1: \documentclass{pspum-l}
2: \usepackage{amssymb, amscd}
3: \bibliographystyle{unsrt}
4: %\baselineskip=24pt plus 2pt
5: %\raggedbottom
6: %\input epsf.tex
7: \overfullrule=0pt
8: \begin{document}
9:
10:
11: \font\nt=cmr7
12: \def\note#1
13: {\marginpar
14: {\nt $\leftarrow$
15: \par
16: \hfuzz=20pt \hbadness=9000 \hyphenpenalty=-100 \exhyphenpenalty=-100
17: \pretolerance=-1 \tolerance=9999 \doublehyphendemerits=-100000
18: \finalhyphendemerits=-100000 \baselineskip=6pt
19: #1}\hfuzz=1pt}
20:
21: \def\note#1{}
22:
23: \newcommand{\bignote}[1]{\begin{quote} \sf #1 \end{quote}}
24:
25: %I took this out \renewcommand{\theequation}{\thesection.\arabic{equation}}
26: %\let\myection=\section
27: %\renewcommand{\section}{\setcounter{equation}{0}\mysection}
28:
29: % Peter's chapters have no blank pages between.
30: %
31: %\def\cleardoublepage{\clearpage}
32:
33: %\topmargin=-.5in
34: %\oddsidemargin=.1in
35: %\evensidemargin=.1in
36: %\vsize=23.5cm
37: %\hsize=16cm
38: %\textheight=23.5cm
39: %\textwidth=16cm
40:
41: %\def\upbracketfill{
42: % $\leaders\hrule\hfill$}
43: %\def\underbracket#1{\mathop{\vtop{\ialign{##\crcr
44: %$\hfil\displaystyle{#1}\hfil$\crcr\noalign{\kern3pt\nointerlineskip}
45: % \upbracketfill\crcr\noalign{\kern3pt}}}}\limits}
46: %\def\upslurfill{
47: % $\bracelu\leaders\vrule\hfill\braceru$}
48: %\def\underslur#1{\mathop{\vtop{\ialign{##\crcr
49: %$\hfil\displaystyle{#1}\hfil$\crcr\noalign{\kern3pt\nointerlineskip}
50: % \upslurfill\crcr\noalign{\kern3pt}}}}\limits}
51: % Usage: \underslur{}
52:
53: %use when boldfacing terms in equations-use \bold{xxxx}instead of
54: %{\mbox{\boldmath$(\p_i g^{ij})$}}
55: %\def\bold#1{\mbox{\boldmath$#1$}}
56:
57: %\def\bop#1{\setbox0=\hbox{$#1M$}\mkern1.5mu
58: % \vbox{\hrule height0pt depth.04\ht0
59: % \hbox{\vrule width.04\ht0 height.9\ht0 \kern.9\ht0
60: % \vrule width.04\ht0}\hrule height.04\ht0}\mkern1.5mu}
61: %\def\bo{{\mathpalette\bop{}}} % box
62:
63: % Young tableaux: \boxup<a>{\boxes<a>...\boxes<b>}
64:
65: %\def\boxes#1{
66: % \newcount\num
67: % \num=1
68: % \newdimen\downsy
69: % \downsy=-1.5ex
70: % \mskip-2.8mu
71: % \bo
72: % \loop
73: % \ifnum\num<#1
74: % \llap{\raise\num\downsy\hbox{$\bo$}}
75: % \advance\num by1
76: % \repeat}
77: %\def\boxup#1#2{\newcount\numup
78: % \numup=#1
79: % \advance\numup by-1
80: % \newdimen\upsy
81: % \upsy=.75ex
82: % \mskip2.8mu
83: % \raise\numup\upsy\hbox{$#2$}}
84:
85:
86: %%%%%%%%%%
87: %\font\mybbb=msbm10 at 8pt
88: %\font\mybb=msbm10 at 12pt
89: %\def\bbb#1{\hbox{\mybbb#1}}
90: %\def\bb#1{\hbox{\mybb#1}}
91: %\def\id{{1 \kern-.30em 1}}
92: %\def\id{\protect{{1 \kern-.28em {\rm l}}}}
93: \def\I {\mathbb{1}}
94: \def\Z {\mathbb{Z}}
95: \def\pRe{\mathbbb{R}}
96: \def\Re {\mathbb{R}}
97: \def\C {\mathbb{C}}
98: \def\pC{\mathbbb{C}}
99:
100: \def\H {\mathbb{H}}
101: \def\ni{\noindent}
102: \newcommand\noi{\noindent}
103: \def\nn{\nonumber}
104: \newcommand\seq{\;\;=\;\;}
105:
106: \def\eq{\begin{equation}}
107: \def\eqe{\end{equation}}
108: \def\eqa{\begin{eqnarray}}
109: \def\eqae{\end{eqnarray}}
110: \def\bea{\begin{eqnarray}}
111: \def\ena{\end{eqnarray}}
112:
113: \def\st{\star}
114: \def\dZ2p{\frac{dZ_1}{2\p i}}
115:
116: %\def\half{\frac{1}{2}}
117: %\def\({\left(} \def\){\right)} \def\<{\langle } \def\>{\rangle }
118: %\def\[{\left[} \def\]{\right]} \def\lb{\left\{} \def\rb{\right\}}
119:
120: \newcommand{\be}{\begin{equation}}
121: \newcommand{\ee}{\end{equation}}
122: \newcommand{\ba}{\begin{eqnarray}}
123: \newcommand{\ea}{\end{eqnarray}}
124: \newcommand{\al}{\mbox{$\alpha$}}
125: \newcommand{\als}{\mbox{$\alpha_{s}$}}
126: \newcommand{\s}{\mbox{$\sigma$}}
127: \newcommand{\lm}{\mbox{$\mbox{ln}(1/\alpha)$}}
128: \newcommand{\bi}[1]{\bibitem{#1}}
129: \newcommand{\fr}[2]{\frac{#1}{#2}}
130: \newcommand{\sv}{\mbox{$\vec{\sigma}$}}
131: \newcommand{\gm}{\mbox{$\gamma_{\mu}$}}
132: \newcommand{\gn}{\mbox{$\gamma_{\nu}$}}
133: \newcommand{\Le}{\mbox{$\fr{1+\gamma_5}{2}$}}
134: \newcommand{\R}{\mbox{$\fr{1-\gamma_5}{2}$}}
135: \newcommand{\GD}{\mbox{$\tilde{G}$}}
136: \newcommand{\gf}{\mbox{$\gamma_{5}$}}
137: \newcommand{\om}{\mbox{$\omega$}}
138: \newcommand{\Ima}{\mbox{Im}}\newtheorem{thm}{Theorem}[subsection]
139: \newtheorem{mthm}{Theorem}
140:
141: \newcommand{\Rea}{\mbox{Re}}
142: \newcommand{\Tr}{\mbox{Tr}}
143: \newtheorem{cor}[thm]{Corollary}
144: \newtheorem{prop}[thm]{Proposition}
145: \newtheorem{lemma}[thm]{Lemma}
146: \newtheorem{claim}[thm]{Claim}
147: \newtheorem{conj}[thm]{Conjecture}
148: \newtheorem{definition}[thm]{Definition}
149: \newtheorem{proposition}[thm]{Proposition}
150: \newtheorem{condition}[thm]{Condition}
151: \numberwithin{equation}{section}
152:
153:
154: %\newcommand{\beta}[2] {\mbox{$\bar{\beta}_{#1}^{\hspace*{.5em}#2}$}}
155: %\newcommand{\homega}[2]{\mbox{$\hat{\omega}_{#1}^{\hspace*{.5em}#2}$}}
156:
157: \def\a{\alpha}
158: \def\b{\beta}
159: \def\c{\chi}
160: \def\d{\delta}
161: \def\e{\epsilon} % Also, \varepsilon
162: \def\f{\phi} % \varphi
163: \def\g{\gamma}
164: \def\h{\eta}
165: \def\i{\iota}
166: \def\j{\psi}
167: \def\k{\kappa} % Also, \varkappa (see below)
168: \def\l{\lambda}
169: \def\m{\mu}
