first derivative of W. Similarly, R,,.° and R,s.? are related to R, G, W, and the first derivative of G. Combining (17.15) and (17.16) allows us to solve for the second derivatives of R and G and the first derivative of W in terms of the supergravity multiplet M, b, y, and e. 136 XVII. 6 = 8 = 0 COMPONENTS All we have left to compute are the first derivative of R and the second derivative of W. The first derivative of R is related to the first derivative of G through the Bianchi identities. It is computed in Exercise 8. The second derivative of W is outlined in Exercise 10. With the results of this chapter, we have what we need to compute the torsion and the curvature. The first step is to find the components of R, G, and W in terms of the supergravity multiplet. From this, we can then derive the torsion and curvature through the solutions to the Bianchi identities. Those components of R, G, and W that we will need are col- lected below. REFERENCES B. Zumino, in Recent Developments in Gravitation, M. Levy and S. Deser, eds. (Cargése 1978), New York, Plenum (1979). K. S. Stelle and P. C. West, Phys. Lett. 74B, 330 (1978). EQUATIONS OE’ = —Dy&* — ST gy* + Ex? Lp’. (17.1) 1 em") 5 Ym(X) > mil) 2 Ey“(2)\o-a-0 = . . (17.2) 0 62 0 0 0 5, mn 1 1_ eax) 3 Wal(x) > Vail) E4M(D\o-0=0 = 0 oe 0 . (173) 0 0 5 Soma = —ERema® + bua Lc? = (HL Mbuc® = Oy Ly? (17.5) = Oma (X) , a (1716) = PF 2)|o-a-0 = 0. Dim? = Ont + Wy Png? Onn (178) Van? = De — Dy 2 XVII. @ = 6 = 0 COMPONENTS 137 1 i a: ay, i ay, ay, ame = 5) ~3 Leal nV n — Vn Wn) = 5 mal noe = WoW) i + - + 5 eval em — Wm" We) — CcalOnem’ — Omen) = CmalOcen’ — One") + Cna(Omee” — erat (17.12) 6R = —EDR (17.13) 5G, = —£BG, + GL. 1 RG)loav-0 = —g MCs) \ (17.14) G,(2)|o=0=0 = “3 ba(x). GpOma? ~ On Ona? + Oma One! — Oya Ome = Rina» (17.15) 1 . , Weal = sg, LY Wai’ + thaybs’)- 6) 2-4! pissy QD 1 7 1 vy i * Gaal = 4 Val sia + D 85 5 — é WaisM + 5 WasPbas + Daphbar — Tshabas BG. j-ty ia + 1 ssh aig M baa 4 yam 12 ‘bahay 6 od: i — [gpl bas + Wps?baz — WPiabps)- ”) 1 ia i D,R| = 5 (0 bang + 5 (OU JM — 2 aad” (8) m4 1 it i j i BR| = — 50 Dal + ZV IM* + ZTE". 138 XVII. 6 = 6 = 0 COMPONENTS EXERCISES (1) Show that E,,4 may be gauged into the form (17.2). Use the freedom available in the higher components of €: A= E (0-04 (0) + OME, 1-4(x) + BHO DAL) foes (2) In ordinary four-dimensional relativity, the torsion takes the form Tam! = Ox€m! — Om@n’ + Dum? — Omn* Impose the constraint T,,,," = 0 and solve for m in terms of e. (3) Solve (17.7) and (17.10) for @pne. Use the fact that Omne = Cran Ome’ = — Omens (4) Use the definitions Wain = F5;°Ca" Winy . = go. £e,le™ Woayia = G55 Ty5 Ca Co V ama along with (17.7) and (17.11) to verify 1 1 Toiyiel = 5 Waivia ~ 5 Wai? Tovial — Wri? Train) Liye 6 + 3 os Tyial — V5? T posal)» (5) Show that Exercise 4 and the solutions to the Bianchi identities give 1 i Woyal = ya on (isiy?a + Wojyby’). XVII. 6 = 0 COMPONENTS 139 (6) Use the solutions to the Bianchi identities to show that 1 9G. =—e®T., AediT DsGui OPT aaj + 6 b5g8 ET p54, 2 1 ; digs via ~ B58, jad = 1 FiGus = 5 0°T (7) Compute Y,G,,| and G5G,,| in terms of the supergravity multiplet: 1_, 1 Ty i DsGuil = rae + DR b5V a5 — 6 vaaoM* + 55 Vadis t Vop?bax — Vs abap) too is DGaal = ZW iyi + Jp Cie in + G Vass i — yy oat as + Vips! Bas — W'sadya)- (8) Use (15.37.2) and the results of Exercise 7 to compute Y,R| and G*R*| in terms of the supergravity multiplet: i 1 ; FAR] = 3 Masp + GUM — 5 Wend™ _ 1 i i. PR*| = 30 hal + (USMY + EDD". (9) Denote F by Rand write the Bianchi identities (12.31) in the following form: ECEPEND pRyca” + Trp’ Rrca’} = 0- Show that this implies D Rex? + DaRea? + DRiar! + Toa’ Rees! + Tac’ Reig? + Tei" Rea’ = 9. (10) Use the solutions of the Bianchi identities and the result of the previous exercise to show that ZW may be computed. Warning: the actual calculation is tedious! XVIII. THE SUPERGRAVITY MULTIPLET We are now ready to derive the transformation law of the supergravity multiplet @,,7, Wn2, Wis 0a and M. We start with the general transforma- tion law of the vielbein SEy’t = —Dyb4 — ET yy + EyPLy4 (18.1) and evaluate its lowest component. The 6 = 6 = 0 components of & and &, parametrize gauged supersymmetry transformations. We shall focus on these by setting (2)|o-3-0 = 0 &(\o-a-0 = Cx) Elo-a-0 = Fal) Lan(2)\o=0=0 =0. (18.2) Higher components of &4 and L,, will be chosen to preserve the gauge (17.2) and (17.6). To preserve (17.2) we must require 5E,4| = 6E*4| = 0. (18.3) From (18.1) and the constraints (14.25), we find é = std = (~~ Pot — yriy + Be) a ‘a agit + ios? = 0 6 a —w, § E*| + 2ifPop,%e** = 0. (18.4) XVIII]. THE SUPERGRAVITY MULTIPLET 141 Equations (18.4) are satisfied if &* = 2i(00°C — CoO). (18.5) No further conditions follow from (18.3). To preserve (17.6) we must demand 5, 4°| = 5h" 4"| = 0. (18.6) From (16.20), we have Sys] = —EReya?| = OyLs"|- (18.7) The curvature term does not vanish. From the solutions (15.39) of the Bianchi identities, we know that it contains M and b: Ruse] = AC jab up + Suatip)R |, 2 * = 3 (Efup + buat pM Riuxpl = ~(ExeG pj + FupGus)| 1 = 5 CaPpy + Sysba) (188) Substituting (18.8) into (18.7), and imposing (18.6), we find 2 3 Cote + Cpt aM 1 a + 3 Cube + Eygb,3)0". (18.9) This tells us that a gauged supersymmetry transformation ¢, must be accompanied by a field-dependent Lorentz transformation 1 ; Lip = 3 [0,(26pM* — bys?) + O,(2¢,M* — b,sb*)] ' (18.10) Lap = 3 L025 5M — Cbyp) + O,(2E,M — 67,4) to preserve the gauge (17.6). Equations (18.5) and (18.10) are the only conditions that follow from the gauge fixing. 142 XVIIL THE SUPERGRAVITY MULTIPLET To summarize, we have found a set of transformations, parametrized by ¢, which include gauged supersymmetry transformations and pre- serve the gauge (17.2), (17.6). We shall call these the supergravity transformations: HQ = 00) El) = Les S%(2) = 2i[Bo%(x) — x)o°O] Lage) = 5 (0,[2% 0M) ~ bys OE] + 6pL25,0)MM(x) — bu sCORN]} Lalo) = 5 Os jooMCx) ~ CO 40] + DiL2ZoIMex) — O60} Ly = 500 Lag - 5 (00,24)? Lap (18.11) We are now ready to compute the transformation laws of the compo- nent fields. We start with the vierbein. From (18.1) and (18.11), we find Seg! = SE gt| I -Iyqk"| — ET ya = PT gp'| — SpT?n'|- (18.12) The terms proportional to €* do not contribute for 0 = 0 = 0. The torsion terms may be evaluated with the help of the constraints: Tom’ = Eno T pc — J"? = Ep iog;" . (18.13) Tin’ = Exe T(r"? = —2iE,,’6,4". Their 0 = 0 = O components are specified through (17.2): Tpm'| = idgg” mb aml = Foes" (18.14) Tim'| = Wp! Opp Inserting (18.14) into (18.12) gives the transformation law of the vierbein: 5en'(x) = mo — Con)» (18.15) XVIII. THE SUPERGRAVITY MULTIPLET 143 We now turn to the gravitino: Uy? = 5Ey!| = —DyL*| — CT an’ (18.16) Nie As before, we write Tint = En€Tjc%( — J"? Tmt = Em oT gc)". (18.17) The solutions (15.38.2) and (15.38.4) of the Bianchi identities give T,,* and T;,,* in terms of R and G,: T! 2 = —ig?*.R i (18.18) Tyet = gP°ASEGpe — 35y'Gy, + Bey. - Restricting to 9 = 6 = 0, we find | i a ).% Te = j (c6,.)4°M (18.19) ifi Tye | = 3 {i (G46 Jp" + ayn bt. Substituting (18.19) in (18.16) gives the transformation law for the gravitino: i OW? = —2Dyk™ + 3 en '(€0 go °M 1 (18.20) + ie,PC? (an + (oi) be. A similar calculation holds for : > i OV mg = —2D yg ~ 3 em Cap * (18.21) > 1 ~ tenth (sts +3 (0) De 144 XVII. THE SUPERGRAVITY MULTIPLET The transformation laws for M and b, follow from (16.9): 1 — GoM = 5R| = —2F,R| 1 (18.22) —3 dba = 8G, = (GD, — GG). The term proportional to €, in 6M drops out because R is chiral. The Lorentz transformation (18.10) does not contribute to 5G, for 0 = 6 = 0. In Chapter XVII, Exercises 7 and 8, we computed Y,R| and 9,G| in terms of the supergravity multiplet. From here it is only a short calculation to find 6M and 6b,. In conclusion, we collect our results for future reference: Bet = Wao Co") Bat = —20g6" + eg | Meo + bar + $ACor Bhs = —20qhs = ieg |S MMO, + Ble — $H.0ey} 5M = —C(o°S", + ib*W, — io*V,M) 3 1 os i i . Sbea of Va! sia + a eV a3 — 3 M*Wais + Vad? bos . 3543 1 5 + V5j Das _ Taradh -? \; W spice + 4 bia yy i i +5 Mais — 4 Voi? bas + Wi? bax — Wubyah (18.23) REFERENCES K. S. Stelle and P. C. West, Phys. Lett. 77B, 376 (1978). S. Ferrara, F. Gliozzi, J. Scherk, and P. van Nieuwenhuizen, Nucl. Phys. B117, 333 (1976). EQUATIONS 2) = C(x) Ez) = Gx) &%(z) = 2i[00°C(x) — ¢(x)o°O] XVII. THE SUPERGRAVITY MULTIPLET 145 Lgl 2) = 5 (OAL 0) MC) — By C0] + OpL2oM x) — baslxK0)}} Lage) = 5 (OL ALIMC) ~ OC, 9] + DL 2ECOM(H) — Ob a(0]} 1 Liv = 5 1 5 FaOre) Leg — 5 (60,05) Lap. (18.11) Sem = iW — CoD) Wyk = —2D yl + in {5 Meo ff + bf + zocor 1 8 mg = —2Dubq — iby’ {i M*(Co.)q + bbe ~ ; Waa.) 6M = —{(0°S" a, + ib, — io,M) < 3 1 | i is 5b, = 0 \j Wa! aie + re) - 3 Maia + Yaa a mm ton nor choos + Wai baz — Varadast — F 4 W sya + 4 cela “yy i, - i +5 MY a5 ~ a (Wpa?bas + Wi? dxa — Wabyah (18.23) EXERCISES (1) Compute ée,," for €* = &, = Lyg = 0, Ez) = &(x). Compare this with a general coordinate transformation and a local Lorentz rota- tion in ordinary relativity. (2) Compute dy,,, by conjugating dy, in (18.20). (3) Show that the supergravity transformations (18.11) can be augmented by terms higher-order in (0,0) and still preserve the transformations (18.23) of the component fields. XIX. CHIRAL AND VECTOR SUPERFIELDS IN CURVED SPACE In Chapter XXI we shall construct a Lagrangian, invariant under super- gravity transformations, which reduces to (7.24) in the limit of flat space. Before we do this, however, we must define chiral and vector superfields in curved space. We start with chiral superfields, which satisfy the covariant constraint condition Gm =0. (19.1) This reduces to D, ® = 0 in flat space. Chiral superfields contain three component fields. We could define them as the coefficient functions of a power series expansion in 6 and 0. This decomposition, however, is coordinate-dependent, for 0 and 6 carry Einstein indices. It is much more convenient to define them in analogy to Exercise 4 of Chapter V: A = Oy-5-0 o V2 a*|0=0=0 (19.2) 1 F = ~{ 99-0. These components carry Lorentz indices. They are related to the 0,0 expansion coefficients through a transformation which depends on the supergravity multiplet. The transformation laws of the component fields are found from the transformation law of the superfield ®: 50 = —649,0, (19.3) XIX. SUPERFIELDS IN CURVED SPACE 147 The parameters &4 are specified in (18.11). Since ® is chiral, we have: 6b = —E9O — &G. (19.4) The change in A follows immediately: 6A = 60| = —29,0| = —/20%7,. (19.5) The change in ¥ requires a little more work: 1 5fa = PP, D,0| (ED, — BG)9,0|. (19.6) ale 8 To proceed, we must evaluate J,9,0| and G,9,®|. This may be done with (16.14): (DDy — (-P"DyBN4 = —Tex? DpV4 + (JV? Regy'. (19-7) For 9,9,®|, Eq. (19.7) and the constraints (14.25) imply {PpF,}® = 0, (19.8) so 1 DD = ~ iy,D'DS 2 (19.9) D,D@| = —264,F. For Gj, we use (19.1), (19.7), and the constraints (14.25). These give GjD& = [G93 = -2i9,;'9.0. (19.10) The derivative DM is related to the components (19.2) through the definition (16.6) of the covariant derivative: DO = E"D,~D + Ej'D,O + Ey DO DW = EDO + EfD,O + Ey, G'O (19.11) BO = E™G,@ + EXD,O + EX, F*O. 148 XIX. SUPERFIELDS IN CURVED SPACE Restricting to 0 = 0 = 0, we find I 1 Da = 6" (240 -5 1019.9) 1 ee (04 - R Watt) (19.12) D,®| = 59,8 = 2, FO| = 5,F'O| = 0. From (19.11) we see that spacetime derivatives e7"9,, of the component fields are always accompanied by extra terms proportional to the gravi- tino field and higher components of the matter multiplet. We shall combine these terms into supercovariant derivatives D,. Equation (19.12) provides our first example D,A =e,” (a4 - a va): (19.13) Combining the above results, we find the change in x: 5%, = —V20,F — iV20,;°'D,A. (19.14) All we have left is the change in F. We start from (16.9): OF = tO — BG) De. (19.15) In Exercise 2 we show that B_DD Yd = ; {BaD} DP 2 =-3 R,?, De. (19.16) Inserting (15.39.1) for R,g,5, we discover the very important result, DAD'D, ~ 8R*)® =0 19.17 G{G,F' — 8Ryv* = 0. (eo) XIX. SUPERFIELDS IN CURVED SPACE 149 This tells us that (¥’Y, — 8R*) and (GZ; — 8R) are the covariant generalizations of the chiral projection operators D’D, and D,D’, pro- vided the superfields on which they act carry no Lorentz indices. [If the superfields carry Lorentz indices, (19.17) changes because of the curvature term in (19.7).] The @ = 8 = 0 component of (19.17) gives the first term of (19.15): 4 DPD = -3 V2y,M*. (19.18) The second term is computed in Exercise 3. Combining the two results, we have 1, wll BR BF = = 3 2M, + (t VD yal? — 1/20) (19.19) The supercovariant derivative D,z, is defined as follows: a“ 1 i a Duka = ea” (20 — eV nal? — = Wn! D, 4) (19.20) V2 V2 ° where Dida = Onda — Oma! kp- Equations (19.5), (19.14), and (19.19) give the transformation law of the chiral multiplet: 6A = ~ 20% Sta = —V2,F — i/20,;°'D,A SF = — 5 VIM, (19.21) well op es GA oe +¢ (j V Dai? — iV Deak ) Vector superfields in curved space obey the usual constraint, Vevt. (19.22) As with chiral superfields, their components may be defined through 150 XIX. SUPERFIELDS IN CURVED SPACE covariant derivatives: Cc=)| b,= -iD,V\| $x = iG;V| 1 == = PF. -~F,PV\ N= 799, + FF WV| Yai 1 [2nF:IV| (1923) a= iW) 7, = -iM| 1 1 — D=—_9'W} = 3 5 PW) = 5 DW Here we have used the superfields W, and W,, where 1 - W, = = 4 (Fie? — 8R)9,V (19.24) W, = ~j1% — 8R*)DV. These superfields are chiral and gauge invariant. Chirality is proven in Exercise 4: GW, =0, 9,W, =0. (19.25) Gauge invariance follows from (19.7) and (15.38.2): V=aA+A*, GA=9,A* =0 1 OW, = —7(GpD? — 8R)9,A = -59; {BDA + 2RD,A = 0. (19.26) Since W, is gauge invariant, we may compute its components in the WZ gauge: Vi=9V| =GV|=G,PV\=FGjV\=0. (19.27) Higher derivatives of V are computed in Exercise 7. These lead to the XIX. SUPERFIELDS IN CURVED SPACE 151 following results: W,| = —id, @W,| = —2D (DW, + DeW,)| = —4i(o"e).,D va (19.28) 1 pti in G22 Wd = —9,i'DF! + 5 (2gM* + by!79), where Bova = 65" {nt + rte + Tile) + 5 aiarnt / (19.29) Byft = eg" fo," +5 TniD — (Fatah Equation (19.28) gives all the components of W,. REFERENCES S. Ferrara and P. van Nieuwenhuizen, Phys. Lett. 76B, 404 (1978). S. Ferrara, D. Z. Freedman, P. van Nieuwenhuizen, P. Breitenlohner, F. Gliozzi, and J. Scherk, Phys. Rev. D15, 1013 (1977). EQUATIONS Ge =0. (19.1) A= D\o-0-0 = 9,0). da 2 aPla=a=0 (19.2) 1 F = ~{ PF Qy-0-0- 50 = —49,0. (19.3) (19.13) (19.20) 152 XIX. SUPERFIELDS IN CURVED SPACE 6A = ~\20%q Sfq = ~J2GF — i20,;°C?DA afl = oP = ~h amet, + C(Z Vbart ~ 12a) Vevt. c=V| $b, = -i9,V| 8, =i, | M = 5 Pa. -~9,2)V| N= 1 O°, + DP V| 1 Ya = 5 [F.