170: \def\n{\nu}
171: \def\o{\omega}
172: \def\p{\pi} % Also, \varpi
173: \def\q{\theta} % \vartheta
174: \def\r{\rho} % \varrho
175: \def\s{\sigma} % \varsigma
176: \def\t{\tau}
177: \def\u{\upsilon}
178: \def\x{\xi}
179: \def\z{\zeta}
180: \def\D{\Delta}
181: \def\F{\Phi}
182: \def\G{\Gamma}
183: \def\J{\Psi}
184: \def\L{\Lambda}
185: \def\O{\Omega}
186: \def\P{\Pi}
187: \def\Q{\Theta}
188: \def\S{\Sigma}
189: \def\U{\Upsilon}
190: \def\X{\Xi}
191: \def\del{\partial}
192: \def\pa{\partial}
193:
194: % Calligraphic letters
195:
196: \def\ca{{\mathcal A}}
197: \def\cb{{\mathcal B}}
198: \def\cc{{\mathcal C}}
199: \def\cd{{\mathcal D}}
200: \def\ce{{\mathcal E}}
201: \def\cf{{\mathcal F}}
202: \def\cg{{\mathcal G}}
203: \def\ch{{\mathcal H}}
204: \def\ci{{\mathcal I}}
205: \def\cj{{\mathcal J}}
206: \def\ck{{\mathcal K}}
207: \def\cl{{\mathcal L}}
208: \def\cm{{\mathcal M}}
209: \def\cn{{\mathcal N}}
210: \def\co{{\mathcal O}}
211: \def\cp{{\mathcal P}}
212: \def\cq{{\mathcal Q}}
213: \def\car{{\mathcal R}}
214: \def\cs{{\mathcal S}}
215: \def\ct{{\mathcal T}}
216: \def\cu{{\mathcal U}}
217: \def\cv{{\mathcal V}}
218: \def\cw{{\mathcal W}}
219: \def\cx{{\mathcal X}}
220: \def\cy{{\mathcal Y}}
221: \def\cz{{\mathcal Z}}
222:
223: \def\vecnab{\vec{\nabla}}
224: \def\vx{\vec{x}}
225: \def\vy{\vec{y}}
226: \def\arrowk{\stackrel{\rightarrow}{k}}
227: \def\kbar{k\!\!\!^{-}}
228: \def\karrow{k\!\!\!{\rightarrow}}
229: \def\arrowl{\stackrel{\rightarrow}{\ell}}
230: \def\var{\varphi}
231:
232:
233: %\fig 3.15in by 3.88in (15 scaled 800) make figure smaller
234: %Use this when scanning figures into an equation file - example -\fig 1.24in by
235: %1.31in (3.4)
236: %\def\fig #1 by #2 (#3){\vbox to #2{
237: % \hrule width #1 height 0pt depth 0pt\vfill\special{picture #3}}}
238:
239: % Usage: #1 = width, #2 = height (see pictures window),
240: % #3 = name of picture, or name "scaled" magnification ¥ 1000
241: %
242: % For best bitmaps, do at 72dpi, then scale 240
243:
244: %\def\hook#1{{\vrule height#1pt width0.4pt depth0pt}}
245: %\def\leftrighthookfill#1{$\mathsurround=0pt \mathord\hook#1
246: % \hrulefill\mathord\hook#1$}
247: %\def\underhook#1{\vtop{\ialign{##\crcr % |_| under
248: % $\hfil\displaystyle{#1}\hfil$\crcr
249: % \noalign{\kern-1pt\nointerlineskip\vskip2pt}
250: % \leftrighthookfill5\crcr}}}
251:
252: %\def\under#1#2{\mathop{\null#2}\limits_{#1}}
253:
254: \def\ttZ{\tilde{\tilde{Z}}}
255: \def\ttz{\tilde{\tilde{z}}}
256: \def\tz{\tilde{z}}
257: \def\tZ{\tilde{Z}}
258: \def\tq{\tilde{\q}}
259: \def\del{\partial}
260: \def\half{\frac{1}{2}}
261: \def\st{\star}
262: \def\stam{\stackrel{\rightarrow}{\m}}
263: \def\stab{\stackrel{\rightarrow}{\b}}
264: \def\half{\frac{1}{2}}
265: \def\m{\mu}
266: \def\n{\nu}
267: \def\g{\gamma}
268: \def\d{\delta}
269: \def\in{\infty}
270: \def\nn{\nonumber\\}
271: %%%%%%%%%%%%%%%%
272:
273: \title[Supersymmetry in a simple model]{ Supersymmetry, supergravity, superspace\\ and BRST symmetry in a simple model}
274: %\title {in a simple model}
275: \author{ Peter van Nieuwenhuizen}
276: \address{ C.N. Yang Institute for Theoretical Physics,
277: Stony Brook University, Stony Brook, NY, 11794-3840, USA}
278: \copyrightinfo{2000}% % copyright year
279: {American Mathematical Society}% copyright holder
280: %\dedicatory{To Dennis Sullivan on his 60-th birthday}
281: \keywords{Supersymmetry, superspace, action, Hamiltonian, spinors, conserved
282: charges, constraints, gauge fixing, Faddeev-Popov ghosts, supergravity, BRST quantization}
283: \subjclass[2000]{81T60}
284: %\date{July, 2002}
285: \maketitle
286: %\pagenumbering{roman}
287: %\mathversion{bold}
288: \tableofcontents
289: %\mathversion{normal}
290:
291: %\pagenumbering{arabic}
292:
293: %\baselineskip=24pt
294:
295: %\setcounter{secnumdepth}{-2}
296: \section{Introduction}
297:
298: %\setcounter{secnumdepth}{10}
299:
300: In these lectures we shall introduce rigid supersymmetry,
301: supergravity (which is the gauge theory of supersymmetry) and
302: superspace, and apply the results to BRST quantization. We assume
303: that the reader has never studied these topics. For readers who
304: want to read more about these ``super" subjects, we give a few
305: references at the end of this contribution, but the whole point
306: of these lectures is that one does not need additional references
307: for a self-contained introduction. The reader should just sit
308: down with paper and pencil. We could have decided to begin with
309: the usual models in $3+1$ dimensional Minkowski space with
310: coordinates $x$, $y$, $z$ and $t$; this is the standard approach,
311: but we shall instead consider a much simpler model, with only one
312: coordinate $t$. We interpret $t$ as the time coordinate. For
313: physicists, the $3+1$ dimensional models are the ones of interest
314: because they are supposed to describe the real world. For
315: mathematicians, however, the simpler model may be of more
316: interest because the basic principles appear without the dressing
317: of physical complications. Let us begin with three definitions
318: which should acquire meaning as we go on.