-Di\V| A, =iW,| 4, = —iW, 1 1. D= -39°W) = -59,W" Ww, = 1G, 9 — spa = —4! p2" — 8R)D,V. Wi) = -i, ow) = -2D (DW + DpW,)| = —4i(o'*e),pDyv, 1 aay i GP IW| = ~o,BI! + 5@M* + b,/A;), D2! I! + Sah ~ CV aTudDa. EXERCISES (1) Verify Y (-"F,F,%, ® = 0. PUaby) Use (19.8) to write this in the following form: (DD pD, + DpD,Dy + D,DzDy)® = 0. Duta = oi"| 2 + Waki + Drida) + 5 ataouth of (19.21) (19.22) (19.23) (19.24) (19.28) (19.29) XIX. SUPERFIELDS IN CURVED SPACE 153 Show 1 D_DjDD = Dn I} Dy ~ (DpF,}FIp)®. (2) Use (19.7) and Exercise 1 to prove (19.16). (3) Show GD,D® ={G,D}D,D — DGD} when © is chiral. Use (19.7), the constraints(14.25) , and the solutions of the Bianchi identities to confirm DDD} ® = —2io,;!D,D,0 = -2i6,3D,D, 0 — 2io"[D,,D,]o —2i6,;°D,D, + 5 [eaGra — 3G — 36 4G] HI {Gp DDY = —2i6,;'D, DW — [&,Gox + & 564 ]D"O. Take the 6 = 6 = 0 components of these expressions using (19.11) and (19.12). Combine these results with (19.18) to prove (19.19). (4) Use (19.7) to check that W, is chiral. (5) Show that the transformation law for a chiral multiplet reduces to (3.10) in flat space. (6) Use (17.12) to show that (9,W, + D,W,)| in Eq. (19.28) is invariant under ordinary gauge transformations vz > vg + 4" Omf (x). (7) Prove as many of the following relations as you wish. (Be sure to work in the WZ gauge.) (a) Fp2.V| = —9.FjV| = v4 (b) DAV| = DIV| = 5 ads" G,.9,V| = 2G.V\ = Wen (C) D,DyV| = D,B,V| 1 _ =—4 VWa'o® + Wea") 154 XIX. SUPERFIELDS IN CURVED SPACE (d) Ss is > ® S a! iV (e) iV BV SS () 9,F,F4V\ = (g) (h) 9,2,2,V () DIDI V GQ) PDpBaV (k) () DPI V D,BiDV (m) 9,F,2.V| DDIN = — ith, = —2i€ hs = itil, — io,j'WP0ss = ieyghy — iops'v,3h 2 1 5 — 2itsp ( 73 nase.) = = 2iey (%, = 5 = 50,40" We ‘) =0 =0 1 = -auh(a—lecea) . 1 j - nate tavn) i (7, 3g oyege i iM* = iby | Ay — 5 oae Ve) G Paints M i, 1. ; i , = iby" (2 -3 O45 ven) -§ Oyzq¥" pM _ - - 1 = ig ( 525! soveate + Sear) N = tog Vata — 5 one — 5 vt 5g Fa Ux (ides — Zixdcs — 3e5ebex) + oF Gleb — 3€aDre — BexeD Qe) ala 94 0 Valen? as an Dek EaD xe = Datos + 5 Fl tesbre — Vaibes a i_, = —Dadas” + § Ga" Waibia — Vidas) XX. NEW © VARIABLES AND THE CHIRAL DENSITY In the previous chapter we defined the covariant components (19.2) of chiral superfields. In this chapter we introduce new © variables. These new variables are defined such that the expansion coefficients of chiral superfields are precisely the covariant components (19.2): © = A(x) + (20*y,(x) + ©*O, F(x). (20.1) In this expression, the © variables carry local Lorentz indices rather than Einstein indices. The transformation law for a chiral multiplet is given in (19.21). Our goal is to reproduce this law in the following form: 560 = —1"(x,0) dy. (20.2) The differential operator dy acts on the spacetime coordinates x” and the new variables ©*. The new transformation parameters 1™(x,0) = no (x) + OM yx) + O70. (a(x) (20.3) must be found in terms of the old parameters {(x) and f(x). The ansatz (20.2) will be justified by the fact that (19.21) may indeed be written in the form (20.2). Because (20.2) involves a linear differential operator, a product of chiral superfields still transforms as a chiral superfield. We shall now compute the parameters 7. From (20.1) and (20.2), we see 6A = = 1"0)8mA — 2m" oye: (20.4) Comparing with (19.21), 6A = ~/26%7,, (20.5) 156 XX. NEW © VARIABLES AND THE CHIRAL DENSITY we find 0) = 0 Toe (20.6) Moy = o. Next we consider 5y. From (20.1) and (20.2), we have 25% = = N12 OmA _ V2! ayake ~ 21" 0) Onda ~ 2Noyak- (20.7) Comparing with (19.21), = 1 by = VRE ~ 1YautChee"(Ou ~ Vuln) 208) V2 and using 79) and %(o) from (20.6), we conclude: NM" a)2 = 2idgi"EeQ” : (20.9) Maya = iG aj CP eg Wy? - The computation of 6F is left as an exercise. All told, we find that (19.21) may be written in the form (20.2) with the following parameters 77: nl” = 2100" + OOP, a"0'C mt =F — 10," + “00! M*O+— t 6°. a(eo"C)* — iw,,"*(a"Z) - Suelo (20.10) The variables © may be used to construct invariant actions. Before we do this, however, we must introduce the concept of a chiral density. Chiral densities are functions of superspace with the following trans- formation law: 6A = —duln" A(—)"] = 1 yA — (—Y"Oun™)A. (20.11) XX. NEW © VARIABLES AND THE CHIRAL DENSITY 157 This law is chosen so that the product of a chiral density and a chiral superfield is again a chiral density: 5AD = —Ay[™ A(—)"]@ = Ay Oy = ~duln™ A0(— J"). 0.12) This fact allows us to construct invariant actions from chiral superfields: of =5 fate 20 Ag(®) = -fatxao aul Ag(®\(—Y"] = 0. (20.13) Here g is a chiral function of ®. Chiral densities may be decomposed in terms of component fields: A =a + \2@p + @Of. (20.14) The transformation laws of the component fields follow from (20.10) and (20.11): ba = —/2tp + iapot Jf — iv2D,(o"Ca), + ipotp, o 1, + 0"). WmP 3 ¥ 2tM*a - 5 V2alot Tada +5 JV WrV moma 5f = 0,[—aij,a"o"l + on (20.15) The expression for df shows again that | d*x f is invariant. There is a special chiral density & connected to the vielbein. We shall construct this density from its lowest component: a= se= 5 det ent. (20.16) 1 2 The transformation law of e,," was given in (18.23). From this it follows that de = ce," bem! ice d"(Wmo"% — on) (20.17) 158 XX. NEW © VARIABLES AND THE CHIRAL DENSITY Comparing (20.17) with (20.15) gives the middle component: p= GN 200": (20.18) To find the remaining component, we need only compute the terms in 6p proportional to ¢: Pa = 5 V2 i810." = -V20,f - evieMt, + [€ terms]. (20.19) From (18.23) we have: 5p, = i V20,i8¢ | —i(Co"p,)(e,"e.” = een? + imacoar + [Eterms]. (20.20) Comparing the two results gives: f= -5 eM* — 5 ei arom = so"), (20.21) It requires a lengthy calculation to show that (20.16), (20.18), and (20.21) transform as a chiral density under the full transformation law (18.23) of the supergravity multiplet. In the next chapter we shall couple supersymmetric models to super- gravity. We shall find that the chiral superfield R is the Lagrangian of the supergravity multiplet. In Chapter XVII we discovered how to com- pute the components of R. Here we shall use the transformation laws of the chiral and gravity multiplets to derive the same results. We start from the lowest component, R| = -2M, (20.22) and build the full superfield in analogy with (4.11). From (18.23) we know that 5R| = ELA + ib"), — io%JM). (20.23) XX. NEW © VARIABLES AND THE CHIRAL DENSITY 159 From (19.5) it follows that 6R| = —C9,R\, (20.24) so 1 _ B,R\ = -% (0°, + ib", — io™h,M),. (20.25) This agrees with Chapter XVII, Exercise 8. In a similar way, we find: 1 2 nc DOR = 3 Cy Bim? + 5 WOW inn 1 tmnt = + Fan + VATA mn] 2 4 2 _fiomg p42 * 4 5 py 3 Hea ‘Dy +9MM + 9 Oba +5 UM — + ano ob. (20.26) REFERENCES V. Ogievetsky and E. Sokatchev, Phys. Lett. 79B, 222 (1978). J. Wess and B. Zumino, Phys. Lett. 74B, 51 (1978). EQUATIONS = A(x) + /20%,(x) + ©°O,F(x). (20.1) 6® = —7"(x,0) 5,0. (20.2) = 210o"C + OOY,a"0"F P= F = 00", + eoltM M* + gbleot T)* — iw,,2"(0"Z)p - svete". (20.10) 160 XX. NEW © VARIABLES AND THE CHIRAL DENSITY oA = —éyl[n™ A(—)"] = —1 OyA — (—)"Oun™)A. (20.11) A=a+ ./20p + OOF. (20.14) ba = —\/2p + iapo® 5pa = —V26,f — iV2D, (0% a), + ipotp, ~ 1s + HO"D)aWmp — 3 /2CM*a 1 Lo ne i GV oT abe tav Wi mood of = 6,[—ap,o"0"C + ix/2p0"C]. (20.15) atea!acte,* 20.16) a=5e= 5 dete,,”. (20. p= 7 v 20" (20.18) 1 1 a f= —zeM* — 5 el,(a"o" — oo", (20.21) 1 R= ard M. (20.22) 1 _ DR = —Z (Oa + ib", — io"TM)y- (20.25) 1 2, DOR = —5 Cae Rg + 3 OV mn 1 _ +p MTT M mn + VFM nn) 2 4 2 — ie™G pt + — * 4 = pa 3 Hea" Dad + 9 MM + 5 bbe + 5 uM - 5 Uno (20.26) XX. NEW © VARIABLES AND THE CHIRAL DENSITY 161 EXERCISES (1) Compute dF using (20.2) and (20.10). Compare the result with (19.20). (2) Derive (20.21) from (20.19) and (20.20). [You may wish to use (A.17) to simplify the yy terms.] (3) Show DDR -~ DPR =|PDYG 4 = 4i9,G". (4) Use the Bianchi identities to verify 3 a ese 5" Rn | = = (WAR + G,G*R*)| + 48 RR*| + [UnO"VDR + (Wmo")xF*R* J| ~ mF" TR + Vimo" enR * | + 66,G"| + 2129, — Vox G"| + 5 lina") = Vr") n) Gri (5) Use the results of Exercises 3 and 4 to reproduce (20.26). Beware: The calculation is tedious! XXI. THE MINIMAL CHIRAL SUPERGRAVITY MODEL We now have what we need to construct the supergravity matter cou- plings. The general case is rather involved, so we shall start here with a simpler example. We take the Lagrangian to be given by & = [00d°00,*0, + | feo(ao, +5 + 5m, + 5200/0) + ne} (21.4) This Lagrangian was first introduced in Chapter V; it is the most general renormalizable supersymmetric Lagrangian involving only chiral super- fields. In what follows, we will extend (21.1) to curved space. The tech- niques we introduce in analyzing this model will prove useful in discussing the general case in later chapters. Since the result we derive reduces to (21.1) in the limit of flat space, we call it the minimal chiral supergravity model. We start our construction by writing down an invariant action for the supergravity multiplet, Ls6.= -§ [Cor + he. (21.2) Here x? = 82Gy is the gravitational coupling, which we set equal to one. The chiral density & and superspace curvature R were computed in Chap- ter XX. Their © expansions are listed later in this chapter. Inserting these expressions into (21.2) gives Zs ¢. in terms of the supergravity multiplet: 1 1 1 Ls, = ~5 eR — 5 eM*M + 5 eb*d, + Sek Dtle — VioFnie). (213) XXI. MINIMAL SUPERGRAVITY MODEL 163 The curvature # was introduced in (17.15), R= Ofer (GQ — Og Oy? + Om One? — Oy Omee)s (21.4) while the covariant derivative %,,¥,, was defined in (17.8). From (21.3) we see that %;,, contains the Einstein action for the gravitational field. It also contains the Rarita-Schwinger action for the spin-3 gravitino. The fields M and b, do not propagate; they are the auxiliary fields of the super- gravity multiplet. Note that they enter (21.3) with opposite signs. The Lagrangian (21.1) is easily extended to curved superspace. We first write it in chiral form, 1. Ga feo -j DDO,*®, + a, 1 1 FZ Mj PO; + 3 ix POM | + he., (21.5) as outlined in Exercise 6 of Chapter IX. We then add the supergravity action (21.2), and replace 0+ 0, @?0+d?@26, and —1DD— — 8R). This gives the action (21.1) in curved superspace: ge Jeers] -ae - 499 — 8R)0,*o, 1 -3 (GD — 8R)[c®; + &O,*] + d + a; 1 1 + zmsPO, + n0.00| + he. (21.6) This Lagrangian describes the minimal chiral model. It reduces to (21.1) in flat space. The c and d terms are included because they arise from shifts in the superfields ®;. They vanish in flat space. Note that gauge invariance restricts c,; = 0 unless ®; is neutral. Equation (21.6) contains two types of terms: those with the chiral projector (YP — 8R), and those without. The terms with projector are curved-space generalizations of the chiral kinetic energy. Those without are curved-space extensions of the usual superspace potential. We will 164 XXI. MINIMAL SUPERGRAVITY MODEL emphasize this distinction by writing (21.6) in the following form, 1. L= feo 24| 7199 = 8R)Q(0,0*) + a +he., (21.7) where Q(0,0*) = ©,*@, + cP; + &@;* — 3 is the superspace kinetic energy, and P(®) = d + a,@, + 4m,O, + 49:,0,0,®, is the super- space potential. In Chapter XXIII we shall see that this distinction is pre- served in the general case, where Q and P are arbitrary functions of their respective superfields. The Lagrangian (21.7) has a long expansion in terms of component fields. To find it, we need the © expansions of ®,, &, R: ©, = A, + J207, + OOF; 26 = ef1 + Oop, — OO[M* + Fe", ]} R= 4 | + O[o76*W,, — iot,M + ip.b*] 1 _ 2 + 00-5 R + iV + | MM* | ah — ico. + tgam — iWeb + 3 b'ba — ie! ‘DD’ + 3M 3 Waa bb + fetta + vara 218) The components of 2; = (GJ — 8R)®,* can be computed with the help of (19.7): 1 _ — 4907 PATD* — 8RY0;*|. (21.9) The necessary ingredients are given in Exercise 1. We find: XXI. MINIMAL SUPERGRAVITY MODEL. 165 4 a = —4FF + 3 MAK + © | —4i,/20°D.Z, — 5 Ro" z, 4 * ab, italy i a) + ZAN20% an — fo°GeM + itfab') + eo} —4e,"D,D°A¥ — : ib,D*A* 2 7 xabs bi Bz. 8 — ZV + 22D, — 5 MOPS 2. 57 e 157 oy en — F iv Tb" + 5 iVDO Tb, 4 1 sranb se +3 A¥ 58 F iO" Way — ie!" D mb" 2 1 1 4 . + 5 MPM + 5 bbt + 5 EM — 5 0° 3 3 + fetes + Vara |} 21.10) Here we have used the following supercovariant derivatives: 1 ope DyAt = 02" WA? ~ 5\ 2 0h eT = "Dyin + 53ND PopeD AT — 52a? 241411) where PnZis = Omiig + Tip Onbiee The superfields &, R, ®;, and =; allow us to compute any supergravity Lagrangian involving only chiral fields. The expansion of the Lagrangian (21.7) contains kinetic terms for the physical fields A;, ¥;, @m“, and ,,”, as well as terms involving the auxiliary fields M, b,, and F;. The Lagrangian also has higher-order interaction terms, such as nonrenormalizable four-fermion couplings, which are sup- pressed by powers of Newton’s constant. For ease of exposition, we will write the full Lagrangian as follows, L = Lyin + Los + Lourie» (21.12) 166 XXI. MINIMAL SUPERGRAVITY MODEL where 1 Lain = GOR — CAs OA i = 510" Dal + Zid” Pt] 1 a8 7 "Dp C0" FN nn _ VO Von] + [ebm QA; ~ nO AD Oat 1 - = V2 GAIL OAs + Wud" OA] 1 a + V2 D0" Van + QF" — 5 V2eP ino". — 5 V2eP RA", 1 1 — 3 eP ikki — 5 Pit. — ePY,e by — eP* WoW, (21.13) A is the kinetic part of the Lagrangian, Laux = 5com — 3(log Q).F¥|? + eQ(log Q),;.F FF - ; Qb,b* — 52 OA; — Qe Dy A¥)D™ 1 i 5 = HIB + 2A ~ Qe TNO 6 — ePM* — eP*M + eP;F, + eP&F* (21.14) is the auxiliary field contribution, and Yjyqni¢ Contains four-fermi terms that we ignore for the moment. In (21.13) and (21.14), Q and P are the same as before, except that the superfields ©, and ®;* are replaced by their lowest components A; and A¥. The subscripts on Q and P denote derivatives with respect to the scalar fields. For example, P; = (6/0A,)P and Q» = (3/0A*)Q. XXI. MINIMAL SUPERGRAVITY MODEL 167 To proceed further, we must eliminate the auxiliary fields from Y,,,. This is most readily done by shifting, N = M — 3(log Q),F¥. The shift decouples N and F;, and allows the equations of motion to be easily solved: N = 9PQ71 Qlog Q);.F; = —Ph + 3P*(log Q) b, = 310,24, = 0.0,Ax)Q7! = 3 5 - + GN QLba — QF". (21.15) Substituting (21.15) into (21.14), we find Laue = —IePP*Q™! — e(log Q);'[P; — 3P(log Q),][P# — 3P*(log Q), JQ? _ Fel2,4i — Qn 0,AF IQ; 0A; — 2p 0"A¥) = ino; — J2Q4W" — QTWYJQ* + --, (21.16) The dots denote additional four-fermi terms that we absorb in Lyuartic- Equation (21.16) contains derivative terms, fermion masses, Yukawa cou- plings, and the scalar potential, which we shall call /(A,A*). The above expressions are not quite ready for model building. We must still check the normalizations of the physical fields. From (21.13), we see that the gravitational action has an unconventional Brans-Dicke form. This normalization can be fixed by performing a field-dependent Weyl rescaling of the gravitational field: en’ > e,” exp(A), (21.17) where 2a) = 3 (21.18) exp(24) = O° . This transformation restores the canonical normalization (21.3) for the Einstein action. The matter-field normalizations can be restored through a field-dependent redefinition of the spinors, Xi > exp(— 4/2) x; Wn > eXP(A/2) Uns (21.19) 168 XXI. MINIMAL SUPERGRAVITY MODEL followed by an additional shift of the gravitino, Vin > Vim + ix/26 ni (21.20) Adding the four-fermi terms from Pyuaic, and performing the transforma- tions (21.17)-(21.20), we find our final result for the supergravity matter coupling: 1 La 5 eR ~ Kip yA OMAP i - on — 5 eK Lo" Indy + Lj" Init} 1 os ~ + 3D FL nin = WF eLnVn) 1 - +q ec!" KOA; — Ki APU Om i - ~ GAKi WK Gnd — Kur OyAB) = AK joe OnAn — Kijene On AB) Z:0"%; 1 lg _ - 5 VK iy O,AF LOEW mn — 3 V2eKip O,AZjF"0"D 1 - -, + 4 Kili oa m + Vn Wi F niki ! = Fy 7 +36 CLK ijeKuee — 2K jones + 2K KinseK jrcor Limi sla” Le te exptk2){ PHY oy — PUG, = 5 V2DPro, ~ 5 /2DPPTT, 1 - 3 Pu + KiP + K;D\P + Kj;D;P — K;K;P 1 — KK iy.D,P nin; — 3 [Pie + KyppP* + KpDpP* + KpDypP* — Kp K,P* — KK er DaP"Itis} — e exp(K)[K""(D,P)(D;P)* — 3P*P], (21.21) XXI. MINIMAL SUPERGRAVITY MODEL 169 where D,P = P; + K;P, K(A,A*) = —3 log(—Q/3), and K'” = Kj’. In this expression, Q(A,A*) = A¥A; + c(A¥ + Aj) — 3 and P(A) = d — aA, + 3m4;A; + 49;A;A;Ay. In Chapter XXIII we will derive a similar result for the general chiral coupling. Equation (21.21) gives the full supergravity coupling of the minimal chiral model. It has properly normalized kinetic energies for all physical fields, and the full set of four-fermi terms is included. Equation (21.21) is automatically invariant under supergravity transformations (up to total derivatives) because it was derived from a superspace formalism. It also has the correct flat-space limit. From (21.21) we see that the supergravity scalar potential emerges in a form that will turn out to be quite general: ¥ (A,,A¥) = exp(K)[K""(D,P)(D;P)* — 3P*P]. (21.22) Note that this expression is not positive definite, so the connection between the potential and supersymmetry breaking is more subtle than before. In Chapter XXIII we shall see that the signal for spontaneous supersymmetry breaking is # 0. Equation (21.22) shows that supersymmetry can be spontaneously broken with zero vacuum energy. The preceding expressions are all written in terms of the real function K(A,A*). In the coming chapters, we shall see that this function is called a Kahler potential, and that (21.21) and (21.22) have a natural interpreta- tion in the language of complex geometry. REFERENCES S. Deser and B. Zumino, Phys. Lett. 62B, 335 (1976). D. Z. Freedman, S. Ferrara, and P. van Nieuwenhuizen, Phys. Rev. D13, 3214 (1976). EQUATIONS 6 2 . Lo, = —y [POR + he. (21.2) 1 1 1 Ls6.= 308 - 3 °M*M + 3 ebb. + ebm nite — Wo Dati @13) 170 XXI. MINIMAL SUPERGRAVITY MODEL 2 lL =e +, ge fa ©28| ~3R - (GF — Ryd, 1 - _ gF — 8R)[c®; + &®,*] + d + a, 1 + m,0,0, +} 3m 3 In 9] + he. (21.6) ©, = A, + /20x, + OOF, 26 = ef1 + io%), — OO[M* + HF" y,]} 1 - R=; {w + O[o°S Wy, — io,M + iW,b"] 1 a 2 + oo] -52 + IVF Way + 3 MM* + Lb'b, — ie," Ogb* + PUM — A y,orbbe 37 Oa a Pm +5 Wh 3 Mao" e 1 ~ - + go WF othea + vain (21.8) - 4 meeps 2 pap = E, = —4Fh + 5 MAP + ©) —4i/20D,Q, — 5 V20"b.S, 4 aT . + AP 20% hay — io M+ wan} + 60} - 4e."D. mD AF _ 5 ib DtAa* 2 pe 5 8 = Vf + 22D — 5 MEE a ae Loe na eo -3 i 2 ab" + 3 iV 2p, 60Gb. 4 1, . a + 34% 3 R+ We Way — ica" Db 1 __ + 5 MM + ye bt + 00M — a + fies + Hara]. 21.10) XXI. MINIMAL SUPERGRAVITY MODEL 171 a 1p _, DAP = 6." — 5 V2 act a“ 1 Dike = Ca" Pid + 5 Vivos'B,At = 5 V2baFP. 2111) -} CR — eK ip OA, O"AF _ 5 eK nl10" ul) +H" Duis] + beet Dall = WoIDnVn) + pe, adi — Ke OAPW Onn _ Fe KiplKy Oye = Ky 0,A*) = UK pe Ome — Kijoee OmADIUO"Z) ~ pV2K Op AFLiO"O" Vm — SV2eK 8,4 LjF"O"Wm + pki oelm + mo" 10,7, + ig Kine = 2K ipener + 2K K nseK jecor iO mile Le + cexpik/2)| —Phyats — PUP - 5 V2DPx0%b, - 5 2D.PYZa%, _ Str) + KjP + K,D,P + K,D;P — K,K;P — KK ijDyP tit; — 3 (Pte + KiapP* + KD pP* — e exp(K)[K'"(D,P)\(D,P)* — 3P*P]. (21.21) 172 XXL. MINIMAL SUPERGRAVITY MODEL ¥(A,,A*) = exp(K)[K'"(D;P)(D;P)* — 3P*P]. (21.22) EXERCISES (1) Verify as many of the following relations as you wish: (a) @*| = A* (b) F,o*| = 0 () 9o*| = /2, (d) 9,D,0*| = 0 (0) GeDj@*| = egg F* (1) G,P,o*| = 0 (2) D,DO*| = —2iog;4D,A* (h) 9D,9,0*| = 0 (i) «PD ,D*| = 0 (i) G,GD0*| = 0 2 5 = 3 V2:p%aM* jD,0*| =0 (m) 9,9,9,0*| = —i2)20,;D.i, 1 + 6 Veins — 37ab.g — 3€spb.5%') (1) 2,9,j0"| = 2resj0u5BP +4 reba (0) 9. | = B,A* = 0." DyA* - 3 Tosi (p) DD w*| =0 @ 9.9.0" = 7 /2o.57M* XXI. MINIMAL SUPERGRAVITY MODEL 173 () DI*| = 205s : .. = Vi(enoni + eu doB,A* — = ia) y2 v2 i 5g EMP — Babyy — Seybnit} (9) F,90*| = V2 {ox = (0) ZB" | = e." PDA + 55 VBE o.g7 M* Lief i gy + ZV Dyn — 34% (Zab oa — 3%sbpp — Seba} (u) F9,9,0*| = 0 (¥) 9,9.D,0*| = 0 ; 1 a (Ww) D,F,90*| = iV 26ei7/D,R* | - 3 (0°0°e).pM*D,A* (x) PD,G,PO*| = 16e."D_D,A* + » ib*D,A* A 8 - = 829557 + MAF +S 2 8 pe 4 pe +5 iJ2p,7b" — 3 V2.5" 7.. (2) Use the above relations to derive (21.10). (3) Verify that Y,,, and Y,,, are given by (21.13) and (21.14), respectively. (4) Show that the Weyl rescaling (21.17) takes 1 3 -29 34 -1 gp? > oom 25 Fegmm . 5 ek 42 6p Q O"Q. + 3 OmLeg”"Q~ * 6,Q] Fn + - (5) Use the result of Exercise 4 to show that (21.17)-(21.20) restore the proper kinetic energies in (21.21). (6) Use (21.14) and (21.17)-(21.20) to check that the potential (21.22) is indeed correct. 174 XXI. MINIMAL SUPERGRAVITY MODEL (7) Show that for Q(4,A*) = A,;*A; — 3, the field redefinition, A, +a Awl — ; A=, ae SS 1), 1" 1+ 4fa*A,’ ‘ 1+ 4a*A, G40 induces the Kahler transformation, 1 = 4lal? K(A',A*) = K(A,A*) — 3 log 2. (4.4) = K(A,A*) ~ 3 log ye As discussed in Appendix C, this is an isometry transformation— it leaves the Kahler geometry invariant. Note that after such a trans- formation, the supergravity potential Y can again be related to a superpotential P of third order. This feature characterizes the mini- mal chiral supergravity model. (8) LetQ = A*A — 3and P = p(1 + 4/34). Show that the potential /* vanishes. Since the potential does not determine the expectation value where = V3a. XXII. CHIRAL MODELS AND KAHLER GEOMETRY In the previous chapter, we constructed the minimal coupling of chiral superfields to supergravity. We found that the resulting Lagrangian could be written in terms ofa Kahler potential and its derivatives. In what follows, we will begin to explore the relation between matter couplings and Kahler geometry. We will work in flat space, where the connection first appears, leaving the curved-space generalization until Chapter XXIII A brief in- troduction to Kahler geometry is given in Appendix C. We start by studying the most general Lagrangian that can be built from chiral superfields ®‘, for i = 1,...,n. This Lagrangian takes a very simple form, ga fvrodoK@o*) + [ feraro + ne} (22.1) Here K and P are superfields, with power series expansions in terms of the chiral superfields ©’, K(,O*) = Yi iygjy eo yg DY DOH tw Jase iM PO) =P Gis. in BO. (22.2) To find the component Lagrangian, we must expand K and P in terms of the 6 variables. The expansions of the individual fields were given in Eggs. (5.3) and (5.5). For the superpotential P, Eqs. (5.7) and (5.8) im- mediately extend to OP(A) oAt + wo\r OP(A) _ P(®) = P(A) + 267! (22.3) Nie x eAt 176 XXII. CHIRAL MODELS Here all component fields are functions of y" = x" + i000. The conju- gate superpotential P* has an analogous expansion, oP*(A*) ) oA ap*(A*) 1 _,_, 0 P*(A*) gate FR cast gani(> P*(O*) = P*(A*) + 207 — + 00 {re (22.4) where the fields now depend on y*. The 6 expansion of K(®,®*) can be computed starting from the monomial Kyy = B8 + DDH... Dia, (22.5) Its 6606-component may be found with the help of (5.9) and an appro- priate interpretation of (22.3) and (22.4), _ a, ; 1 0>(Al ++» AIX) Kyy =-** + 00002] FE (iss + gisy — = ykye nm * {| Bae A AN yl aaa ge é \y OAR. Aine [Pao (At any = 3 ie A 4 _ apes a ; a(ait + als ig tah atin, ptama, (AA)! ang where we have used partial integration. This result can be rewritten more elegantly in terms of derivatives of Kyy|, the lowest component of the superfield (22.5), Kael pi Kya = 00° + + 000 a FSi _1__@Kynd izigk 1 O° Kyatl _ prai,ink 2aadaniaa®” "" ~ 294% Gaia hF 1 d*Kyadl 4 aa AoA awe MEET © Kyul 5 giam gai — , 2’ Kuml ojoms vi ~ Garages Om OMAN — Tria gj HO Oni PK yal ceemi a ~ igang CO" ana. eon XXII. CHIRAL MODELS 177 Equation (22.7) is also true for the full polynomial K. The expression simplifies in the notation of a Kahler manifold, where 0 8 fr = 54 aan Xl Dijk = OAe Gip = Imp ik Ginnae = Fak Iie = Gime prs (22.8) One finds = : 1 : K =::+ + 0000 {iP ~ 5 Gil FieP BT 1 mt Erik 1 ji yk Sf Ty Ime RELL + 7 Gin nek el = Gije OA! OM A* — ig, je Zi5™ Om — mel Rory ana’. (22.9) We now have all we need to write the full Lagrangian in terms of com- ponent fields. Substituting (22.3) and (22.9) into (22.1), we find ; 1 ‘kp L = OjpF'F* + GIrwcok LEE 1 —_ oP —F {; Ginel Pe 7 sat 1 op* — Fe {5 Imi LL — sah = Gijp FAO" AN — igi ZE"Dnyi 2 2 px 1 @P 1 oP 22.19) iyi —— zz, ~ aaa ** ~ 2 Gaviaaei Here D,,7' = Onyx’ + Ti, 6,A/z* is a covariant spacetime derivative, as- suming y' transforms like a contravariant vector under the transforma- tions (C.1) on a Kahler manifold. 178 XXII. CHIRAL MODELS The auxiliary fields in this expression may be eliminated by their Euler equations, a ; oP’ 9inF — 5 Gj mee + gaa = 9 (22.11) Substituting into (22.10), we obtain the final form of the component Lagrangian, L = ~Gip dg A! OPA — igi LE"D yl 1 ae + Rime CLT Il i a — 5 DD Prix) — 5 DeDy PATE 2 — g'*D,PD»P*, (22.12) where a DP = aaP Dp,p = ——p ~ 142, P (22.13) ee 0A'oAl ry ee " Equation (22.12) describes the most general supersymmetric coupling of chiral multiplets. We have used Kahler notation to illustrate the geo- metrical nature of the result. Invariance under Kahler transformations is manifest. Each term in the Lagrangian (22.12) has a natural interpretation in the language of Kahler geometry. The scalar fields should be thought of as the coordinates of a Kahler manifold, and the fermions as tensors in the tangent space. The Lagrangian (22.12) is a supersymmetric version of the sigma model, expressed in geometrical form. In superspace notation, the appearance of the Kahler geometry can be traced to the invariance of the Lagrangian (22.1) under the superfield Kahler transformation: K(®,*) + K(®,0*) + F(®) + F*(®*). (22.14) This invariance will play an important role in what follows. XXII. CHIRAL MODELS 179 REFERENCES B. Zumino, Phys. Lett. 87B, 203 (1979). D. Z. Freedman and L. Alvarez-Gaumé, Comm. Math. Phys. 80, 443 (1981). EQUATIONS L= faroaK@o”) + [ ferore + ne} (22.1) L = ~GipbyA' OMA” — igipZE"D yy! 1 + 4 Rio COL 1 jiyd 1 ZIT - 3 PiDiPx'x ~ 3 Pie x P* LY. — g DPD »P*. (22.12) DP = ppp -———p ro (22.13) Oy @AigAl SOAR " EXERCISES (1) Check that the Lagrangian (22.12) is invariant (up to a total derivative) under the following supersymmetry transformations: dA = V2Ey op* ix OP* dex’ = iV20"E8,,A' — Ty, deAly* — 24) aaa I et K = A*‘A! an = 2,A' + $m,A'A! + 4g,,,A'A/A® Show that 2) Let K = A*A! and P = 1,4! + 4mA'Al + 4g, ‘AIA Show thi (22.12) reduces to the renormalizable Lagrangian given in (5.13). XXIII. GENERAL CHIRAL SUPERGRAVITY MODELS Having discussed the role of Kahler geometry in flat space, we will now compute the most general coupling of chiral superfields to supergravity. As in flat space, we will find that the component Lagrangian has a natural interpretation in the language of Kahler geometry. Motivated by our discussion in Chapter XXI, we take our superspace Lagrangian to be 1 3 ad + 2 ga [vor ° (GG — 8R)exps—— K(oo*)$ + x2P)| + he, K 8 3 (23.1) where K(®,0*) is a hermitian function of the superfields ©! and ®*/, and P(®) is the superpotential. The exponential form is suggested by the rela- tion between K and Q below (21.21). Expanding in x”, we see that K is the flat-space Kahler potential, Go -§ [wear + Jvors| a9 — 8R)K(0,0*) + nro | K 8 +o the. (23.2) In this chapter, we will find that K is a Kahler potential in curved space as well. The Lagrangian (23.1) is manifestly invariant under supergravity trans- formations. Its component form can be found using the techniques in- troduced for the minimal case in Chapter XXI. The steps are virtually identical; there are just a few extra terms that follow from the general nature of K. At the end of the computation, one finds precisely Eq. (21.21), where K is now an arbitrary real function of the scalar fields A’, the lowest component of the superfield K(®,0*). Equation (21.21) gives the component Lagrangian in terms of K and its derivatives. It can be written more compactly if we use gj and Rj jexer, the metric and curvature of a Kahler manifold. In this form the geometric XXII CHIRAL SUPERGRAVITY MODELS 181 invariance of the Lagrangian is manifest. Comparing (C.10), (C.18). and (21.21), we find 1 . L = eR ~ Gi. Ai OMAP 2 = iGO" Dyfi + CMT, Dan _t 2 ; 1p : - V2€9;p: OAM YO FY gy — ZV 205 ALE y 1 : ~ ae +] CG ipLie UCD m + Vino OW" yon 1 ik Zige ~ geld dee — Ripe LLL — e exp(K/2) {prc + PYF, + 5 V2D,Py'o, + 52D POU, 1 1 nic) + IDPH! + 5 PD PLE — eexp(K)[g'"(D,P)(D,P)* — 3P*P]. (23.3) The covariant derivatives are defined as follows: a : 1 ; : Dyfi = Opi! + LOm + Upc OmA tl — Z(KjOnA! — Kp Om AM VQ ~ , Log ; Gailn = Onin + Wan + 4 (Kj OmA? ~ Kj OmA® Wn DP = P, + KP QD ,P = Pi + KijP + K,D\P + K\D,P — K\K,P —T%,D,P. (23.4) The covariant derivatives contain the Christoffel symbols for the Kahler geometry, and the spin connection (17.12) for spacetime. Note that they also contain a U(1) connection proportional to Im(K;0,,A4/). The meaning of the U(1) connection will become clear as we proceed. 182 XXIIL CHIRAL SUPERGRAVITY MODELS The Lagrangian (23.3) is invariant under supergravity transformations because it was derived from a superspace formalism. It is useful, how- ever, to verify the invariance directly, using the following supergravity transformations: = Co, + TO% yn) = yx = iV2o"0D,,A' — Ti, 5,Aiy* 1 a + 4 (KjO,A) — Kp d,A*y — V2eKg'"D, PH . 1 Odin = 2Dnk — 4 (Kj 5A! — Kp 5A” Wy i j > 5% me Snio"e + ick? Poe, (235) where Y,,{ includes the U(1) connection, C. (23.6) 1 . Dk = Ono + [Om + 4 (Kena? = Kp 0, AME Note that the transformation for the field y' indicates that supersymmetry is spontaneously broken whenever # 0. In this case, x! shifts by a constant and plays the rol¢ of the Goldstone fermion. To check the Kahler invariance, let us first examine the component Lagrangian (23.3). Under a Kahler transformation, K(A,A*) > K(A,A*) + F(A) + F*(A*), (23.7) the metric, Christoffel symbols and curvature terms are all invariant. The U(1) connection is not: Dutt > yt — 5 Onli Fy Fan > Dain + 5 Om Pi: (23.8) XXIII. CHIRAL SUPERGRAVITY MODELS 183 The Kahler invariance is restored if Kahler transformations are accom- panied by Weyl rotations of the spinor fields, yi > exp +5(m F)yx Wy > exp 5 (Im F) py. (23.9) With this rule, the kinetic terms in (23.3) are invariant under the combined Kahler-Weyl transformations. The Kahler-Weyl invariance insures that the kinetic terms are invariant under field redefinitions (such as isometries) that induce Kihler transformations of the Kihler potential K. The superpotential contributions to the component Lagrangian con- tain explicit factors of K, so their invariance is not automatic under the Kahler-Weyl transformations. For example, the scalar potential V = eX[g'"(D,P)(D,P)* — 3P*P] (23.10) is not invariant unless Pe Pp (23.11) as well. With this choice, the D,P transform covariantly, D,P > e~*D;P, (23.12) and the full Lagrangian is invariant. Note that (23.11) does not in general preserve a polynomial structure in the superpotential (see Exercise 21.7). In mathematical language, the transformations (23.9) and (23.11) imply that the spinors and the superpotential are not ordinary functions, but rather sections of appropriate line bundles over the Kahler manifold. If the manifold is nontrivial, the combined Kihler-Weyl invariance is neces- sary for the Lagrangian to be globally well defined. Locally, however, we can simply think of the geometrical notation as giving a convenient short- hand that is useful for describing the full set of supergravity couplings. The Kahler-Weyl invariance of the matter couplings can also be seen from the superspace Lagrangian (23.1). Now, however, the superfield 184 XXIII. CHIRAL SUPERGRAVITY MODELS Kahler transformation K(0,0*) + K(®,®*) + F(®) + F*(@*) (23.13) must be accompanied by a super-Weyl transformation of the vielbein. A super-Weyl transformation is defined to be a superfield rescaling of the vielbein, consistent with the torsion constraints (14.25). In the exer- cises, it is shown that the most general such transformation is of the form bEy! = (2+ DEy! SEy? = (25 — D)Eyt + 5 Eutlooss*E, (23.14) where and ¥ are chiral superfields, r= 9X =0. (23.15) This implies ~2(22 — X)R — pat 4a = —(2 + DG, + iD,(Z — BD). (23.16) Wt OR 6G, The transformations (23.14) and (23.16) determine the super-Weyl trans- formations of the supergravity multiplet. The super-Weyl transformations of the matter fields are also paramet- rized by & and =. They are defined in such a way as to preserve the appropriate constraints. For example, a super-Weyl transformation of a chiral superfield is given by 5b = wdO, (23.17) while that of a hermitian vector superfield is just Vv =w(h + Dy. (23.18) In these expressions, w and w’ are called the Weyl weights of the respective superfields. XXIII CHIRAL SUPERGRAVITY MODELS 185 Equations (23.14)-(23.18) allow one to find the super-Weyl scalings of the various component fields. These, in turn, can be written as variations of superfields in the new © variables. After a small computation, one finds 6 56 = 658 + (S78) 30" 30 = wed — +o 5(DD — BRU = (PF — 8R)[(w' — 4)Z + (w’ + 2Z]U a. _s DG —8R . aor (22 )U, (23.19) where S* = O2E — | + CODE], (23.20) and U is an arbitrary hermitian superfield of weight w’. The transformations (23.19) induce a variation of the superspace Lagrangian. For w = 0, we find 3 8 - bf = foro 26|5 (GF — 8RY& + De-*3 + czr| + he. (23.21) This is precisely a Kahler transformation, bf = f wore| a9 — 8R\F + F083 — FP] + he, (23.22) where P is scaled to e~*P in accord with (23.11). Comparing the two transformations, we see that (23.21) cancels (23.22) if F = 6. With this choice, the superspace Lagrangian is invariant under combined Kihler- Weyl transformations. It is a useful exercise to show that superspace Kihler-Weyl transformations induce local Weyl rotations (23.9) of the component fields. An arbitrary super-Weyl transformation can be used to change the form of the Lagrangian (23.1). In particular, a super-Weyl transformation with 186 XXIII. CHIRAL SUPERGRAVITY MODELS. a finite parameter —6Z = log P simplifies that component expression by rescaling the superpotential to one. Of course, this is an allowed trans- formation only if the expectation value