319:
320: \noindent {\bf Supersymmetry} is a symmetry of the action (to be
321: explained) with a rigid (constant) anticommuting
322: parameter%
323: \footnote{ Technically: a Grassmann variable. ``Constant'' means
324: ``independent of the spacetime coordinates $x,y,z$ and $t$''.}
325: (usually denoted by $\epsilon$)
326: between bosonic (commuting) and fermionic (anticommuting) fields
327: (again to be explained). It requires that for every bosonic
328: particle in Nature there exists a corresponding fermionic
329: particle, and for every fermionic particle there should exist a
330: corresponding bosonic particle. So supersymmetry predicts that
331: there are twice as many particles as one might have thought. One
332: may call these new particles supersymmetric particles. These
333: supersymmetric particles will be looked for at CERN (the European
334: high-energy laboratory) in the coming 8 years. So far not a
335: single %\note{*}
336: supersymmetric particle has been discovered: supersymmetry
337: is a theoretical possibility, but whether Nature is aware of this
338: possibility remains to be seen.%
339: \footnote{ This is not the first time a doubling of the number of particles has been
340: predicted. In 1931 Dirac predicted that for every fermionic
341: particle a fermionic antiparticle should exist, and these
342: antiparticles were discovered in 1932. We consider in these
343: lectures only real fields, and the particles corresponding to
344: real fields are their own antiparticles. Hence in these lectures
345: the notion of antiparticles plays no role.}
346:
347: \noindent {\bf Supergravity} is the gauge theory of supersymmetry.
348: Its action is invariant under transformation rules which depend on a local (space- and time-dependent)
349: anticommuting parameter $\epsilon(x,y,z,t)$, and
350: there is a gauge field for supersymmetry which is called the gravitino field.
351: It describes a new hypothetical particle, the gravitino. The gravitino is the fermionic partner
352: of the graviton. The graviton is the quantum of the gravitional field (also called the
353: metric). The astonishing discovery of 1976 was that a gauge
354: theory of supersymmetry requires gravity: Einstein's 1916 theory
355: of gravity (called general relativity) is a product of local
356: supersymmetry. Phrased differently: local supersymmetry is the
357: ``square root of general relativity'', see~(\ref{moya}). (Likewise,
358: supersymmetry is the square root of translation symmetry, see
359: (\ref{Jacobi})).
360:
361: \noindent {\bf Superspace.} In Nature fields can be divided into
362: bosonic (commuting) fields and fermionic (anti-commuting)
363: fields. This is one of the fundamental discoveries of the
364: quantum theory of the 1920's. The anticommuting fields are
365: described by spinors and the bosonic fields by tensors according
366: to the spin-statistics theorem of the 1930's. (Spinors and
367: tensors refers to their transformation properties under Lorentz
368: transformations). One can also introduce in addition to the
369: usual coordinates $x^\mu$ anticommuting counterparts $\theta^\a$.
370: The space with coordinates $x$ and $\theta$ is called
371: superspace. In the case of a four-dimensional Minkowski space
372: (our world) there are four coordinates ($x,y,z$ and $t$) and also
373: four $\theta$'s, namely $\theta^1$, $\theta^2$, $\theta^3$ and
374: $\theta^4$, but in other dimensions the number of $x$'s and
375: $\theta$'s are not the same.%
376: \footnote{ The $x^\mu$ transform as vectors under the Lorentz group while the
377: $\theta^\a$ transform as spin $1/2$ spinors.}
378: These $\theta$'s
379: are Grassmann variables \cite{Berezinfa}, for example $\theta^1
380: \theta^2 =- \theta^2 \theta^1$ and $\theta^1 \theta^1 =0$. In
381: our case we shall have one $x^\mu$ (namely $t$) and one
382: $\theta^\a$ (which we denote by $\theta$). In superspace one can
383: introduce superfields: fields which depend both on $x$ and
384: $\theta$.Because $\theta^2=0$, the superfields we consider can be
385: expanded as $\phi(t,\theta)=\varphi(t)+\theta\psi (t)$. This
386: concludes the three definitions.
387:
388: Supersymmetric quantum field theories have remarkable properties.
389: Leaving aside the physical motivations for studying these
390: theories, they also form useful toy models. We present here an
391: introduction to supersymmetric field theories with rigid and
392: local supersymmetry, both in $x$-space and in superspace, in the
393: simplest possible model%
394: \footnote{ Actually, an even simpler model than the one we present in these lectures
395: exists. It contains constant fields, so fields which do not even
396: depend on $t$ and $\theta$. These so-called matrix models are
397: important in string theory, but they do not have enough structure
398: for our purposes, so we do not discuss them.}.
399: To avoid the complications due to ``Fierz rearrangements"
400: (recoupling of four fermionic fields $A,B,C,D$ from the structure
401: $(AB) (CD)$ to $(AD)(CB)$)\footnote{Here $(AB)$ means contraction of spinor fields $A$ and $B$.} \note{*}
402: we consider one-component (anticommuting) spinors. Then
403: $(AB) (CD)$ is simply equal to $-(AD) (CB)$. The simplest case in
404: which spinors have only one component is a one-dimensional
405: spacetime, i.e., quantum mechanics. The corresponding superspace
406: has one commuting coordinate $t$ and one anticommuting coordinate
407: $\theta$. Both are real.
408:
409: We repeat and summarize: one can distinguish between rigidly supersymmetric field
410: theories,
411: which have a constant symmetry parameter, and
412: locally supersymmetric field theories whose symmetry parameter is an arbitrary
413: space-time dependent parameter.
414: For a local symmetry one needs a gauge
415: field. For supersymmetry the gauge field has been called the
416: gravitino. (The local symmetry on which Einstein's theory of gravitation is based is diffeomorphism invariance. The gauge field is the metric field $g_{\mu\nu} (x)$). Gauge theories of supersymmetry
417: (thus theories with a local
418: supersymmetry containing the gravitino) need curved spacetime. In
419: other words, gravity is needed to
420: construct gauge theories of supersymmetry, and for that reason local
421: supersymmetry is usually called
422: supergravity. In curved space the quanta of the metric $g_{\mu\nu}$ are massless particles
423: called gravitons. They are the bosonic partners of the gravitinos. Neither gravitons nor gravitinos have ever been directly detected. Classical gravitational radiation may be detected in the years ahead, but the gravitons (the quantized particles of which the gravitational field is composed) are much harder to detect individually. The detection of a single gravitino would have far-reaching consequences.
424:
425: In the last chapter we quantize the supergravity action which we
426: obtained in chapter~3. There are several methods of quantization,
427: all in principle equivalent, but we shall only discuss the BRST
428: method. It yields the ``quantum action", which is the action to
429: be used in path integrals. This method has a beautiful and
430: profound mathematical structure, and that is one of the reasons
431: we chose to include it.
432:
433: The author wrote in 1976 with D.Z. Freedman and S. Ferrara the
434: first paper on supergravity, soon followed by a paper by S. Deser and
435: B. Zumino. However, we will not discuss past work
436: and give references; rather, the present account may serve as a
437: simplified introduction to that work. For readers who want to
438: read further, we include a few references at the end.