is nonzero. This is not an innocent assumption: it gives a nonvanishing contribution to the cos- mological constant. This contribution can be canceled only if supersym- metry is spontaneously broken. The transformation with —62 = log P changes the Kahler potential K to a new potential G = K + log P + log P*. Since this is a Kahler transformation, the geometry is left invariant. Therefore, to find the new Lagrangian, we simply replace P by 1, K by G, D;P by G,, and 9,D,P by Gi + GG; — TG. This gives 1 i ; L= “3 CR — eGix0,A' OMA = 1G: LE" Dyfi + eT Dyn 1 2 V2€9;5 A ZI" 1p = 5 V 2607 CAML O"O Vm — 1 - a + 4 eaiplic™Wso lin + Uno" D" On? 1 | 2 iykaist ~ g LGirduer — Rime LPT — Ly + Uae + 5 V2Gr0%G, + 5 2G 1 aa + 5 [Gj + GG, — TRG]? 1. * + 5 LGry + GGp — Vip Gyl7'% — ee [g"GiG» — 3]. (23.23) Let us now examine the physical content of (23.23). We first note that the kinetic terms are properly normalized if gj. = 6, + +++. We shall always assume this to be true. We then remark that the potential V = e[g"GG» — 3] (23.24) XXII CHIRAL SUPERGRAVITY MODELS 187 is extremized if <@¥/0A') = 0, and that the resulting cosmological con- stant is zero if = 0. Taken together, these conditions are satisfied if (GG = 3 = 0, (23.25) where G' = g'"G;, and V,G; = 6,G; — Tj,G). The scalar mass matrix is found from the second variation of ¥. It is of the form Mi M. ‘) i ii)’ (23.26) (us fe May with M2, = CLV,GV.G* — RijanG'G” + gij-Je% Mi; = # 0, this mixing must be removed to find the physical mass matrix. Of course, it is always possible to diagonalize the coupling by redefining the fields. It is more instructive, however, to recognize that the mixing has an important physical origin. 188 XXIII. CHIRAL SUPERGRAVITY MODELS To see this, let us consider the supergravity transformation of the field n = G,z', given by (23.5) with the appropriate substitutions, 6&n = ~/2e9?G'G,E + = 3 Imo +>, (23.29) where we have used the fact that the cosmological constant is zero, = 3. We see that 7 transforms by a shift. This indicates that 7 is a Goldstone fermion, and supersymmetry is spontaneously broken. Exactly as in ordinary gauge theory, the Goldstone fermion can be gauged away through a supersymmetric analog of the Higgs effect. In this “unitary gauge,” all terms proportional to G,z' vanish identically. This removes the gravitino-Goldstino mixing, and allows one to read off the mass matrix for the spinors y', subject to the constraint G,z' = 0. Of course, it is also possible to find the spinor mass matrix in a gauge- independent manner, by diagonalizing the terms quadratic in the fermion fields. The necessary field redefinition is suggested by the above arguments. We find Tn =n + 1 amg on + 52a. (23.30) With this choice, the mixings are eliminated and the mass terms are diagonal: my {io + ha", 1 1. Lj + 3 exp +5 (Im F)z Wn > exp (im F) Up. (23.9) Y = &[gi"(D,P\(D,P)* — 3P*P]. (23.10) Str M2 = Yo (-1)™Qs + 1) TrM? sph s = An — Amz — 2R,.G'G* m2. (23.34) EXERCISES (1) Given variations 6E,,4 and dy;54 of the vielbein and the connection, show that the most general variation of the torsion Tc,“ is given by XXIII. CHIRAL SUPERGRAVITY MODELS 191 OT cy* = DcHy* — (—)"PgH4 + Qen* — (=) Quc4 + Teg? Hp* — He?Tpp* + (—)* Ha? Tc*, where H® = E4™ Ey ® and Qcg* = EM bb yg". (2) Use the results of Exercise 1 to show that the most general Weyl re- scaling of the vielbein, consistent with the torsion constraints, is of the form (23.14). (3) Find the Weyl rescalings of R and G,;,. Check your results against (23.16). (4) Show that the conditions (23.25) imply that the scalar potential (23.24) is extremized with vanishing cosmological constant. (5) Compute the scalar mass matrix (23.27). (6) For infinitesimal n, show that € = ny/2/6my transforms y to zero. (7) Show that the matrix (23.32) has a zero eigenvalue, with eigenvector proportional to G'. (8) Verify the mass sum rule (23.34). (9) Show that Str M? = 0 for the minimal chiral model, where G = —3 log(1 — $4;*Aj). This is an important property of the model because most radiative corrections are proportional to Str M?. XXIV. GAUGE INVARIANT MODELS In Chapter XXII we studied the most general coupling of chiral super- fields in flat space, LH f 22020 K(O1*4) + [ fear + he} (24.1) We found that K has a natural interpretation as the Kahler potential for a Kahler manifold ./. We also noted that the action (24.1) is invariant under the Kahler transformations, K(@',0*4) + K(@ib*) + F(@') + F*(b*), (24.2) where F is an analytic function of the superfields ©‘. In this chapter we will gauge the analytic isometries of the Kahler geometry, and in this way generalize (24.1) to include vector fields. We will take advantage of the fact that K transforms by a Kéhler transformation under each of the analytic isometries of 7. The analytic isometries of a Kahler manifold are generated by holo- morphic Killing vectors, XY = Xgl) oO 6a 5 XH = XH GH) aa (24.3) where the index (b) runs over the dimension d of the isometry group G. As shown in Appendix D, this implies that Killing’s equation reduces to the statement that there exist d real scalar functions D(a,a*), such that ip X81 = i D ir aa gip XO = i D®, (24.4) éa*i XXIV. GAUGE INVARIANT MODELS 193 The D are known as Killing potentials. They are defined up to constants c®, D® + D® + c. In what follows, we shall see that the freedom to redefine the potentials is related to the Fayet-Iliopoulos D term intro- duced in Chapter VIII. The Killing vectors X and X* generate independent representations of the isometry group G. They obey the Lie bracket relations [X@,X] = —faexo [XO x40] = —frex*o [X,X*] = 0, (24.5) where the f“” are the structure constants of G. In Appendix D it is shown that the Killing potentials D can be chosen to transform in the adjoint representation, ai [x 5 + xxi a D® = —f%* Do, (24.6) This fixes the constants c for non-Abelian groups. For each U(1) factor, however, there is an undetermined constant c. Under an isometry in G, the variations of K and P are determined by the Killing vectors X, 5K = [eX 4 ok OX*O] K OP = XP, (24.7) The variation of the superpotential must vanish for the action to be in- variant. The variation of the Kahler potential, however, does not need to vanish. As shown in Appendix D, it can be cast in the following form: OK = FO 4 gXOpH@ — j(g — gD, (24.8) where F® = XK + iD is an analytic function of the coordinates. For real parameters ¢), (24.8) is just a Kahler transformation. For complex e, it is not of Kéhler form; there is a change in K proportional to the Killing potential D. The fact that (24.8) reduces to a Kahler transformation for real e® im- plies that the action (24.1) is invariant under the rigid isometries of the manifold ./. For local motions, however, the story is more complicated. This is because the parameter e must be promoted to a chiral superfield 194 XXIV. GAUGE INVARIANT MODELS, in superspace. In this case, the variation of the action (24.1) is 3 = [d0d05K = frorstarre + AFOFAO — (AM — At) DO] = -1 fPoe5(a” — A*™)D, (24.9) where A“ is a chiral superfield with lowest component ¢, and the D“ are hermitian functions of the chiral superfields ®' and ®*/. In the rest of this chapter, we will see how to construct a supersym- metric gauge theory, invariant under the isometries parametrized by A“. We will add a term to the action whose variation exactly cancels (24.9), using the formalism developed in Appendixes E and F. We will find that the counterterm involves the vector superfield V = VT, where the T are the hermitian generators of the isometry group G. Since « is complex, we must study the complexification of G, which we call Y. An arbitrary element of Y can be written in the form gH eT }urT (24.10) where u and v™ are real, and as above, the T™ are the hermitian gen- erators of G. Equation (24.10) splits g into the product of a hermitian and a unitary matrix, which can always be done. Given the complexification Y of G, the space Y/G is constructed by identifying elements g and g' € G if g = g'u’, for some w’ € G. Thus a point of the coset can be represented by pa wor, (24.11) The matrix v is an element of Y, and the v are coordinates of Y/G. The group & acts naturally on the cosets 4/G by left multiplication on v: v' = gov. To find the transformation of v, it is useful to examine two cases, the first with gy = uy € G, and the second with gy = v) €Y (but not in G). For a transformation parametrized by up, we have V > Ugd = UpbUp ‘Uy = v'u'" (24.12) where v' = ugvug! and wu! = up. In terms of the coordinates v, this implies OT — yyehOTy=t (24.13) XXIV. GAUGE INVARIANT MODELS. 195 and we see that the v transform linearly under elements ug € G. In con- trast, for a transformation vo, we have v > vv Sv’. (24.14) Taking the hermitian conjugate, we find v > ve = uty’. (24.15) Combining the two expressions, we see that v'? = v9070 9. (24.16) In terms of the coordinates v, this implies POTS = yoerhTy (24.17) a manifestly nonlinear transformation law. Note that the v can be trans- formed to zero if we take vg = e3%°T, For infinitesimal variations, the transformations (24.13) and (24.17) can be combined to give d exp(vbT) = -5 [uh + ivf? ]T exp(XT) + 5 expo Ts? — io] = Riek eHM 4. jt, (24.18) where we have set ¢ = &T, with < = 4(u — iv). If we identify « with the lowest component of a chiral superfield A®, and v with the lowest component of vector superfield V, we see that the transforma- tion (24.18) is precisely the lowest component of the gauge transforma- tion (7.15): de” = —iAte” + ieVMA. (24.19) Comparing (24.13) and (24.17) with (E.17) and (E.23), we see that the transformation law of a vector superfield is just a nonlinear realization, corresponding to the coset Y/G. As in Appendix F, we can exploit this fact to construct a fully gauge invariant theory. Recall that previously we found 6g =- J @20d20(A® — A*®)D®, (24.20) 196 XXIV. GAUGE INVARIANT MODELS To cancel this variation, we need to find a function I(,6*/,V) such that oP = i[A® — At ]D®, (24.21) Then L= f d?6.d?6[K(@,0"/) + Ti*1,V)] + [ f 26 P(@!) + ne | (24.22) will be a fully gauge invariant action. To find the counterterm I, we first restrict to its lowest component T(a‘,a*iv), We then write the variation dT in terms of differential operators, OF = OXOT + AOXHOT 4 Fy or dv 1 2 (9 + XO POP + Zee — eX NOOL, (24.23) where P and © include the variations of the coordinates a‘ and a*’/, as well as the appropriate variations of the v. For (24.23) to agree with (24.21), we must demand POT =0 OT = 21D”. (24.24) Furthermore, we also require that I satisfy the boundary condition T(aia*i,0) = 0. (24.25) With these ingredients, it is not hard to integrate (24.24). Following the steps of Appendix F, we find error 1 rao vOD® $009 f doer yO 0), (24.26) XXIV. GAUGE INVARIANT MODELS 197 In this expression, the operator O is the same as © but without the variations of the v: O® = X@ — xx, (24.27) It is a useful exercise to check that T indeed obeys (24.24). Having found the counterterm I’, we are now ready to write the gauge invariant action in superspace. We first promote T to a superfield, re- placing a’, a*/ and v” by superfields ®', b*/ and V. In a symbolic nota- tion, we have T(@iotiy) = f da 2X0 YOp, (24.28) where the differentiations O are performed before the fields are replaced by superfields. Substituting this expression into (24.22), we obtain the com- plete gauge invariant action in superspace: L = [aa OK WO") + [da fd?0d79 er rOmyorpH + [ f 20 P(®!) + ne } (24.29) The action (24.29) is manifestly supersymmetric because it is written in superspace form. By construction, it is also invariant under the local isometries in G: 50! = AMX!) beY = —iAT TMM" 4 ieVAOTO, (24.30) Note that the explicit appearance of the Killing potentials in (24.29) implies that their global existence is necessary for gauging of the isometry group G. To write this action in components, we add the kinetic term for the vector multiplet, lop, , f= gage fa Tr WW + he., (24.31) 198 XXIV. GAUGE INVARIANT MODELS and then pass to the WZ gauge. It is a straightforward exercise to eliminate the auxiliary fields and cast the remaining terms into geometrical form. We find i ; x 1, L = ~Gip DAG A® — i 0"G, 10 — 3 gD i; ; 1 = igin{io™ Dy — i F yy OF + GV 2GiplXOZT $ XOYVO] 1 1 “a 735 DD ,Pyiy) — 37 PieDpPAZ e — gi" D,PDp»P* 1 +] Rime LOK (24.32) where Dy, Ai = dA’ = gdp ®X* axia Pee ax Duh! = Opi! + VpPnA li — gem a eh Qj) = 9 J — gfarey 70) Dyb OmA' Gf Og OH oP DP = ai 2 D,DjP = oP 1, OP (24.33) dAidAi 7 AAP? and we have rescaled V > 2gV. The action (24.32) is invariant under the following gauge transformations: 6Ai = eX exi@ by = _ yi ee Gai * SA = faregr yo 5p = gmt Ane + fe, (24.34) XXIV. GAUGE INVARIANT MODELS 199 The covariant derivatives (24.33) are fully gauge covariant, as is evident from the transformations (24.34). The Lagrangian (24.32) includes the following scalar potential: 1 ¥ = 59D + gD PDpP* (24.35) The first term is the sigma-model generalization of the “D-term” introduced in Chapter VII. Equation (24.35) implies that supersymmetry is sponta- neously broken if either <(D) #4 0 or = —f*™D, (24.6) 5K = OPO 4 ckOpH) — (2 — #1), (24.8) VOTO _ y pOTO AT, (24.13) POT = yi