439:
440: \mathversion{bold}
441: \section{Rigid $N=1$ supersymmetry in $x$-space}
442: \mathversion{normal}
443:
444: The model we consider contains in $x$-space (or rather $t$-space)
445: two point particles which correspond to a real bosonic field
446: $\varphi (t)$ and a real fermionic field $\l(t)$. We view them
447: as fields whose space-dependence (the dependence on $x$, $y$, $z$)
448: is suppressed. The function $\varphi (t)$ is a smooth function of
449: $t$, so its derivative is well-defined, but for every value of
450: $t$ the expression $\l (t)$ is an independent Grassmann
451: number~\cite{Berezinfa}.%
452: \footnote{For a more recent mathematical treatment,
453: see ``Five Lectures on Supersymmetry'' by D.~Freed (AMS)
454: and articles by Deligne \& Morgan and by Deligne \& Freed in
455: ``Quantum Fields and Strings: A course for mathematicians'' (AMS)} \note{*}
456: So $\l (t_1) \l (t_2) =- \l (t_2) \l
457: (t_1)$. We assume that the concept of a derivative of $\l (t)$
458: with respect to $t$ can be defined, and that we may partially
459: integrate. As action for these ``fields" we take $S = \int L dt$
460: with
461: \eqa L \; ({\rm rigid}) = {1 \over 2} \dot\varphi^2 + {i
462: \over 2} \l \dot\l \label{1one}. \eqae
463: The $\dot\l = {d \over dt} \l$
464: are independent Grassmann variables, so $\dot\l (t) \l (t) =-
465: \l (t) \dot\l (t)$ and $\dot\l (t_1) \dot\l (t_2) =- \dot\l (t_2)
466: \dot\l (t_1)$. In particular they anticommute with themselves
467: and with each other at equal $t$:
468: \eqa \{ \l (t) , \l (t) \} =0, \quad
469: \{ \l (t), \dot\l (t) \} =0, \quad
470: \{ \dot\l (t), \dot\l (t) \} =0.
471: \eqae The symbol $\{ A,B \}$ is by definition $AB+BA$. Later we
472: shall define Poisson brackets and Dirac brackets, which we denote
473: by $\{ A,B \}_P$ and $\{ A,B \}_D$ to avoid confusion. At the
474: quantum level the Poisson and Dirac brackets are replaced by
475: commutators for commuting fields and anticommutators for
476: anticommuting fields, which we denote by $[A,B]$ and $\{ A,B \}$,
477: respectively, and which are defined by $[A,B] =AB-BA$ and $\{
478: A,B \} = AB+BA$.
479:
480: We introduce a concept of hermitian conjugation under which $\varphi (t)$ and $\l (t)$ are
481: real: $\varphi (t)^\dagger = \varphi (t)$ and $\l (t)^\dagger = \l (t)$. Also $\dot\l (t)$
482: is real. Furthermore $(AB)^\dagger = B^\dagger A^\dagger$ for any $A$ and $B$. We define the
483: action by $S=\int L(t)dt$.
484: The action should be hermitian according to physical principles
485: (namely unitarity%
486: \footnote{``Unitarity'' means ``conservation of probability'':
487: the total sum of the probabilities that a given system can decay into any other system should be one.}).
488: We need then a factor $i$ in the second term in (\ref{1one}) in order that $({i \over 2}
489: \l \dot\l )^\dagger = -{i \over 2}
490: \dot\l \l$ be equal to ${i \over 2} \l \dot\l$. In the action we need $\l (t)$ at different $t$. We repeat that for
491: different $t$ the $\l (t)$ are independent Grassmann variables. Thus we need an infinite basis for all Grassmann variables.%
492: \footnote{ In some mathematical studies one takes a finite-dimensional basis for the Grassmann variables.
493: This is mathematically consistent, but physically unacceptable: it violates unitarity.}
494:
495: Physical intermezzo which can be skipped by mathematicians: The
496: term ${1 \over 2} \dot\varphi^2$ is a truncation of the
497: Klein-Gordon action to an $xyz$ independent field, and the term
498: ${i \over 2} \l \dot\l$ is the truncation of the Dirac action for
499: a real%
500: \footnote{ Real spinors are called Majorana spinors,
501: and complex spinors are called Dirac spinors.
502: Already at this point one can anticipate that $\l$ must be real because we
503: took $\varphi$ to be real, and we shall soon prove that there
504: exists a symmetry between $\l$ and $\varphi$.}
505: spinor to one of its components. In higher dimensions the Dirac action in curved
506: space reads (as discussed in detail in 1929 by H. Weyl) \note{*}
507: \eqa \cl \; ({\rm Dirac}) = - ( \det e_\mu{}^m ) \bar\l
508: \g^m e_m{}^\mu D_\mu \l \label{1two}, \eqae where $D_\mu \l =
509: \del_\mu \l + {1 \over 4} \o_\mu{}^{mn} \g_{mn} \l$ with $\g_{mn}
510: \equiv {1 \over 2} [ \g_m , \g_n ]$ the Lorentz generators
511: (constant matrices) and $\o_\mu{}^{mn}$ the spin connection (a
512: complicated function of the vielbein fields $e_\mu{}^m$). The
513: matrices $\gamma^m$ (with $m=0,1,2,\dots ,d-1$) in $d$ spacetime
514: dimensions satisfy Clifford algebra relations,
515: $\{\gamma^m,\gamma^n\}=2\eta^{mn}$. The ``vielbein" fields
516: $e_\mu{}^m$ are the square root of the metric $g_{\mu\nu}$ in the
517: sense that $e_\mu{}^m e_\nu{}^n \eta_{mn} = g_{\mu\nu}$, where
518: $\eta_{mn}$ is the Lorentz metric (a diagonal matrix with
519: constant entries $(-1, +1, +1, \cdots , +1)$). Furthermore,
520: $e_m{}^\mu$ is the matrix inverse of $e_\mu{}^m$, and
521: $\bar\lambda$ is defined by $\l^\dagger i \g^0$. However, in one
522: dimension there are no Lorentz transformations, hence in our toy
523: model $D_\mu \l$ is equal to $\dot\l$. Furthermore, $\det
524: e_\mu{}^m = e_\mu{}^m$ in one dimension, and this cancels the
525: factor $e_m{}^\mu$. Thus even in curved space, the Dirac action
526: in our toy model reduces to ${i \over 2} \l \dot\l$
527: (for real $\l$;
528: the factor $\frac12$ is used for real fields, just as for
529: $\frac12{\dot\varphi}^2$ in~(\ref{1one})). As a consequence, the
530: gravitational stress tensor, which is by definition proportional
531: to ${\delta \over \delta e^m_\mu (x)} S$, vanishes in this model
532: for $\l (t)$. Also the canonical Hamiltonian %
533: \footnote{The momenta are defined by left-differentiation: $p=\tfrac{\partial}{\partial\dot{q}}S$.
534: Hence $p(\phi)=\dot{\phi}$ and $\pi(\lambda)=-\tfrac{i}{2}\lambda$, where $\pi$ denotes the conjugate momentum of $\lambda$.}
535: $H= \dot{q}p-L$
536: vanishes for $L$ given in (2.1) and $q = \l (t)$. This will play a role in the
537: discussions below. The sign of the term ${1 \over 2}
538: \dot\varphi^2$ is positive because it represents the kinetic
539: energy, but the sign of the fermion term could have been chosen
540: to be negative instead of positive. (Requiring $\l$ to be real,
541: we cannot redefine $\l \rightarrow i \l$ in order to change the
542: sign of the second term and still keep real $\l$.) The $+$ sign
543: in (2.1) will lead to the susy anticommutator $\{Q,Q\}=2H$
544: instead of $\{Q,Q\}=-2H$ with a hermitian $Q$. End of physical
545: intermezzo.
546:
547: The supersymmetry transformations should transform bosons into fermions, and vice-versa, so $\varphi$ into
548: $\l$, and $\l$ into $\varphi$. Since
549: $\varphi$ is commuting and $\l$ anticommuting, the parameter must be
550: anticommuting. We take it to be a
551: Grassmann number $\epsilon$, although other choices are also
552: possible.%
553: \footnote{ The author has proposed long ago with J. Schwarz to consider
554: $\theta$'s which satisfy a Clifford algebra,
555: $\{ \theta^\a , \theta^\beta \} = \gamma^{\a\b}_\mu x^\mu$.
556: %One can then introduce not only the super-curvatures
557: %familiar in quantum groups but also super-torsions.
558: %They satisfy a Yang-Baxter algebra (see work with Bouwknegt and McCarthy)
559: .}
560: %\bignote{This footnote looks too special for this exposition. }
561: One might then be tempted to
562: write down $\delta \varphi = i \epsilon \l$ and $\delta \l = \varphi
563: \epsilon$ (where the factor $i$ is
564: needed in order that $\delta \varphi$ be real, taking $\epsilon$ to
565: be real) but this is incorrect as
566: one might discover by trying to prove that the action is invariant
567: under these transformation rules.
568: There is a more fundamental reason why in particular the rule $\delta
569: \l = \varphi \epsilon$ is
570: incorrect, and that has to do with the dimensions of the fields and
571: $\epsilon$ as we now explain.
572:
573: The dimension of an action $S \equiv \int L dt$ is zero (for
574: $\hbar =1$).%
575: \footnote{ More precisely, the dimension of $H$ and $L$ is an energy,
576: and $t$ has of course the dimension of time. In quantum mechanics the Planck constant
577: $\hbar \equiv h/2 \pi$ has the dimension of an energy $\times$ time
578: (discovered by Planck in 1900). Since $S$ has the dimension of
579: an energy $\times$ time, one can define dimensionless exponents
580: of the action by $\exp {i \over \hbar} S$. Such exponents appear
581: in path integrals. Physicists often choose a system of units such that $\hbar =1$.}
582: Hence $L = {1 \over 2} \dot\varphi^2$ should
583: have dimension $+1$, taking the dimension of $t$ to be $-1$ as
584: usual for a coordinate. It follows that the dimension of
585: $\varphi$ is $-1/2$ and that of $\l$ is $0$
586: \eqa
587: [\varphi ] =- 1/2 ;\quad [\l] =0 .
588: \eqae
589: From $\delta \varphi = i \epsilon \l$ we then conclude that
590: $\epsilon$ has dimension $-1/2$
591: \eqa
592: \; [ \epsilon ] =- 1/2 .
593: \eqae
594: Thus $\delta \varphi = i \epsilon \l$ is dimensionally correct: $[
595: \delta \varphi ] =- 1/2$ and $[
596: \epsilon \l] =- 1/2$. Consider now the law for $\delta \l$. The
597: proposal $\delta \l = \varphi
598: \epsilon$ has a gap of one unit of dimension: $[\delta \l ] =0$ but
599: $[ \varphi \epsilon ] = - 1/2 - 1/2
600: =-1$. To fill this gap we can only use a derivative (we are dealing
601: with massless fields so we have no
602: mass available). Thus $\delta \l \sim \dot\varphi \epsilon$. We
603: claim that the correct factor is minus
604: unity, thus
605: \eqa
606: \delta \varphi = i \epsilon \l ; \quad \delta \l = -\dot\varphi \epsilon.
607: \label{1five}
608: \eqae
609: By correct we mean that (\ref{1five}) leaves the action invariant as
610: we now show.
611: It is easy to show that $S$ (rigid) is invariant under these
612: transformation rules if $\epsilon$
613: is constant (rigid supersymmetry). Let us for future purposes
614: already consider a local $\epsilon$
615: (meaning $\epsilon (t)$) and also keep boundary terms due to partial
616: integration. One finds then if one successively varies the fields in $S$ according to (\ref{1five})
617: \eqa
618: \delta S &=& \int \left( \dot\varphi \delta \dot\varphi + {i \over 2} \delta \l \dot\l
619: + {i \over 2} \l \delta \dot\l \right) dt \nn
620: &=& \int \left[ \dot\varphi {d \over dt} (i \epsilon \l) - {i
621: \over 2} (\dot\varphi \epsilon) \dot\l - {i \over 2} \l {d \over
622: dt} ( \dot\varphi \epsilon ) \right] dt \nn &=& \int \left[
623: \dot\varphi i \dot\epsilon \l + \dot\varphi i \epsilon \dot\l -
624: {i \over 2} \dot\varphi \epsilon \dot\l - {i \over 2} {d \over
625: dt} (\l \dot\varphi \epsilon) + {i \over 2} \dot\l \dot\varphi
626: \epsilon \right] dt \nn &=& \int \left[ \dot\epsilon (i
627: \dot\varphi \l ) - {i \over 2} {d \over dt} ( \l \dot\varphi
628: \epsilon ) \right] dt \label{1six} .
629: \eqae
630: We performed a partial
631: integration in the third line and used $\dot\l \epsilon =
632: -\epsilon \dot\l$ in the fourth line. We now assume that
633: ``fields" (and their derivatives) tend to zero at $t= \pm
634: \infty$. (If there would also be a space dimension $\s$, we
635: could consider a finite domain $0 \leq \s \leq \pi$, and then we
636: should specify boundary conditions at $\s = 0, \pi$. This
637: happens in ``open string theory".) It is clear that neglecting
638: boundary terms at $t= \pm \infty$, and taking $\epsilon$ constant
639: ($\dot\epsilon =0$), the action is invariant. (A weaker
640: condition which achieves the same result is to require that the
641: fields at $t \rightarrow + \infty$ are equal to the fields at $t
642: \rightarrow - \infty$). This assumption that fields vanish at
643: $t= \pm \infty$ is not at all easy to justify, but we just accept
644: it.
645:
646: The algebra of rigid supersymmetry transformations reveals that
647: supersymmetry is a square root of
648: translations, in the sense that two susy tranformations (more
649: precisely, a commutator) produce a
650: translation. On $\varphi$ this is clear
651: \eqa
652: && [ \delta (\epsilon_2) , \delta (\epsilon_1) ] \varphi =
653: \delta (\epsilon_2) i \epsilon_1 \l - \delta (\epsilon_1) i \epsilon_2 \l
654: % 1 \leftrightarrow 2
655: \nn
656: && = i \epsilon_1 (-\dot\varphi \epsilon_2 )- i \epsilon_2 (-\dot\varphi \epsilon_1 ) =
657: (2 i \epsilon_2 \epsilon_1 ) \dot\varphi.
658: \eqae
659: We recall that the symbol $[A,B]$ is defined by $AB-BA$, so $[ \delta (\e_2) , \delta (\e_1) ]$ is a commutator of two supersymmetry transformations.
660: We used in the last step that $\epsilon_1 \epsilon_2 = - \epsilon_2 \epsilon_1$.
661: The result is a translation $(\dot\varphi)$ over a distance $\xi =
662: 2i \epsilon_2 \epsilon_1$. The same result is obtained for $\l$
663: \eqa
664: && [ \delta (\epsilon_2 ), \delta (\epsilon_1) ] \l =
665: -\delta (\epsilon_2 ) \dot\varphi \epsilon_1 + \delta (\epsilon_1 ) \dot\varphi \epsilon_2
666: % 1 \leftrightarrow 2 =
667: \nn
668: && = -{d \over dt} ( i \epsilon_2 \l) \epsilon_1 + {d \over dt} ( i \epsilon_1 \l) \epsilon_2
669: = ( 2 i \epsilon_2 \epsilon_1 ) \dot\l.
670: \eqae
671: We used that $\epsilon_2$ is constant, so ${d \over dt} \epsilon_2 =0$, and $\l \e_1 = - \e_1 \l$.