QvOTy (24.17) T= eo — vO po TyOOe 500! = fp daemon, (24.26) 0 = XO — x*o, (24.27) jzaV OC) 17(b) py(b) £ = [@0aOK Wo) + [ da foe + [ f 20 P(®’) + ne} (24.29) 50! = AMXMH/) be” = —iAT@TM” + ieVAPT. (24.30) 202 XXIV. GAUGE INVARIANT MODELS L = —GipDyA' DAY — i20"G, J — , gDe? 2 = gino" DyZi — — iF (Rmn(a) + an + XO AO} 1 ot = 5 PD,PHd — 5 DpDp PT — g"DPD»P* 1 tee + Riek CEE (24.32) Dy Al = AA’ — Gvy(X ita) ; é Dh = Oni! + Vx DaA Ls — G0nl? — x yh = Oy A = of 0,270 oP Ai ap oP D,D,P = = — TS OA'0AT OAK (24.33) bAi = MX oxi” ee aal ~xi 5A = forego Sv_ = G7 Oye + fv, (24.34) a i 1 i” ¥ = 59D? + gDPDpPr. (24.35) EXERCISES (1) Prove that the Killing potentials can always be chosen to satisfy (24.6). This can be done by first differentiating the left-hand side with re- spect to a’, and then using the relations introduced above to obtain XXIV. GAUGE INVARIANT MODELS 203 the a‘ derivative of the right-hand side of (24.6). The proof can be completed by repeating the procedure, this time differentiating with respect to a*‘, (2) Show that xi o D® 4 x*i®) a D® = 0. ca ca (3) Verify that the differential operators, é @) — _jgi Tok XM = —iai TO, — ca ; 6 (a) igti Ts X* = ja*/T' k Bgak? are indeed Killing vectors, where the commutation relations of the T* are given in (7.14). Show that their Lie brackets close into (24.5). (4) Let @ be the complex plane. In this exercise we will gauge translations in the y-direction on ./@. (Note that one could have chosen to gauge translations in the x-direction, but because of (24.6), one cannot gauge both simultaneously.) As above, take K = a*a + d,s0 gage = Land Rogeaqe = 0. For D take the Killing potential D = m(a + a*). Find the Lagrangian and the mass spectrum in unitary gauge. (5) Show that (24.29) reduces to (24.32) in the WZ gauge. XXV. GAUGE INVARIANT SUPERGRAVITY MODELS Having discussed the geometrical interpretation of supersymmetric theo- ries, we are now ready to write down the general coupling of matter fields to supergravity. The Lagrangian we derive is the starting point for the phenomenological study of supergravity theories. We present the Lagrangian in superspace (25.1), in two-component spinor notation (25.12), and as a service to the reader, in a more conventional form with four-component spinors (25.24). Readers interested only in the results should feel free to skip to the relevant part of the chapter. The supergravity extension of the gauge invariant superspace La- grangian is easy to find using the material from the previous chapters. As in (24.17), one first adds the counterterm I to the Kahler potential K. Then, as in (23.1), one exponentiates the result to find v= [ror li (DD — 8R) exp —3 [Kio + rent 1 + agi Han) WOW + 0) +he, (25.1) where x? = 1, and W, = WOT = —_ (GD — 8Rye "De" (25.2) 1 4 is the curved-space generalization of the supersymmetric Yang-Mills field strength. In this expression, K is an arbitrary hermitian function of the superfields © and ®*/, P is the superpotential, and T is the counterterm (24.22), which is necessary for gauge invariance, as we will see below. The analytic function H(q) is included for generality. Under a gauge transfor- mation, it must transform as required to render (25.1) invariant. In what follows, we shall take Hiay, = 5,,; the Lagrangian with nontrivial Hay) is presented in Appendix G. The supergravity invariance of (25.1) is manifest because of the super- space formalism. The gauge invariance, however, is a little more subtle. XXV. GAUGE INVARIANT SUPERGRAVITY MODELS 205 To check it, let us recall the gauge transformations for K, T, and P, 6K = A%FO 4 A*OPFHO _ TAM — A*@]DO oF = i[A® — A*@]D bP = AMXp, (25.3) as given in Chapter XXIV. Here FO = XK + iD (25.4) is an analytic function of the ', and A is the superfield gauge param- eter. When applied to the Lagrangian (25.1), the transformations (25.3) induce a variation of the following form: 1 (DF — 8RAPFO + APOP HO. “KEN bL = foo 26) 5 + nox | + he, (25.5) In Chapter XXIII, such a variation is canceled by a super-Weyl transformation, where the Weyl weight of ®' is taken to be zero. Setting the weight of V to be zero as well, we find 3 - , bf = foo 26 E (PP — BRE + Bee + oxp| + he. (25.6) under a super-Weyl transformation with superfield parameter 2. Com- paring (25.5) to (25.6), we see that the variation is canceled if 1 = — AMF@ zx é AF 5P = A@XP = —APFOP, (25.7) The condition on 6P is a nontrivial condition on the superpotential that is necessary for the gauge invariance of the theory. The superspace Lagrangian presented above can be expressed in com- ponents using the techniques developed in the previous chapters. One first 206 XXV. GAUGE INVARIANT SUPERGRAVITY MODELS. passes to the WZ gauge, where 1 i T= vop@ + 5 Gi XOXO YOYO and 1 -- 1 W, = ~4 (99 - ee) fav -35 [var]. One then works out the © expansions for (QZ — eR exp] 3 [K+ ri} and wow, (25.8) (25.9) (25.10) (25.11) After eliminating the auxiliary fields, and rescaling and redefining the other fields as in Chapters XXJ and XXIII, one finds the component Lagrangian in terms of the physical fields. This Lagrangian is the starting point for phenomenological studies of supergravity theories: 1 ~ Aig Ls a2 L = 5h — eg DyA'GMA® — 5 eg°D® _ jh ma _ ieXG"G,, A — igi O"D yt + eT Dy + V2egg ip X*O 71 + V2e9g;-X OPI 1 x 1 7 _ ee + 5 gD YA lis Q Awiyinm=n Lon B gizi=m ny = 7 V2Gi¢PnA oO" Wm — 3 V2» DnA LE" o'W + ; CLVino?o"Z +f, FPE™ VF + Fy] Van a a + capi oem + Vm WC Ond! XXV. GAUGE INVARIANT SUPERGRAVITY MODELS 207 1 kei ~ gy lGirdise — Riper lL 1 a + BIL OLR gh — eA g™ZOZO GZ) — e exp(K/2) {prc + Pye, + 5 VDD Prot, + 5 V2D.P Aza", 1 1 , +5 UDP! + 5 a.b.Pern — eexp(K)[g'"(D;P\(D,P)* — 3P*P]. (25.12) In this expression, the scalars A‘ and the spinors y' and 4 are matter fields, while the vectors v,, are the gauge fields for the gauge group G. The field ,, is the gravitino, and e,,“ is the graviton. In (25.12), K, P, and D are functions of the scalar fields. As before, the metric g;;. is Kahler. The Lagrangian (25.12) contains derivatives covariant with respect to gauge transformations, as well as spacetime and Kihler coordinate transformations: Qt » I A! = Gd, X40) -, : - axto Dy = Opti + LOm + Ty uAYk ~ gon! St _i 4 Dyh = Oy A + BD — Gf PV, OO (KDA! — Kp ZAP yl — 5 guy{0 Im Fy! + 1 (KOA! _ K.P A* A + 594 Im FOZ Drilin = Onn + Wom + q(KiZmA! = KpD,A* iy + 5904 Im Fp, D,P = P, + K,P Q,D,P = Pi + KiP + K,D)P + KjD,;P — K,K,P —ED,P. (25.13) 208 XXV. GAUGE INVARIANT SUPERGRAVITY MODELS The covariant derivatives contain the Christoffel symbols for the Kahler geometry and the spin connection (17.12) for spacetime. They also contain the vector potential v,,. Note that the covariant derivatives contain a coupling between Im F and the vector potential. This is a reflection of the fact that gauge transformations are accompanied by super-Weyl rota- tions of the component fields. The above Lagrangian is invariant under the gauge group G. The gauge transformations of the component fields are given by 6A = gOXI@ 7 oxi i byi = i gt 5a" Im Fz ; s ; i 5 SAD = fare — 58” Im FO bv, = ga d,e + fev, © On i (a) *(a), 2 Im Py. (25.14) It is automatically invariant under supergravity transformations because it was derived from a superspace formalism. It is instructive, however, to verify the invariance directly, using the following transformations laws, Seen" = COV m + CFV n) 5A = 26! bef! = if20"CD,,A — Ti, 5AM 1 ni I> KI. ijt + (KSA! — Ky d AM)! — V2eK? gi" D wP*l SP = (C62 + CEA) a 1 s yay] 5A = Fy Mol — 4 (KoA — Kp d,A*)A — igDOC 5m = 2G — 5 Onl GLO" + 5 (Grn + make AOA 1 g(Kj5cA! — Ky 5AM hy + 08? Poyk. (25.15) XXV. GAUGE INVARIANT SUPERGRAVITY MODELS 209 Here J,¢ is defined to be Dh = Ok + CO + [KGaA —KpFpA*NC, (25.16) while the supercovariant expressions B,,Ai and F,, are given by B,,Ai = By, A! — gt, 9X = TyAi — 53 Yi F, @ = 5,0, — Byvp,! _ Foyy(® _ 5 dwell + VinF AO = Wb yA = VFA]. (25.17) The action (25.12) differs from that of Chapter XXIII by the addition of the gauge supermultiplets. The additional fields change the form of the scalar potential from (23.10) to 1 ays V = a + eX[g'"(D,P)\(D;P)* — 3P*P]. (25.18) They also change the trace formula from (23.34) to Str M2 = Yo (—1)%(2J + 1) Tr M? spins J = (n= N[2my — g?¢D?y] + 292g" DQD> — 2¢RieG'G" mj. (25.19) From the form of the transformation laws, we see the condition for spontaneous supersymmetry breaking is either #0 (25.20) or

) #0, (25.21) 210 XXV. GAUGE INVARIANT SUPERGRAVITY MODELS for some value of i or (a). Depending on the relative magnitudes of (25.20) and (25.21), an appropriate linear combination of x! and 4 plays the role of the Goldstone fermion. Note that the Lagrangian (25.12) explicitly contains the Killing poten- tials D. Their existence is both necessary and sufficient to gauge the group G. If the group G contains a U(1) factor, we know from previous arguments that the D® are not uniquely defined. There is an arbitrary integration constant associated with each U(1) factor, DoD+é. (25.22) In the globally supersymmetric case, shifts of these constants give rise to the Fayet-Iliopoulos D-term £,,. The same is true in supergravity. By shifting the functions D, we find the gauge invariant supergravity version of Ley Duy = —Fegre? ~ eg D ~ 5045 no™T — Gn”). (25.23) Note that the shift (25.22) changes the spinor covariant derivatives as well as the transformation laws (25.15). New terms proportional to € are in- duced in all expressions involving the Killing potentials D. In the rest of this chapter, we will present the Lagrangian (25.12) in four-component notation, following the conventions described in Appen- dix A. Care should be exercised in comparing this formula to those in the references; conventions vary throughout the literature. With this said, we write the Lagrangian as follows: 1 1 g= 3 eh — CG jp D yA'DM AN) — zg De? _ fF anP — icH yD, A = iegipTiy" Dug, + eS yD niin + egy 2GipX HOTA + egy 2GieX TIO 1 7 5 1 7 ; + 5 eqD ini"Zf — al 1 7 ~ 5 200: ¢ LA TRI Wim — 52241 F A'Zhy" "em XXV. GAUGE INVARIANT SUPERGRAVITY MODELS, 211 i. - - + Zim AD + VamoyAR Fao + Fes] 1 . 7 r = + ipl tim — Von Wile niabkk 1 Zi yk oj yt — gL Deer — Rijuee RTL 1 ] (a). AO 4 3 ) (a)aym (a) 7 (b), (b) 3 CG ie LLY LAR mA + IG A AR mA — eexp(K/2) {Pinan + PUT ey ip i “ia +5 V2DP Ties + 5 V2 oP Tin 2 — eexp(K)[g'"(D,P)(D;P)* — 3P*P], (25.24) 1 i, 1 ; + 5 DDjiPL RL + 3 9D PF} where yi. = 4(1 + 7s)z', and similarly for 2 and y,,. The covariant derivatives are defined as follows: Q mA = d,Ai = guy Xi (a) Drift = Onil, ~ Ont + Ve PmAlth, — Bn th 1 ~ a ~ F i i = GK iD! — Kp DyA* Yh, — 5 Gml? tn FY, Dyh = Og — Oy A — gf Pv, PAO? Io. + (KiOndl — Ky AAO 4 5 om © Im FOAM Fon = Ont in — On in 1 ~ ~ i + G(KiGnA! — Kp DyA* Wan + 5 9n'? Im FO, D;,P = P; + K;P Q,D;P = Py + KyP + KD,P + K,D,P — K:KjP —TSD,P. (25.25) 212 XXV. GAUGE INVARIANT SUPERGRAVITY MODELS The Lagrangian (25.24) is invariant under the supergravity transforma- tions, 8: = UCL im + Ferm) 6A = Vert 5e7i, = VIM DpAi — TH bel 1 . : x va + 4 (KjdAl — Kp dA), — 2K? gi*D.P*C, Benl™ = CLAD + Catrall < a 1 . 5AY = Fyol, — ik j6AF — Kp 5, A*)A — igDC, < >, i, i, i + Fla).nj bem = 2b — 5 Smnbe ie" te + 5 (Gun + Smnder ARIA 1 ; — 4K A! — Kj dA Wim + 108? Pinles (25.26) where %,,¢, is given by 1 ~ {KiGud! — Kp GyANy. (25.27) SQ mbt = Ob, — Omer + REFERENCES J. A. Bagger, Nucl. Phys. B211, 302 (1983). E. Cremmer, S. Ferrara, L. Girardello, and A. van Proeyen, Nucl. Phys. B212, 413 (1983). EQUATIONS L= Jeore| a9 — 8R) exp |—3 [K(®,0*) + reo + Tigi Hoos + | + he. (25.1) 1. W, = WET = — (GD — BR" D0". (25.2) ~ ~ 0. yp = Opit + 10m + Vip uAiGe = Gq!” XXV. GAUGE INVARIANT SUPERGRAVITY MODELS 213 1 = yigmani L = ~Feh ~ eG. FyA'G"A¥) — 5 eg? DO m _ 5PM = ieTG"G,, 4 = eg iO"D yy + 8! E,D yn + Begg X12 + Pegg, pi ZT® 1 = 1 7 - 7 gD oO + 7 gD 7 ~ I V2e5-P AML 0" — 5205p /A'LO"OTy Nie + Lg + yao IEF + Pel] 1 zo kémn, Ip nym) jin Zi + egiplie Wom + Uno WV Gnd 1 i ~ glindeer — Ripnel ALT ! zigmyiT@e p(a\ 3 F(a) 1(b)- F(b + 3 esinrke ROE AO — 16 ea} 1G, A” — e exp(K/2) {Pvc + PUT in ig), i />, a Fi =a, + yD Pro. + 5 V2DHPTLO 1 id , + 5 GD Pri + janbsprzeh — e exp(K)[g'"(D;P)\(D,P)* — 3P*P]. (25.12) Ai = 6,At = gvy(®X i(a) xi Loe ay = pene tl ; ~ {KG mA! = Kp Fp 5 In” Im Fy 214 XXV. GAUGE INVARIANT SUPERGRAVITY MODELS Bh = Pgh ® + 2g ~ Gf BV ORO 1 + GK Dud! — Ky FAO + 5 ota Im F®2 Daly = Onlin + Val tn ~ i + G(KPmd! — KFA" Wy + 500 Im Foy, D,P = P, + KP DD,P = Pi; + KP + K;D;,P + KD;P — K;K;P —T%D,P. (25.13) bAi= OX i(a) : 6x’ 7 : byi= Fi yit 5" Im Foy SAM = force JO — 5 Im FO1@ Sp = gran + fareeHy, Sy 5 2 Im FY,. (25.14) “= ilo, + Ty) = V2, i = iV30"CB,,A° — Ti, 6, Aly* 1 ; ap ie + (Kydd! — Kp 5A) — 208? gi* DP*C 8m = H(L,A + yh) a 1 . . 62 = Py ol — i (Kj6,A) — Kp 5-A®)A — igDl 1 ~ 7 = Balm = 2b — 5 Sak IrLO"E + 5 (Gn + Orde AOA 2 _ aK 5;AI — Kj 5A" thm + ick? Pon’. (25.15) XXV. GAUGE INVARIANT SUPERGRAVITY MODELS 215 __t i ¥ = 59D + eX[g'"(D,P\D;PY* — 3P¥P). (25.18) StrM? = Yo (—1)*(2J + t)Tr M? spins J = (n — 12m — g?] + 297g" DYED — 2(Rj-G'G" mj. (25.19) EXERCISES (1) Show that — 8R)(e~"F,e") is gauge covariant under the following non-Abelian gauge transfor- mation: V gWih* pV oid | where DA = G,A* =0. (2) Verify that - 1 e"De" =DV — al. DV in the WZ gauge. (3) For an Abelian group, the components of W were given in (19.28). Use the results of Exercise 7 in Chapter XIX to show (G,D* — 8RVG,V]| = GF — 8RV.P,V]| = DP D(D,D* — &R)V.PV | = 0 8 is DG (oe) pa¥aPs —16i[ v4") — 8ifo% va]. Then find all the components of W for a non-Abelian group G. 216 XXV. GAUGE INVARIANT SUPERGRAVITY MODELS (4) Compute (G,G* — 8R\B*V = 2fiA*2A + oS v,) — /20°%Fv,} + 00 {12/274 + 4iD.A** + A*L=2D + 2ie"D,yv° — 4 ab" + VG — IEW, — Woh wd} and (G,D% — 8Ry®*V? = 200 A*0,v". (5) Check that (9,2* — 8R)@* V3 = 0 in the WZ gauge. XXVI. LOW-ENERGY THEOREMS In the study of chiral dynamics, nonlinear realizations of chiral sym- metries have proven to be useful tools for constructing low-energy effec- tive Lagrangians. In this chapter we shall see that similar techniques can be used to describe the low-energy effects of spontaneously broken supersymmetry. The resulting low-energy theorems describe the effective couplings of Goldstone and matter fields at energies far below the scale of the symmetry breaking. The fact that the low-energy theorems hold for supersymmetry might seem surprising, for the usual proofs in chiral dynamics rely on the finite volume of a compact group. For the case of supersymmetry, the anti- commuting nature of the group parameters makes such volumes vanish. Nevertheless, we shall see that alternative proofs can be supplied which validate the supersymmetric versions of the low-energy theorems. In chiral dynamics, the low-energy theorems apply when a group G is spontaneously broken to a subgroup H. The subgroup H is linearly represented on the physical fields, while the remaining generators of G are realized nonlinearly in terms of the coset parameters for G/H. The coset parameters can be interpreted as Goldstone bosons associated with the spontaneous breaking of G down to H. The nonlinear realizations of G are determined up to field redefinitions. They are often parametrized in certain canonical forms known as stan- dard realizations. These realizations linearize on the subgroup H. Any linear representation of H can be promoted to a standard realization of G. Conversely, any realization of G that linearizes on H can be decom- posed into a set standard realizations and Goldstone fields. For the case of supersymmetry, the Lorentz group plays the role of the subgroup H. The remaining generators generate pure supersymmetry transformations. In Chapter XI we used this construction to find a non- linear realization for the Goldstone fermion /, Bedale) = © E, ~ ieBla9 Oat) Blo) = © Ey ~ ieBx) Ol) v(x) = K[A(xo™ — Co™A(x)], (26.1) 218 XXVI. LOW-ENERGY THEOREMS where x is a constant that parametrizes the supersymmetry breaking scale, analogous to f, in chiral dynamics. These transformations can be lifted to superfield form using the techniques introduced in Chapter IV. The relevant construction is given in (4.11); for the case at hand, it gives a superfield A whose lowest component is the Goldstino 2: A,(x,0,0) = exp(0Q + 90) x A,(x) A,(x,0,0) = exp(0Q + 00) x 7,{x). (26.2) The superfield A is built out of A, its derivatives, and the constant x: A,{x,0,0) = 2,(x) + Yo, tee KAx,0,0) = T(x) + 8, + 0°. (26.3) I K It is a short exercise to show that the transformations (4.10) reduce to (26.1) when applied to the lowest component of A. The Goldstone superfield A can also be defined as the solution to a certain set of constraints. These conditions can be found with the help of the identity, D, exp(0Q + 00) x = exp(0Q + 00)Q, x D, exp(0Q + 6Q) x = exp(0Q + 6Q)Q, x. (26.4) Applying (26.4) to (26.2), and using (26.1), we find 1 omaha DA, = Paz + ixoggA* 0,,A, DjAg = —ixAPogp” Ona. (26.5) These constraints are consistent with the D algebra (4.7). Their solution is the superfield A as defined in (26.2). To derive the low-energy theorems, we need the supersymmetric ana- logs of standard realizations. We shall define a standard realization of supersymmetry to have the following transformation law: é Befle) = —i0800) 5-5 F109), (26.6) XXVI. LOW-ENERGY THEOREMS 219 where v? is given in (26.1). In the exercises, you will show that (26.6) closes into the supersymmetry algebra. The field f is free to carry an arbitrary set of Lorentz or internal symmetry indices. As with the Goldstone fermion 2, we would like to promote f to a superfield F whose variation reduces to (26.6) when restricted to its lowest component. Using the construction of Chapter IV, we find F(x,0,0) = exp(0Q + 00) x f(x) ; = Fle) — i880) S109) +. 26.7 In (26.7), the superfield F carries the same indices as f. Its component fields are built out of A, f, and their derivatives. It is also possible to derive (26.7) from the constraint equations, D,F = ix(o"A), OnE D,F = —ik(Ao"), OnF (26.8) In the case of chiral dynamics, it is well known how to convert any non- linear realization into a standard realization. As shown in Appendix E, one simply applies a finite group transformation with the field-dependent parameter that would transform the Goldstone fields to zero. This pro- cedure also works for supersymmetry. To see this, let f be an arbitrary nonlinear realization of supersymmetry, and let F be its superfield exten- sion. A standard realization F’ is obtained by taking F'(x,0,0,2) = e&*52F(x,0,0) a (26.9) where the Q’s are the differential operators (4.4), and the substitution € = —x/is made after all the differentiations are performed. We can also write (26.9) in a more explicit form, avoiding derivatives on A, by changing arguments as follows: - 3 5 F(x,0,0,4) = exp [an | exp] —x(200 ant iw) x F(x,0.0)|,-)- (26.10) In this expression, we are able to separate the exponents because of the fact that =0. (26.11) 220 XXVI. LOW-ENERGY THEOREMS, To show that f" is a standard realization, we must compute the change in F’ from a supersymmetry transformation. This is most easily done using (26.10). The variation of vj' follows from (26.1): Sevp(y) = 609 — Oa"E — Wk) 5 ot): (26.12) Using (26.1), (26.12), and (4.4), we can then compute the change in F’, 5F(x.0,0,4) = {[ cord — 00m = — ieh(y) J] ene» ai 6 é x. 6\|Jx O2 x exp] ~ «(20 Dat Ay) | F(x,0,0) 6 a 7 6 + experi zoe [{] -£ 59 — op - MOV | @ 5.8 _ A é é + exp] ito =| exp] -* (A) 25 ot Ay) 5 >) emt ER! itEo"B — 00") = sey = ~ite 2 F'(x,0,0,A). (26.13) Taking the lowest component, we see that f’ indeed transform as a stan- dard realization, é Sef (XA) = —iv(x) Zea TA) (26.14) As above, the fields f’ and F’ can carry any Lorentz or internal sym- metry indices. With these results, we are now in a position to supersymmetrize any Lorentz invariant Lagrangian. The first step is to find a Lagrangian for the Goldstone spinor A. Two obvious choices are 2 Ly = 7 feow ann (26.15) XXVI. LOW-ENERGY THEOREMS 221 and L,= 5 faroaan +). (26.16) It is not hard to show that the highest component of A? + A? is a total spacetime derivative, so (26.16) is unsuitable for a supersymmetric action. In contrast, (26.15) is perfectly fine, and coincides with (11.11) when ex- panded in terms of component fields. We shall take it to be the Lagrangian for the Goldstone fermion. The next step is to construct the matter superfields. We start with the original matter fields, which have well-defined transformations with re- spect to the Lorentz and internal symmetry groups. We assign the fields supersymmetry transformations via (26.6), and promote them to superfields via (26.7). In this way we build a superfield out of each matter field in the original theory. The final step is to construct the supersymmetric matter coupling. We start with the original Lagrangian Y, and replace all the matter fields by their corresponding superfields. This gives a superfield Lagrangian whose lowest component is the original Lagrangian. We then turn this lowest component into a highest component by multiplying the superfield ex- pression by A?A?, < 1 ~ NR? = ah font. (26.17) This gives a fully supersymmetric Lagrangian, gant f @00202K2F, (26.18) whose /-independent part is just the original Lagrangian L. As usual in the theory of nonlinear realizations, it is always possible to include higher-derivative terms in the effective action. For example, a con- tribution of the form Lr= [@0a20(D2A2)(D7A?) ~ (Aq AG™ OA GAG" 8,2) + +>* (26.19) adds a higher-derivative interaction to #. The coefficients of such terms are not determined by symmetry, and must be regarded as parameters of the theory. The leading term in the derivative expansion is the only term 222 XXVI. LOW-ENERGY THEOREMS that is unique. At high energies, where the higher-order terms become important, the predictive power breaks down. The Lagrangian (26.18) describes the low-energy interactions in a theory where supersymmetry is spontaneously broken at some scale much greater than the energies involved in the low-energy effective theory. For example, the formalism would apply to the situation where all the supersymmetric partners of the physical fields are very heavy (except for the Goldstino). In this case, the low-energy scattering amplitudes are determined by the effective theory. The only signals of supersymmetry are the nonlinear cou- plings of the Goldstino to the physical fields. To illustrate this construction, let us consider the case of a free scalar field a(x) and a free spinor field (x), We supersymmetrize the Lagrangian by assigning transformations to a and w via (26.6), and lifting them to superfields A and ¥ that satisfy the constraints (26.8). We then replace the fields in (26.20) by A and ¥, to find the superfield Lagrangian Y, a a2x2| 1 1 = 2, 2, 2A2] __ a" 4 — —m2A2 = ty + [aOrInn| 5 oyAcrd zm —iPor d,,P — pHi? + ¥)|. (26.21) The Lagrangian (26.21) should be expanded in terms of the Goldstone spinor 4. A helpful trick is to replace d?0d?0 by D?D?/16 and use the con- straints (26.5) and (26.8) to compute the D and D derivatives. To second order in /, the resulting Lagrangian is of the form, < i -, & L = 5 — iO Oyh + “5 107 Oy P K 2x? + = (468° — 8"20"2)T yy +o (26.22) At low energies, the Goldstino couples to the energy-momentum tensor T nn, independent of the details of the symmetry breaking. This is the low- energy theorem for supersymmetry. XXVI. LOW-ENERGY THEOREMS 223 REFERENCES E. A. Ivanov and A. A. Kapustnikov, J. Phys. Al1, 2375 (1978). S. Samuel and J. Wess, Nucl. Phys. B22, 153 (1983). EQUATIONS cals) = EE — iaPl0) Oya) 5 A(x) = ae — WWE (x) bypAsl) v(x) = K[A(x)o™E — EoA(x)]. (26.1) DyA, = te + ino,j"A? A, Ny K DjA, = —ixNPo 4)" OAs. (26.5) . 6 Oeflx) = — 10) Ba fO9- (26.6) x’ D,F = ik(o"A), OmF D,F = —ix(Ao"),0,F. (26.8) 2 Lo= —* [arod0nrne. (26.15) 2 Geax froaonrarZ. (26.18) 1 - ives =. — ido’, m9 g L ye to™ 6A + 2 26"2 OmL + so" aT — O00") Ty, Hoo. (26.22) EXERCISES (1) Show that (26.2) satisfies the constraints (26.5). (2) Check that the transformation law (26.6) for a standard realization closes into the supersymmetry algebra. 224 XXVI. LOW-ENERGY THEOREMS (3) Verify that (26.7) is a solution to the constraints (26.8). (4) Show that (26.15) coincides with (11.11) when expanded in terms of component fields. (5) The Lagrangian (26.21) is supersymmetric because the derivative of a superfield is still a superfield. However, the derivative of a standard realization is not a standard realization. Use the techniques intro- duced here and in Appendix E to find a “covariant derivative” A that preserves the transformation properties of a standard realiza- tion. The Lagrangian ~ 57) 1 1 = 20 J2h A2A2| —— m4 © 2 42 ¢ = feoronn| 5 Awl A"A ue — 1 7 NOMA — 5 (P+ ¥)| is another possible extension of (26.20). It differs from (26.21) by higher-order terms in J. The derivative A is natural to use when gauging an internal symmetry if the vector superfields belong to a standard realization. (6) Show that (26.21) reduces to (26.22) in terms of component fields. APPENDIX A NOTATION AND SPINOR ALGEBRA We use the metric 7, ~ (—1,1,1,1) throughout these lectures. Further- more, we work with Weyl spinors in the Van der Waerden notation. To begin, we define M to be a two-by-two matrix of determinant one: M €SL(2,C). The matrix M, its complex conjugate M*, its transpose inverse (M*)~+, and its hermitian conjugate inverse (M*)~' all represent SL(2,C). They represent the action of the Lorentz group on two-com- ponent Weyl spinors. Two-component spinors with upper or lower dotted or undotted indices transform as follows under M: We = Mp Us = M* ED, (Al) yt = Mo ey? v= (M1594, oo Spinors are denoted by Greek indices. Those with dotted indices trans- form under the (0,3) representation of the Lorentz group, while those with undotted indices transform under the ($,0) conjugate representation. The connection between SL(2,C) and the Lorentz group is established through the o-matrices -1 0 01 0 te ° -( 0 ) ° -(; 0) 0-7 1 0 2 3 ae ° -(; 0) ° ( i) in complete analogy to the relation between SU(2) and the rotation group. These matrices form a basis for two-by-two complex matrices: (A.2) —Pot Ps Py — iP P = P,o" = mo ( : (A3) P,+iP, —P)— Py; Any hermitian matrix may be expanded with the P,, real. 226 APPENDIX A From any hermitian matrix P, we may always obtain another by the following transformation: P' = MPM*. (A.4) Both P and P’ have expansions in o, o"P, = Mo™P,,M*. (A.5) Since M is unimodular (det M = 1), the coefficients P,, and P,, are connected by a Lorentz transformation: det[o"P',] = det[o"P,,] = P2 — P? = P? — P?. (A6) Vectors and tensors are distinguished from spinors by their Latin indices. From (A.1) and (A.5), we see that o” has the following index structure: oi". (A.7) With these conventions, wW,, W,%, and w*o,;"0,,/* are all Lorentz scalars. Since M is unimodular, the antisymmetric tensors ¢”* and c,, (¢2, = e'? = 1,6,, = 6?! = —1,6,, = &22 = 0) are invariant under Lorentz transformations: fap = M,'My°t,5 (A.8) ef = 6M AMI. Spinors with upper and lower indices are related through the ¢-tensor: Wey Wy = ap (A9) Note that we have defined ¢,, and ¢*? such that ¢,,c"” = 6,’. An analogous treatment holds for the ¢-tensor with dotted indices. The e-tensor may also be used to raise the indices of the o-matrices: mix — pith yaBgm ap: (A.10) From the definition of the c-matrices, we find (o"a" + oa"),P = —2n™ 5, A (A.11) (a"o" + ao" = — 20%, APPENDIX A 227 as well as the following completeness relations: Troms" = —2y" 4;"F py!" = —25,P5,!. 1) These relations may be used to convert a vector to a bispinor and vice versa: mia, 1 Ugg =F yi" qe UO = — BV. (A.13) at The generators of the Lorentz group in the spinor representation are given by 1 - are — gg amt = 4 (0,:'0" — 0,,"0") (A.14) 1 ams = qos" = o"o,;"). Other useful relations involving the o-matrices are 0a — aoa = —2ie™“G, ; (A.15) o'G'a® ~ oOo" = 2ie”*“4o, where &,;.3 = —1, as well as abc cab aa cob be ga ab a 60 + 66’ a" = Ano” — na" — na") ; : « ; (A.16) oS + Kors = Ano? = ot = ns), and G43 F ph" — Oya" Oph" = 2[(0""e) p6 3p + (co) p66] (A.17) O43" ph" + Fax" pg" = —NEgglag + 408) gl CF) a The equations (A.11) make it easy to relate two-component to four- component spinors. This is done through the following realization of the Dirac )-matrices: 0 oa” = ( 0 } (A.18) 228 APPENDIX A We shall call this the Wey! basis. In this basis, Dirac spinors contain two Weyl spinors, ha wp, =(%), (A.19) -) while Majorana spinors contain only one: Py = (%): (A.20) Throughout these lectures we shall use the following spinor summation convention: WL =v v7 = 0. = hil = LW, = X= —WIs = 1" = 1. (A.21) Here we have assumed, as always, that spinors anticommute. The defini- tion of #7 is chosen in such a way that (A.22) (ah* = h,)* = Note that conjugation reverses the order of the spinors. REFERENCES E. M. Corson, Introduction to Tensors, Spinors and Relativistic Wave Equations, London, Blackie and Son (1953). W. Thirring, Supplemento del Nuovo Cimento 14, no. 2, 415 (1959). EXERCISES (1) Compute P;, in Eq. (A.5) for M = exp(si¢o3) and M = exp(}y03). (2) Show that M, (M7)! form equivalent representations of SL(2,C). (3) Demonstrate: (4) Verify Eqs. (A.11) and (A.12). APPENDIX A (5) Show: om = 9 a" Peg, = o™ Peps. (6) Verify: eh , = — Diet” etG , = 26”. (7) Demonstrate: in the Wey] basis. (8) Show that the canonical basis for the Dirac )-matrices, 40 = -1 0 ok = 0 of ve o ip Vlg of is related to the Weyl basis (A.18) by the following similarity transformation: 1/1 -1 Dy = XPeX7, x= +( ) y2\ot Also show that the Majorana basis, in which pif = —7'h. 40 _( 9 —o wt [9 io* iM =\_ oo 0 ™M = \igs 9 42 =(; 0 ya (/ 0 ~io? Hea (9 Lj vh (ins 0) is related to the Weyl basis by a further similarity transformation: (Ci): 230 APPENDIX A (9) Let ¥y, denote a Majorana spinor in the Weyl basis, ~-@) Convert this to the Majorana basis of Exercise 8: mm (Rey, Yu > V @). “ (187) (10) Prove the following relations: 70" = 5 e00 0,85 = Fe HF = 5 00 1 8,0 = — 000 60000") = -4 0008n™". (11) Use Exercise 10 to show (0004) = ~5 (oyN00), (O5NDD) = —5 GOVE). (12) Verify: Trom™aM = -} (ymin — nrqt®) — seme, where é9123 = —1. APPENDIX A 231 (13) Rewrite the supersymmetry algebra (I) in terms of four-component Majorana spinors: {0,95} = 27%5Pm [0..Pm] = [QxPn] = 0. APPENDIX B RESULTS IN SPINOR ALGEBRA Conventions: mn ~ (—1,11,1) ese 1, =e =-1, ey =e, =0 £9123 = —1 We = Wg, Wy = bag” WH = We = Wak = Ha = 1 UT = Va! = Vs = Kal* = 1 (nh)* = 0) = Wait = 07 = vo-() ae (29) y= yyy = ( ‘ (B.1) Sigma Matrices: (B.2) a = 0° (B.3) APPENDIX B Tr o"G" = —2n" ai"? = ~25,P 5,8. G, (o" + o"a™),! = —2n5,? (o" + ao"iy = — 2 5%. om! = —(6,,"6"% — o,,manil) 4 1 am, = 5 (0%0,4" — 0," ome = 0 oho py = 0 Pepa. abed, ethetg., = —2io” eed) = 216%. Ga" Opp" — xa O pp" = (Ge) apeag + (EF )eplap] Oxi" Opp" + Fs" Opp" = —N'™Eaplag + 4(0°"8) gpl F™),g. Tromot = — Soren _ n™nt®) 7 sen, o°GPo! + oa" = no" — Hho" — n**o") G0°G + Go's" = 2Ay*G — yi G* — °S*). aaa — ooo" = Lica. 0° — a ora* = —2ie*a, Spinor Algebra: 00" = 5 er00 1 0.5 = 5 #4900 Oe => #00 ate 1 0,05 = —> £5900. 