672:
673: In higher dimensional theories this commutator on a fermion yields in addition to a translation also
674: a term proportional to the field equation of the fermion, and to
675: eliminate this extra term with the field equation, one introduces
676: auxiliary fields. (Auxiliary fields are fields which appear in the action without derivatives; they are usually bosonic fields which enter as $a (t)^2$). Here, however, the translation $\dot\l$
677: and the field equation of
678: $\l$ are both equal to $\dot\l$. So the result could still have been a sum of the same
679: translation as on $\varphi$, and a field
680: equation, because both are proportional to $\dot\l$. This is not the
681: case: the coefficient of $\dot\l$ is the same as the coefficient
682: of $\dot\varphi$. There is a simple counting argument that explains
683: this and that shows that no auxiliary
684: fields are needed in this model. Off-shell (by which physicists mean: when the field equations are not satisfied) the translation operator
685: is invertible (the kernel of ${\del \over \del t}$ with the boundary conditions mentioned before is empty), hence the commutator
686: $[\delta (\epsilon_2), \delta (\epsilon_1)]$ cannot vanish on field
687: components. It follows that under rigid
688: supersymmetry if ``the algebra closes" (meaning if $[ \delta (\epsilon_2 ),
689: \delta (\epsilon_1 )]$ is uniformly equal to only a translation but
690: no further field equations), each bosonic field
691: component must be mapped into a fermionic
692: one, and vice-versa. Then the number of bosonic field components
693: must be equal to the number of
694: fermionic field equations. In our toy model there is one bosonic
695: field component $(\varphi)$ and one
696: fermionic field component $(\l)$. Thus there are no auxiliary fields
697: needed in this model.%
698: \footnote{ In the 4-dimensional Wess-Zumino model there
699: are 2 propagating real scalars (A and B) and
700: a real 4-component spinor. Hence there one needs two real bosonic
701: auxiliary fields (F and G). In the 2-dimensional heterotic string the
702: right-handed spinors $\l_R$ do not transform under rigid supersymmetry, $Q \l_R =0$.
703: It follows that on the right-hand side of the susy commutator evaluated on
704: $\l_R$ the field equation $(\dot\l)$ exactly cancels the translation $P \l = \dot\l$.}
705:
706: We can construct charges $Q$ and $H$ which produce susy and
707: time-translation transformations. This requires
708: equal-time Poisson brackets for $\varphi$, and Dirac brackets for
709: $\l$, which become at the quantum level
710: commutators and anticommutators. For $\varphi$ these results are
711: standard: the conjugate momentum $p$ of $\varphi$ is defined by $p = {\partial \over \partial \dot\varphi} S$ and this yields $p = \dot\varphi$. The quantum commutator is given by
712: \eqa
713: p = \dot\varphi;\quad [p (t) , \varphi (t) ] = {\hbar \over i}.
714: \eqae
715: For $\l$ the conjugate momentum is (we use left-derivatives) $\pi =
716: {\del \over \del \dot\l} S = - {i
717: \over 2} \l$. The relation $\pi =- {i \over 2} \l$ is a constraint between the coordinates and the conjugate momenta, called by Dirac a primary constraint
718: \eqa
719: \Phi = \pi + {i \over 2} \l =0.
720: \eqae
721: The naive Hamiltonian is $H_L = \dot{Q} P -L = \dot{q} p- {1 \over 2}
722: \dot{q}^2 + \dot\l \pi - {i \over 2}
723: \l \dot\l = {1 \over 2} p^2$. Here $Q$ and $P$ denote the total set of fields and their canonically conjugates. (We must put $\dot \l$ in front of
724: $\pi$, if we define $\pi$ by left-differentiation,
725: $\pi = {\del \over \del \dot\l} L$, because only then
726: $H_L$ is independent of $\dot{q}$ and $\dot \l$. Namely $\delta
727: H_L$ contains no terms with $\delta \dot{Q}$ but only with
728: $\delta Q$ and $\delta P$.) According to Dirac, one must then
729: consider the naive Hamiltonian plus all possible primary
730: constraints \eqa H = {1 \over 2} p^2 + \a \left( \pi + {i \over
731: 2} \l \right), \eqae where $\a (t)$ is an arbitrary anticommuting
732: parameter. Requiring that the constraint $\pi + {i \over 2} \l =
733: 0$ be maintained in time requires $[H, \pi + {i \over 2} \l ] =0$
734: modulo the constraints, which is indicated by the symbol~$\approx$
735: \eqa \left[ H, \pi + {i \over 2} \l \right]_P \approx 0. \eqae The
736: subscript $P$ indicates that we use here Poisson brackets. We
737: define the Poisson bracket by \eqa \{ f (p, q) , g (p , q) \}_P =
738: - \partial f / \partial p {\partial \over \partial q} g + (-)^\s
739: \partial g / \partial p {\partial \over \partial q} f, \eqae where
740: $\s = + 1$ except when both $f$ and $g$ are anticommuting, in
741: which case $\s =- 1$. The basic relations are $\{ p , q \}_P
742: =-1$ and $\{\pi,\l\}_P =-1$. Of course ${1 \over 2} p^2$ commutes
743: with $\pi+ {i \over 2} \l$, but \eqa \left\{ \Phi, \Phi
744: \right\}_P = \left\{ \pi + {i \over 2} \l, \pi + {i \over 2} \l
745: \right\}_P = - {i \over 2} - {i \over 2} =- i. \eqae Note that the
746: Poisson bracket $\left\{ p, q \right\}_P$ is $-1$ for bosons and
747: fermions alike.%
748: \footnote{ In quantum mechanics the sign of the quantum commutator $[p,q] = - i \hbar$ or the
749: quantum anticommutator $\{ \pi , \l \} = - i \hbar$ is not a
750: matter of convention but follows from the compatibility of the
751: field equations with the Heisenberg equations. For example, for a
752: Dirac spinor $\psi$ with mass $m$ one has $L=i\psi^\dagger
753: \dot\psi+ m\psi^\dagger \psi$ and $\pi=-i\psi^\dagger$. For
754: $\psi$ the field equation is $i\dot\psi+m\psi=0$, and the
755: Heisenberg equation is $\dot\psi={i \over \hbar} [H,\psi]$ with
756: $H=-m\psi^\dagger \psi$. Compatibility of the field equation with
757: the Heisenberg equation requires
758: $\{\psi,\pi\}=\{\psi,-i\psi^\dagger\}=-i\hbar$ which agrees with
759: $\{\pi,\psi\}={\hbar \over i}$ in quantum brackets.} It follows
760: that $[H, \pi + {i \over 2} \l ] = - i \a$, and hence $\a =0$.