233 (B.4) (B.5) (B.6) (B.7) (B.8) (B9) (B.10) (B.11) (B.12) (B.13) 234 APPENDIX B 00"000"D = -} 608m". (6a) = ~5 (4100) NBD) = —5 (FO NOD). xo"p = —a"y, yor" = Wo"a"y, WM = -5 (po"Z)(wor)s. (B.14) (B.15) (B.16) (B.17) (B.18) (B.19) APPENDIX C KAHLER GEOMETRY The matter couplings of chiral multiplets are conveniently described in the language of Kahler geometry. It is useful, therefore, to introduce the notion of a Kahler manifold. A Kahler manifold is a special type of an- alytic Riemann manifold, subject to certain conditions that we will discuss below. Since the manifold is analytic, it can be parametrized in terms of complex coordinates a' and a*', where i = 1,...,. Under an analytic coordinate transformation, a@=ala) a = a*(a"), (Cl) the differentials and derivatives transform as follows: i oak" da*" = aael da*! 6 da*i 0 Bak = Bat Baal (2) These transformations preserve the analytic nature of the coordinates. They also define the transformations of covariant and contravariant vector fields, ai Oa’ Viaa*’) = Vfaa*) ati Via a*’) = a Vi(a,a*) , Volatsat’) = 5; Vasc) Va a) = Sa; Vasa"). (C3) The first condition on a Kahler manifold is that it be endowed with a hermitian metric gj. The metric must be positive definite and invertible, 236 APPENDIX C which allows us to raise and lower the indices i and j*: I Vi = gi VO Vee = gaye Vi Viz gh Vy VT = gi VY (C.4) The second requirement is that the covariant derivative must respect the analytic structure. This implies that Ij = Tf, = 0, so the covariant derivative is of the following form: Ye - THM 6 ke iV — Ta Me (C.5) The third condition is that the connection be compatible with the hermitian metric. This imposes the additional restriction, Vigiw = 0 Vegi = 0. (C.6) The transformation law for the connection is chosen to assure that covariant derivatives of tensors transform as tensors. This implies Ca" eal 6a" da" da" a ae a mm A atk Th = sar aca ae Vim: (C7) éa*" da’) Ga" The first of the equations (C.7) tells us that it is consistent to set the torsion to zero, leaving only the symmetric part of the connection. ri = Th. (C8) The second equation implies that it is also permissible to demand ré iy (C9) Equations (C.8) and (C.9) are the two remaining postulates that define a Kahler manifold. On a Kahler manifold, the conditions discussed above imply that the only nonvanishing components of the connection are Ij, and its complex APPENDIX C 237 conjugate I’j,,.. Equation (C.6) can be solved to give 0 n= Bai $00" (C.10) Since Th = Tis (C11) the metric must obey the following integrability condition: 7 Iki (C.12) A similar relation holds for the conjugate derivatives, 6 a Saat Gist = Gaal Site: (C.13) Equations (C.12) and (C.13) imply that the metric is the derivative of a scalar function K, (C.14) The function K is called the Kihler potential; its derivatives determine the metric and the connection. Kahler manifolds are often defined through (C.14), in which case the conditions on the connection are then deduced. The Kahler potential completely specifies the Kahler geometry. Note that the metric g;j is invariant under analytic shifts of K, K(a,a*) > K(aa*) + F(a) + F*(a*). (C.15) Such a shift is called a Kahler transformation of the Kahler potential. The curvature of a Kahler manifold can be defined as the commutator of two covariant derivatives: [VViMi = Rin Me [ViVi] = Rou Ve- (C.16) The upper index on R can be lowered with the help of the metric, giving Rijnex = Imex Rin Rajeres = Imee Riou (C.17) 238 APPENDIX C In Exercise 3 we will see that only R,;,,. and its complex conjugate are nonvanishing. From the definition of the covariant derivative, we find = m Ripper TM a0 sa (0 a = Fgh Bgnt See — I Bag Imex J 75 Duo J- (C.18) Using (C.16), (C.17), and (C.18), it is not hard to show that the curvature obeys the following symmetries: Ripuer = —Rijworu = —Ryrince = Ryriere: (C.19) ij This is all the Kahler geometry we need to discuss the general couplings of chiral fields. REFERENCES M. Bordemann, M. Forger, and H. Rémer, Commun. Math. Phys. 102, 605 (1986). K. Itoh, T. Kugo, and H. Kunitomo, Nucl. Phys. B263, 295 (1986). EXERCISES (1) Verify the transformation law (C.7) for the connection I. (2) Impose the Kéhler conditions (C.8) and (C.9), and solve for the con- nection in terms of the metric. (3) Show that Rj». is the only nonvanishing component of the curva- ture on a Kahler manifold, and solve for the curvature in terms of the metric. (4) Compute the curvature, Ricci tensor, and curvature scalar for the manifold with Kahler potential K = —3 log(t — 4a*‘a’). (5) Show that in the language of differential forms, the Kahler condition (C.14) is equivalent to the statement that the fundamental form Q= Sain da‘ da*i is closed, dQ =0. APPENDIX D ISOMETRIES AND KAHLER GEOMETRY In this appendix we will discuss the isometries of Kahler manifolds. The techniques we introduce will prove useful in constructing gauge invariant matter couplings in flat and curved space. Before specializing to Kahler manifolds, however, we first define the general notion of an isometry group. Consider, therefore, an arbitrary differentiable manifold ./, and a set of parametrized curves that fill the manifold without intersecting. Then construct the map ¢,: > . which takes each point p € .W a parameter distance t along the unique curve that passes through p. This map also induces a map on the tangent space. If the induced map leaves the metric invariant, ¢, is said to be an isometry of the manifold .%. The set of isometries forms a group, called the isometry group of .. Curves and vectors are closely related geometrical objects. Consider a curve A, described by real coordinates x‘ = x(t), and a differentiable function f:. “@ > R. Then the directional derivative of f along the curve 2 is given by df _dx' of a a (D.1) and the operator dx’ @ X=FaG (D.2) maps any function f to its directional derivative along 4. In mathematical language, X is called a vector, and the dx‘/dt are its components. The operator X is the natural generalization of a tangent vector to curved space. This definition of a vector can be applied to a space-filling set of curves as well. The components dx‘/dt become functions on .W, and X = (dx'/dt) 6/dx! is known as a vector field. Alternatively, given a set of continuous functions X‘ on ./, it is always possible to define an associated set of integral curves x'(t) as solutions to 240 APPENDIX D the differential equations, dx! dt = Xx! (D3) The corresponding vector field is just X = X‘0/dx'. Locally, such curves can never cross because the solutions to (D.3) are unique. They are also globally well defined because (D.3) holds at each point of the manifold M. Thus we have seen that sets of space-filling curves are in one-one cor- respondence with vector fields X. The map ¢, defines a motion along the integral curve defined by X. As with any map of a manifold onto itself, ¢, induces a map between vectors in the tangent space. The induced map allows us to compare vectors at different points along integral curves. To construct it explicitly, let x‘ denote the coordinates at p, and x” the coordinates at p’. Then let Y be a vector field, with components Y{(x) at p. The components Y‘(x) at pcan be mapped to components Y(x’) at p’ as follows, ox" vipe(x’)). (D4) Yq) = dxi Equation (D.4) defines a map of vectors at p onto vectors at p’. It is sometimes called Lie transport. The Lie-transported vector field, é Y I (D.5) Yin) vi? ox is defined for all points p’ along the integral curve. Infinitesimally, if x" = x! + X'60, (D4) reduces to 2; ; ox'.. §=6aYF Flee) = Vie) + 35 Wide ~ 5 HSE, (D.6) and the new field ¥ is infinitesimally close to Y. Since Y and Y are both defined at the same points, it makes sense to take their difference and construct the Lie derivative of Y with respect to x: (Ex¥)! (D.7) APPENDIX D 241 The Lie derivative of two vector fields gives a third vector field on the manifold ./. Using similar logic, the definition of the Lie derivative can be gen- eralized to any other tensor field. For example, an expression analogous to (D.4) implies that the Lie derivative of a covariant vector field is as follows: é a (Ex¥) = SGV + Via X. (D.8) In a similar fashion, the Lie derivative of the metric is given by 6 k ra Gin GX + Dix (Exg); = X* ox’ = VX, + VX, (09) where X; = g,,X/ and V;X, = 6,X, — 's,X,, contains the torsion-free con- nection compatible with the metric g;;. A field is invariant under Lie transport if it has a vanishing Lie deriv- ative. If the metric is invariant, then (£x9)ij = ViX; + VjX; = 0 (D.10) for some vector field X. In this case, X generates an isometry of the manifold ./. It is called a Killing vector field, and (D.10) is known as Killing’s equation. The Killing vectors generate the continuous symmetries of a manifold. These symmetries close into the isometry group. Indeed, it is not hard to show that the Lie bracket of two Killing vectors gives another, [X@,X] = — fox, (D.11) where the f* are the structure constants of the isometry group G. Let us now assume that our manifold is Kahler, with metric g; and complex coordinates a‘ and a*‘. We shall focus our attention on the ana- lytic isometries, those that preserve the analytic structure of the manifold. This requires that the associated Killing vectors be holomorphic vector fields, XY = Xi(q) oO éa' i a X*O) = XHOGr) —, (D.12) dar 242 APPENDIX D The index (b) labels the Killing vectors and runs over the dimension d of the isometry group G. Because the X“ are holomorphic, Killing’s equation (D.10) reduces to the following form: VX + VX =0 VX + VX = 0. (D.13) On a Kahler manifold, the first equation is automatically satisfied because of the definition of the covariant derivative. The second is an integrability condition; it is locally equivalent to the statement that there exist d real scalar functions D(a,a*), such that é éaé Jip X* = i DO ise GipeX™ -~i- D®. (D.14) aa*i The D are known as Killing potentials and defined up to constants c, D® + D® + c. In Chapter XXIV we show that the freedom to redefine the potentials is related to the Fayet-Iliopoulos D term for Abelian groups. The relations (D.14) can be inverted to give the Killing vectors in terms of the Killing potentials, it _ ap 9 XH = —igil agai D® xsi = gt © po, (D.15) éa' The requirement that the fields X be holomorphic places a constraint on the D®. Solving this constraint is equivalent to solving (D.13). In general, it may be difficult to find the Killing potentials on a given Kahler manifold. Because of the holomorphic structure, the Killing vectors X and X* generate independent representations of the isometry group G. They obey the Lie bracket relations, [XxX] = — fexo [X# XH] = — fixe [X@,X*0] = 0, (D.16) APPENDIX D 243 where the f*™ are the structure constants of G. The Killing potentials D® also transform under the isometry group. As shown in Exercise 3, they can be chosen to transform in the adjoint representation, [xm & + x 2a] = — eee, 17 0 eat This fixes the constants c for non-Abelian groups. For each U(1) factor, however, there is an undetermined constant c. Let us now turn our attention to the variation of the Kahler potential under an isometry in G. Such an isometry is generated by the Killing vectors X and X*: OK = (OX + ck OX*NK, (D.18) Note that we have used a complex parameter ¢“), and that the hermitian nature of the Kahler potential is preserved. It is straightforward to show that (D.18) can be rewritten as follows: OK = OF 4 ckOpem _ (9 _ geo, (D.19) where the F = XK + iD are analytic functions of the coordinates, OF” i) , OD” Fan = 9X + iT = 05 (D.20) and we have used (D.14). For real parameters «, (D.19) reduces to a Kahler transformation. For complex parameters, however, it is not of Kahler form; there is a change in K proportional to the Killing potential D®. In Chapter XXIV this plays an important role in the construction of gauge-invariant actions. REFERENCES N. Dragon, M. G. Schmidt, and U. Ellwanger, Nucl. Phys. B255, 549 (1985). W. Buchmiiller and W. Lerche, Annals of Phys. 175, 159 (1987). EXERCISES (1) Show that the Lie bracket of two Killing vectors gives another. 244 APPENDIX D (2) Demonstrate that the first equation in (D.13) is automatically satisfied on a Kihler manifold, and that the second is locally equivalent to (D.14). (3) Prove that the Killing potentials can always be chosen to satisfy (D.17). This can be done by first differentiating the left-hand side with respect to a’, and then using the relations introduced above to obtain the a’ derivative of the right-hand side of (D.17). The proof can be completed by repeating the procedure, this time differentiating with respect to a*’, (4) Show that xia) a D® 4 x*i) aoa D®=0. a i (5) Consider the manifold with Kahler potential K = a*‘a'. Verify that the differential operators xXx = ~ iat, © 0 #4) igki TOI X*O = ja*/T' k 5gak are indeed Killing vectors, where the T*, are given in (7.14). Show that their Lie brackets close into (D.16). (6) Given the Kahler potential K = log(1 + aa*), and the Killing potentials 1 a+a* i a—a* 1/1 — ata DY = — Pe pe = 1 87" pay. 2 ( 7 8) 2(1 + a*a) 2(1 + a*a)’ 3G + a) find the Killing vectors X using (D.15). Compute their commuta- tors and identify the isometry group G. APPENDIX E NONLINEAR REALIZATIONS Nonlinear realizations play an important role in theories with sponta- neously broken symmetries. They were first studied in the context of chiral dynamics, where they were used to describe the pion and its interactions. They can also be applied to theories with spontaneously broken super- symmetry, where they are used to derive low-energy theorems for the Goldstone fermion. In this appendix we will develop the necessary for- malism for the case of compact, connected, semisimple Lie groups. This will serve as a guide for our study of spontaneously broken supersym- metry, where similar results can be proved using different techniques. We start by assuming that we have a manifold .# and a group G of transformations that act on ./@, x =g°x, (E.1) where g € G, and x, x’ are points of .@. These transformations induce a realization of G on the coordinates in each neighborhood of ./@. Such realizations clearly include the case of linear representations, but they also include more general realizations that cannot be reduced to linear trans- formations by appropriate coordinates on ./. Given a particular realization, one would like to know whether or not it can be reduced to a linear transformation. For the case of compact, connected, semisimple Lie groups, there is a simple answer: a realization can be linearized (in a given coordinate patch) if and only if it leaves a point in the patch invariant. Now, a linear transformation always leaves the origin invariant, so the first direction is trivial. The other direction, however, is a little less ob- vious. Therefore, let us assume that we have a point x,)¢.