761: Thus, with $\a =0$, there are no further (secondary) constraints,
762: and we have \eqa H= {1 \over 2} p^2. \eqae
763:
764: Whenever a set of constraints $\phi^\a$ satisfies $\{ \phi^\a ,
765: \phi^\beta \} = M^{\a\b}$ with ${\rm sdet} M^{\a\b} \not= 0$, we
766: call these constraints second class
767: constraints.%
768: \footnote{ The expression ${\rm sdet} M$ denotes the superdeterminant of a supermatrix
769: $M = \left( \begin{array}{cc} A & B \\ C & D \end{array} \right)$,
770: where $A$ and $D$ contain commuting entries and $B$ and $C$
771: anticommuting entries. Any matrix can be written as the product
772: of diagonal matrices
773: $\left( \begin{array}{cc} A & 0 \\ 0 & B
774: \end{array} \right)$ and triangular matrices $\left(
775: \begin{array}{cc} I & C \\ 0 & I \end{array} \right)$ and $\left(
776: \begin{array}{cc} I & 0 \\ D & I \end{array} \right)$. Namely, $\left(
777: \begin{array}{cc} A & B \\ C & D \end{array} \right)=\left(
778: \begin{array}{cc} I & BD^{-1} \\ 0 & I \end{array} \right) \left(
779: \begin{array}{cc} A-BD^{-1}C & 0 \\ 0 & D \end{array} \right)\left(
780: \begin{array}{cc} I & 0 \\ D^{-1}C & I \end{array} \right)$. The
781: superdeterminant of the product of supermatrices is the product
782: of the superdeterminants of these supermatrices, and ${\rm sdet} \left(
783: \begin{array}{cc} A & 0 \\ 0 & D \end{array} \right) = \det A/
784: \det D$ while ${\rm sdet} \left( \begin{array}{cc} I & B \\ 0 & I
785: \end{array} \right) =1$. Hence ${\rm sdet} M =\det (A-BD^{-1}C)/\det D $. }
786: It follows that $\phi = \pi + {i \over
787: 2} \l$ is a second class constraint. The Dirac bracket is
788: defined by \eqa \left\{ A,B \right\}_D = \left\{ A,B \right\}_P -
789: \left\{ A, \Phi \right\}_P \left\{ \Phi, \Phi \right\}^{-1}_P
790: \left\{ \Phi, B \right\}_P \label{1bracket}, \eqae where $\{ A,B
791: \}_P$ denotes the Poisson bracket. Its definition is chosen such
792: that $\{A,\Phi\}_D=0$ for any $A$ and any second-class constraint
793: $\Phi$. Since in our toy model $\{ \Phi, \Phi \} =-i$, we find
794: \eqa \{ A,B \}_D = \{ A,B \}_P - i \{ A, \pi+ {i \over 2} \l \}_P
795: \{ \pi + {i \over 2} \l, B \}_P \label{1sixteen}. \eqae
796:
797: We can now compute the basic equal-time Dirac brackets
798: \eqa
799: && \{ \l (t), \l (t) \}_D =0 - i (-1) (-1) =-i, \nn
800: && \{ \pi (t) , \l (t) \}_D = -1 - i \left( {-i \over 2} \right) (-1)
801: =- {1 \over 2}, \nn
802: && \{ \pi (t), \pi (t) \}_D =0 - i \left( {-i \over 2} \right) \left(
803: {-i \over 2} \right) = {i \over 4}.
804: \eqae
805: Recalling that $\pi =- {i \over 2} \l$, we see that these relations
806: are consistent: we may replace $\pi$ by $-{i \over 2} \l$ on the
807: left-hand side. At the quantum level, as first proposed by Dirac, we add a factor $i \hbar$ to
808: the Poisson brackets to obtain the quantum (anti) commutators. Hence
809: \begin{align} \label{1varphi} \{ \l (t), \l (t) \} & = \hbar,\nonumber \\
810: [ p (t) , \varphi (t) ] & = {\hbar \over i}.
811: \end{align}
812:
813: We now construct the susy charge $Q$ as a Noether charge. A Noether charge is the space integral of the time component of the Noether current, but since there is no space in our toy model, the Noether current is equal to the Noether charge. We want to obtain an expression for the Noether charge in terms of $p$'s and $q$'s, and therefore we rewrite (\ref{1one}) in Hamiltonian form, namely as $L = \dot{q} p-H$ where $H$ depends only on $p, \pi, \varphi, \l$ but not on their time derivatives. Since, as we shall discuss, the terms proportional to a derivative of the symmetry parameter yield the Noether current, the latter will only be a function of $p$'s and $q$'s but not of derivatives of $p$'s and $q$'s.
814:
815: The action in Hamiltonian form reads \eqa\label{achamil} L=
816: \dot\varphi p+ \dot\l \pi - {1 \over 2} p^2. \eqae where we took
817: the Dirac Hamiltonian $H = {1 \over 2} p^2$ as discussed above.
818: This action is invariant under
819: \begin{gather} \label{1twenty} \delta \varphi = {i \over 2}
820: \epsilon \l - \epsilon \pi, \quad \delta \l = -p \epsilon, \nonumber \\
821: \delta p =0, \quad \delta \pi = {i \over 2} p \epsilon.
822: \end{gather}
823: These rules follow by requiring
824: invariance of the action, but one can also derive them by adding
825: equation of motion symmetries to the original rules. For example,
826: $p=\dot \varphi$ leads to $\delta p={i \over 2}\epsilon
827: \dot\l-\epsilon\dot\pi$, and to remove $\dot\l$ one may add
828: $\delta({\mathrm extra}) p={i \over 2}\epsilon{\del \over \del \pi}S$
829: and $\delta({\mathrm extra}) \pi=-{i \over 2}{\delta S \over \delta
830: p} \epsilon$. These extra transformation rules form a separate
831: symmetry of any action, so we may add them to the original
832: rules. Then $\delta p=-\epsilon \dot\pi$, and also $\delta
833: \pi=\delta (-{i \over 2}\l)={i \over 2}\dot\varphi \epsilon$ is
834: modified into $\delta \pi={i \over 2}\dot\varphi \epsilon-{i
835: \over 2}(\dot\varphi-p)\epsilon={i \over 2}p\epsilon$. To also
836: remove the term
837: $-\epsilon\dot\pi$ in $\delta p$, one adds another equation of motion symmetry:
838: $\delta(\mathrm{extra})p = {\delta \over \delta \l} S \e$ and
839: $\delta(\mathrm{extra})\l=\tfrac{\partial}{\partial p}S$. The final result
840: is~(2.22).
841:
842: The standard way of obtaining the Noether charge follows from
843: letting the rigid parameter become local and collecting terms
844: proportional to $\dot\e$. The terms with $\dot\epsilon$ in
845: $\delta L$ for local $\epsilon (t)$ are contained in \eqa \delta
846: L &=& p {d \over dt} \left( {i \epsilon \over 2} \l - \epsilon
847: \pi \right) - \pi {d \over dt} (-p \epsilon ) + \dot\l \left( { i
848: \over 2} p \epsilon \right) \nn &=& {i \over 2} p \dot\epsilon \l
849: - p \dot\epsilon \pi + {d \over dt} ( \pi p \epsilon).
850: \label{1twentyone} \eqae So, defining $Q$ as the coefficient of
851: $i\dot \epsilon$, we find \eqa Q = p \left( {1 \over 2} \l +i \pi
852: \right). \label{1twentytwo} \eqae This reproduces (2.22);
853: for example, $\tfrac{1}{\hbar}[\varphi, \epsilon Q]=\delta \varphi$ follows
854: from~(\ref{1varphi}). Using the constraint $\pi =-{i \over 2} \l$
855: we see that $Q$ becomes equal to $Q = p \l$, and (2.22)
856: reduces to (\ref{1five}). However, if one uses (anti) commutators
857: one needs a Hamiltonian treatment, and then one needs
858: (\ref{1twentytwo}).
859:
860: The charge $Q$ which appears in brackets such as $[ \varphi, \e Q
861: ]$ is clearly an operator, so $Q$ in (\ref{1twentytwo}) is an
862: operator expressed in terms of Heisenberg fields. The latter
863: satisfy their own equations of motion. On the other hand, in the
864: action the fields are off-shell. So, in principle one might need
865: extra terms proportional to the equations of motion to obtain the
866: correct off-shell transformations. In this case we do not need
867: such terms. In section 6 we shall discuss the Hamiltonian
868: approach with off-shell fields. This is a very general approach
869: which yields the action in Hamiltonian form and the quantum BRST
870: charge, starting only from the set of first class constraints.
871:
872: The other way of obtaining $Q$ (more precisely, $i\epsilon Q$) is to write
873: it as a sum of terms of
874: the form $\delta \varphi p$ for all fields, plus $-K$ where
875: $\delta L = {d \over dt} K$. From ({\ref{1twentyone}) we read off
876: that $K = \pi p \epsilon$. Hence
877: \begin{align}
878: i \epsilon Q & = \delta \varphi p + \delta \l \pi - K = \left( {i
879: \over 2} \epsilon \l - \epsilon \pi
880: \right) p - p \epsilon \pi - \pi p \epsilon \nonumber \\
881: & = i \epsilon \left( {1 \over 2} \l + i \pi \right) p,
882: \end{align}
883: which is indeed the same result as in (\ref{1twentytwo}).