@ that is invariant under all the transformations in G, 9° Xo = Xo- (E2) We will explicitly construct a set of coordinates that linearize the trans- formation (E.1) in the neighborhood of xp. Since Xp is invariant, we assign 246 APPENDIX E it the coordinate 6, Away from x9, we choose an arbitrary set of co- ordinates, denoting the coordinates of x by X. In terms of these para- meters, the transformation (E.1) has a power series expansion, ‘= 9° X = Dg )X + O18"), (E3) #L where D(g) is a matrix expression. [In Exercise 1 you will show that D(g) is a matrix representation of G.] The constant term is absent because the origin is invariant. We now introduce new coordinates } at the point x as follows: F = faulg\D- "gg. (E4) The integration is over the group G, and is well defined for compact groups. The measure du(g) can be chosen to be left- and right-invariant, Au(Go9) = Au(g99o) = au(g), (E.5) and normalized so that Janta) = 1. £9) With these conventions, it is easy to see that J =X + O(%), (E.7) so (E.4) is an allowed change of coordinates. Let us now study the action of G on the coordinates ¥. We find 40°F = fadnlgD-"@)a- 40° % = fadutgg.)Dt40)D~ GoD "A)g - Go % = Diao) f dug\D- Yaa - & = Diao), (E8) which demonstrates that the coordinates } do indeed linearize the trans- formation (E.1). This construction relies heavily on the properties of group integration. Curiously enough, similar results hold even when group integration can- not be properly defined. For example, in Chapter XXVI we study the APPENDIX E 247 case of supersymmetry, in which a supergroup of transformations acts on superspace. The linearization condition still holds, even though the group volume is formally zero. Given an arbitrary point xo €.@, the transformations that leave the point invariant close into a group H, called the stability group of Xo. In general, H is a proper subgroup of G. We have just seen that the trans- formations in the stability group are precisely those that can be realized linearly in the neighborhood of the point xo. In preparation for what follows, let us now shift our attention to the submanifold 4” of .@ that can be reached by group transformations act- ing on the point xo, X= 9'Xo. (E.9) Clearly, the points in .” are in one-one correspondence with the coset space G/H. This space has a natural parametrization in terms of the group parameters. An arbitrary element of G can be written in the form, T, (E.10) where the parameters ii and é are real, In this expression, the T are the (hermitian) generators of H, while the X are the generators of G in the or- thogonal complement of H. Two elements g and g’ € G correspond to the same point of G/H if they are related by a right H transformation: g ~ g’ if g = g'u’, for some w' of the form a (E.11) This implies that the cosets can be parametrized by the group elements, vae BX (E.12) and that the @ are coordinates of the space G/H. With these conventions, an element go € G acts on the cosets by left multiplication, ig Joe (E.13) The coordinates & are completely determined in terms of & and go, 90: & + B(Ego)- (E.14) 248 APPENDIX E The parameters i’ can be computed as well; they too depend on go and the coset parameters ¢: go: i > W(E,go). (E.15) For elements gg = uo € H, the transformations (E.14) and (E.15) can always be written in closed form. Then VU Ul =UgvUg ‘Uy = v'u', (E.16) where v’ = upvug ' and u' = up. In terms of the coordinates é, this implies e ae Mut, (E.17) = Ue where v’ = ugvug! and wu’ = uo. In terms of the coordinates Z, this implies ug: € -» & = Dlup)é. (E.18) For transformations go € G that are not in H, however, Eqs. (E.14) and (E.15) cannot generally be written in closed form. There is a special case, however, where these transformations can be made more explicit. This is when the structure relations of G admit the automorphism, a > + -X, (E.19) al in which case G/H is called a symmetric space. To see how this works, consider a transformation vo, v > vgn = vu. (E.20) This can be rewritten by first taking the inverse and then applying the automorphism (E.19), voy Sule. (E.21) Combining the two expressions, we find rod? ~lugp. (E.22) APPENDIX E 249 In terms of the coordinates €, this implies eK = ye 2 Ky, (E.23) This is a manifestly nonlinear transformation law. Note that @ can be transformed to zero if we take vg = e* * We will now show that we can use these results to promote any rep- resentation of H to a realization of G, with the help of the coset param- eters 2, We start with a representation D, which acts linearly on a vector space spanned by wp, uy: b + Dluohb, (E.24) for uo € H. Then, using (E.15), this transformation can immediately be extended to a realization of G: 90.0 > Dee“* Tp. (B.25) The variables ii’ parametrize an clement of H, but they are functions of € and go. To show that (E.25) is indeed a realization of G, we compute ge B® = eo ® ge X= ek gage * =e ; = eB" ‘en i To-iw -T (E.26) From this we see that (E.27) which implies Blew”) = Dee“ HDle“# 7), (E.28) since D is a representation of H. In this way we can realize the group G on the space spanned by the vectors . The transformation (E.25) plays a special role in the study of nonlinear realizations. It defines what is known as a standard realization of the group G. The realization is standard because any realization of G that linearizes on H can be reduced to this form with the help of the coset parameters €. 250 APPENDIX E To see this, we assume as before that we have a manifold ./ and a group G that acts on ./# as a group of transformations. We also assume that we can choose coordinates (2%) in some neighborhood % of .%, where the coordinates ¢ parametrize the points in Y that can be reached from (0,%) by the action of G. Because of the construction (E.8), the parameters 7 can be chosen to transform linearly under H. The transfor- mations of the € are completely determined by (E.14). Therefore, under an H-transformation, we have up ED) = (Duo) Duo), (E.29) where @ and 7 transform in the representations D and D, respectively. Now, among the full set of G transformations, there is one that trans- forms € to zero, (E.30) This transformation also takes ¥ to J, which can be computed because we know the action of G on the manifold .#. The parameters y transform in the representation D under H, as follows from (E.29). We shall now construct a new coordinate system on -# as follows. Start at a point x, parametrized by the coordinates (&), and map it to the point (6,/) as in (E.30): ef ¥. (2%) = GW). (E31) Then take the new coordinates at the original point x to be given by (2,y)). This defines an acceptable coordinate transformation on the manifold because the Jacobian of the transformation (€,%) > (€/) is nonvanishing near the origin. In terms of the new coordinates, the transformation (E.31) can be written as e°*- (2) = (6). This allows us to show that the new coordinates #/ transform as a standard realization: 9 (EW) (E32) Together with @, they are the natural coordinates on .W adapted to the action of G. . . In physical applications, the coordinates € and w are x”-dependent fields. The coset coordinates & play the role of the Goldstone bosons that APPENDIX E 251 arise from spontaneously breaking G to H. The standard realizations describe the other fields that transform in representations of the unbroken group H. In this appendix, we have seen that any representation of H can be extended to a realization of G with the help of the Goldstone bosons é To write down invariant Lagrangians we would like to have covariant derivatives that transform as standard realizations. Our general argu- ments tell us that such derivatives must exist. Constructing them provides a straightforward application of what we have just learned, as well as a nice illustration. To find the covariant derivatives, we start from the manifold parame- trized by (Enh.OnZ Ont). As above, we apply a group transformation with go = e* *. This gives EGE Ont) = Ol Ang Ani), (E33) and from our general prescription we know that A,,¢ and A,,\/ are covariant derivatives that transform as standard realizations. To compute A,,¢, we start from the formula (E.13), Joe (E.34) and differentiate with respect to x”, GoOme 2 * = (Bye Xe“ TF 4 eB XG,e-M TF), (E35) As before, the parameters 2” and depend on x” through é. We now choose Jo = e&* &-%, which transforms & and i’ to zero at the point x”. This gives ef Fae BX = ae BF + 807% = 10_o' X — 6, - = -iA,€-X — (E.36) Equation (E.36) allows us to compute Ane as a function of the parameters &. In Exercise 3 we will see that A,, .© indeed transforms as a standard realization. . Similar techniques can be used to find A,,y. One starts by differentiating (E.25), Anil = OmBle~* Py + Dee) Omuh- (E37) 252 APPENDIX E As above, one then takes gy = ef ¥ to find eX. Ob = ~i(0,i TH + OnGlenn-o- (E.38) Comparing (E.38) with (E.36), we find AW = Ont — iV, Typ. (E.39) In Exercise 4 one is asked to show that A,,/ transforms as a standard realization. REFERENCES S. Coleman, J. Wess, and B. Zumino, Phys. Rev. 177, 2239 (1969). J. Wess in Current Algebra and Phenomenological Lagrange Functions, Springer Tracts in Modern Physics 50, G. Héhler, ed., New York, Springer (1969). EXERCISES (1) Show that the matrices D, defined in (E.3), form a representation of the group G. (2) Demonstrate that A,,\/ transforms as a standard realization. Start by eliminating go between (E.34) and (E.35): Xeni T Xa, 2), er Epi TEKH o- EH = (Ae +e Then multiply on the left by e’* ** and e® 7, to find ef Fae ER = ell Tei Tae FN T + ev Tee tF, This shows that and (4) Use the transformation law for V,, to show that A,,/ transforms as a standard realization. APPENDIX F NONLINEAR REALIZATIONS AND INVARIANT ACTIONS In this appendix we will continue our study of nonlinear realizations. We will use the methods introduced in Appendix E to show that an action invariant under a group H can be promoted to a new action invariant under a larger group G > H. The results derived here are used in Chapter XXIV to construct the gauge invariant matter couplings in superspace. We start by assuming we have a Lagrangian &y which is a function of certain fields A’. The Lagrangian is invariant under a symmetry group H. The fields A‘ are arbitrary, except that they have well-defined trans- formations under a group G > H. The Lagrangian %,, however, is not invariant under the full group G. Instead, it has a variation 6%, # 0. In this appendix we will construct a counterterm “cp whose variation precisely cancels that of Y,. We will build the counterterm out of the fields A‘, together with fields €” that parametrize the coset G/H. We impose the condition that %7 must vanish when €@ = 0. In this way the Lagrangian Lo = Ly + Lor (F.1) is invariant under the full group G, and reduces to Yy for €® = 0. As in Appendix E, let us split the transformations in G into two classes, those in H and those not. Under a transformation uy € H, the Lagrangian Ly, is assumed to be invariant: by Ly = —iu?TOL, = 0, (F.2) where the T are differential operators that act on the fields A‘ and gen- erate the transformations in H. Under a transformation vg € G, Zy has an infinitesimal variation of the form SouLu = —iv@XLy = —vOD, (F.3) where the operators X® generate the transformations in G that are in the orthogonal complement of H. We see that we need to find a function 254 APPENDIX F Lop(A‘E™), such that UST Ley = 0 ROL. = ie D®, (4) subject to the boundary condition Lop Ai0) = 0. (E.5) In these expressions, the operators T® and X® act on the fields 4‘ and on the parameters €” in the counterterm Lagrangian. We shall now find £7 as follows. We first compute (Hiv P LON Lop = (—ivP RO)" PD), (F.6) This can be exponentiated to give - afore ~ exe eto Go = Lop + ——aaa jh) iP RO (PD), (7) where, on the right-hand side, the differential operators X¥ reduce to operators 6A,(6/5A;) because the D® do not contain the fields €®. We can now solve for Lor by noting that e~**™ transforms Lop with parameter v'”), In Appendix E we showed that such a transformation with parameter v\? = —é maps €* to zero. Therefore, in conjunction with the boundary condition (F.5), this implies igo Fe gv. =© — ep cr jEOX® § = f da exp(ialX EMD, (F.8) where the derivatives in ¥® do not act on the fields €. It is a useful exercise to check that (F.8) satisfies (F.4), following the steps outlined in Exercises 2 and 3. REFERENCES S. Samuel, Nucl. Phys. B245, 127 (1984). J. Bagger and J. Wess, Phys. Lett. 199B, 243 (1987). APPENDIX F 255 EXERCISES (1) Exponentiate (F.6) to find (F.7). (2) Derive the conditions on D™ that follow from applying the group commutators on Lor, XOPH — XHYO = j¢srpo FOP = if erp, (3) Use the relations of Exercise 2 to show that (F.8) obeys (F.4). APPENDIX G GAUGE INVARIANT SUPERGRAVITY MODELS In this Appendix, we will write the most general gauge invariant super- gravity model in terms of component fields. We start with the superspace Lagrangian, as given in Chapter XXV: ga [ee 26| 5 (BG - sR) exp| — [K(®,*) + reer} +75 ie 5 Han(®)WOW + 10) | + he. (G.1) Then, using the techniques developed in Chapters XXI through XXV, we expand this Lagrangian in terms of component fields. This gives 1 ~ in 1 L= 3 Ch — €G:pDyA'G"AN — ~ eg’DiD® — cgi PO"D yl + eT, Dn 1 1 _ gel Nan Em Pm + 3 eh gy et Fy OF pe 5 Ln yh + Veg F" Dy dO} + = 5 Li op FmlekamZO™] + VegginXtyi2 + /eggipXinv 2 ig * = 7 V2e9 FMayDOP2 + 5 2e9 Onlin D OZR - ny V2€ Ahan Lor 2 Fg? — fvie0 Opal LO LOE yg? 1 x 1 7 my (a - ie + 5 gD ano" lig = si-j=m nT — FV etn DAML Oa m — 5 \2e0,pF,A LG" On APPENDIX G 257 + Luo hay + VFO” a] Fan? + Fay] Io. - ae +4 egieLieK VT Dm + Um” one 1 iykzige ~ gL Ince — Ripe MET 1 cd) A inn. iTax + 76 L2G cen + ARO" Bh Opehina CO"Y HE, A” 1 +; 1 ey, Olah PROM + 3 eV ie Ochi PPAOT + a ehR—! ih, I hray Ay 1 — + FZ CHM Glin Ophfaa ZB LA 1 91) Fle) Fl - Cg Meany 0 phg AO LO TOT 3 ), m Fla) qe ~ 76 CF A A)F me + GV ‘p 2d, ral oY, FR) + = fad e™ 302 | + i 2etultn| ZOE 6A”) 4 — ile oe “7102 — eexp(K/2) [Pees + PU ED, 5 ! Appyord, + 4 5 V2D P*T'G%, on paid ol i + 5 DP + a 1. — {FMD P* Aha AOR — 5g DIP Opie ZA a — eexp(K)[gi"(D,P\(D,P)* — 3P*P], (G.2) 258 APPENDIX G where h®,,,, = Re Hoy)| and h'(q) = Im H,q|. The covariant derivatives are given by TA! = OA! — Gn! X'e ~ oo ~ a Fri = Omi’ + LOm + Tp FnAles = gon! — Ai — Kj Dy Ah — 59m! Im Faz Gy h® = Byh® + 2 Cy — gf rp OKO 1 > eine . i , + {KIA — KpFyA VU + 5 gun! lin Foyh® Dyn = Onn + Var Ione ~ i + GK jG? — Kp FuA* hy + 5 Omi" Li Fath D,P = P, + K,P DD,P = Py + KyP + K,D,P + KjD,P — K;K,P —T%D,P. (G3) In these expressions, the fields in the vector multiplet are defined to have upper gauge indices, such as v,” and 4. The Killing vectors and Killing potentials have lower gauge indices, X{,, and Dj. These indices can be raised and lowered with h*,,,, and its inverse. Using these conventions, one can check that the Lagrangian (G.2) is invariant under the following set of supergravity transformations: 8m" = (COV, + C5) 6Al = 2x! bz! = iv 20"EG,,Ai — Ti, 6,Aly* 1 ag A + G(KjOA! — K 6A yz! — 2087 g'DyP*E 1. a + VIG Opin OZ SV = (CO, + 6,4) APPENDIX G 259 1 . 5,20 = Full — 5 (Kj8.Ai — Kp5, A — igd Ye iaby~1 © 1 Rep, Rir—1 9 pe zit + V2 Ailey A — gv2th ( Oph TAO Sehm = 2Fnh 5 Om GinXio"Z 5 Ona + GpE Aq o"h 1 / . ; — 7 (KOA! — Kp AM Nm + eS? Pog. (G4) The Lagrangian (G.2) is the starting point for phenomenological studies of supergravity theories. REFERENCES G. Girardi, R. Grimm, M. Miller, and J. Wess, Z. Phys. C26, 427 (1984). P. Binétruy, G. Girardi, and R. Grimm, LAPP-TH-275/90 (1990).

You might also like