884:
885: The supersymmetry algebra (rather a superalgebra with commutators
886: and anticommutators) is now easy to evaluate. Using the quantum
887: brackets of (\ref{1varphi}) finds \eqa && \{ Q, Q \} = \left\{
888: \left( { 1\over 2} \l + i\pi \right) p , \left( {1\over 2} \l +
889: i\pi \right) p \right\} = \{ \l p, \l p \} =\hbar p^2 =2\hbar H,
890: \nn && [H, Q] =0 \quad \text{(via Jacobi, or directly)}.
891: \label{Jacobi} \eqae Thus the generators $Q$ and $H$ form a
892: closed superalgebra, and supersymmetry is the square root of (the
893: generator of time-) translations.
894:
895: Dirac was the first to take the square root of the Laplace
896: operator $\Box$, and this led to the famous Dirac equation of
897: 1927. This equation led to the prediction that for every
898: fermionic particle there is a fermionic antiparticle. These
899: antiparticles have been found in the laboratories. Likewise, the
900: square root $Q\sim \sqrt{H}$ predicts that for every particle there
901: should be a superpartner. Not a single superpartner has been
902: found so far, but that may change.
903:
904:
905: \mathversion{bold}
906: \section{$N=1$ supergravity in $x$-space}
907: \mathversion{normal}
908:
909: Having discussed the rigid supersymmetry (= susy) of the action
910: $S \; {\rm (rigid)} \; = \int L dt$ with $L= {1 \over 2}
911: \dot\varphi^2 + {i \over 2} \l \dot\l$, we turn to local susy. We
912: let $\epsilon$ become time-dependent and find then (see
913: (\ref{1six})). \eqa \delta S \; {(\rm rigid)} \; =
914: \int\limits^\infty_{-\infty} \dot\epsilon (i \dot\varphi \l) dt.
915: \label{1seee} \eqae The boundary terms at $t = \pm \infty$ vanish
916: if we require that $\e (t)$ vanishes at $t = \pm \infty$. To
917: cancel this variation we introduce the gauge field for local
918: susy, the gravitino $\psi (t)$. The transformation rule of a
919: gauge field begins always with a derivative of the local
920: parameter. We then couple $\psi$ to the Noether current of rigid
921: supersymmetry, using $\delta \psi = \dot\epsilon + \cdots$ to fix
922: the overall constant of this new term \eqa S \; {\rm (Noether)}
923: \; = \int\limits^\infty_{-\infty} (- i \psi \dot\varphi \l) dt ;
924: \delta \psi = \dot\epsilon + \cdots \label{1already} \eqae If we
925: vary $\psi$ in (\ref{1already}), then the variation $\delta S$
926: (Noether) cancels $\delta S$ (rigid), but the fields
927: $\dot\varphi$ and $\l$ in $S$ (Noether) must also be varied. This
928: yields two further variations \eqa \delta S \; {\rm (Noether)} \;
929: = \int\limits^\infty_{-\infty} \left[ - i \psi \left\{ {d \over
930: dt} (i \epsilon \l ) \right\} \l + i \psi \dot\varphi \epsilon
931: \dot\varphi \right] dt. \eqae
932: %Since $\l \l =0$, the derivative ${d \over dt}$ must cut on $\epsilon$. Then
933: In the first variation the ${d \over dt}$ must hit the field $\l$
934: because otherwise one would be left with $\l\l$ which vanishes.
935: Hence the remaining variations to be canceled are \eqa \delta S
936: \; {\rm (Noether)} \; = \int\limits^\infty_{-\infty} i \psi
937: (\dot\varphi \dot\varphi + i \l \dot\l ) \epsilon dt.
938: \label{1twentyseven} \eqae The last term can be canceled by
939: adding a new term in $\delta \l$ (because this variation is
940: porportional to the field equation of $\l$). However, the first
941: term can only be canceled by introducing a new field $h$ (the
942: graviton) and coupling it to $\dot\varphi \dot\varphi$. Thus the
943: coupling of rigidly supersymmetric matter to the supergravity
944: gauge fields requires for consistency (invariance of the whole
945: action under local susy) also the coupling to gravity. \emph{
946: Local susy is a theory of gravity, and this explains the name
947: supergravity.}
948:
949: There appears, however, an ambiguity at this point: we can also
950: couple this new field $h$ to $-i \l \dot\l$, and the most general
951: case is a linear combinations of both possibilities. Hence we add
952: \begin{align} \label{1however} S\,{\rm (stress)} & = - \int\limits^\infty_{-\infty} h
953: [ \dot\varphi \dot\varphi - i \l \dot\l x ] dt, \nonumber \\ \delta h & =
954: -i \epsilon \psi, \quad \delta \l =- \dot\varphi \epsilon +i (1 +x)
955: \psi \l \epsilon, \end{align} where $x$ is a free
956: constant real parameter. We must now evaluate the old variations
957: in the new action, the new variations in the old action, and the
958: new variations in the new action. The aim is to use these
959: variations to cancel~(\ref{1twentyseven}).
960:
961: The new variation $\delta h =- i \epsilon \psi$ in the new action $- \int h \dot\varphi
962: \dot\varphi dt$ cancels the first term in
963: (\ref{1twentyseven}). The new variation $\delta \l = i (1+x) \psi \l
964: \epsilon$ in the old action $ {i \over 2} \l \dot\l$ in (\ref{1one}) and the
965: new variation $\delta h =-i \epsilon \psi$ in the new action $S$ (stress) cancel the
966: variation $- \psi \l \dot\l \epsilon$ of $\delta S$ (Noether)
967: \begin{gather} i \l {d \over dt} [ i (1+x) \psi \l \epsilon ] - i \epsilon \psi
968: (i \l \dot\l x) - \psi \l \dot\l \epsilon \nonumber \\
969: = (1 +x) \l \dot\l \psi \epsilon - \psi \epsilon \l \dot\l x- \psi
970: \epsilon \l \dot\l = 0.
971: \end{gather}
972: (We partially integrated the first term, and the last term comes from (\ref{1twentyseven})).
973: We find thus at this moment a free parameter $x$ in the action and
974: transformation rules; this frequently
975: happens in the construction of supergravity models, and usually these
976: parameters get fixed at a later stage, or they can be removed by field redefinitions. We demonstrate this later explicitly in our toy model.
977:
978: We are again in the same situation as before: we canceled $\delta
979: S$ (Noether) by introducing a new term in the action, namely $S$
980: (stress). We already took into account the variation of the
981: gauge field $h$ in this new term, but we must still vary the
982: matter fields in $S$ (stress). We first set $x=0$ and later
983: consider the case $x \not= 0$. If $x=0$, we need only vary the
984: $\dot\varphi$ in $S$ (stress) and this yields
985: \begin{align} \label{1streee} \delta L\,
986: {\rm (stress)} = - 2h \dot\varphi {d \over dt} (i \epsilon \l) = -2 h \dot\varphi i \dot\epsilon \l -2 h \dot\varphi i
987: \epsilon \dot\l.
988: \end{align} The first term is
989: proportional to the Noether current $\dot\varphi \l$ in
990: (3.2) and can thus be canceled by a new term in the
991: gravitino law \eqa \delta \; {\rm (new)} \; \psi =- 2 h
992: \dot\epsilon. \eqae Substituting this vatiation into
993: (\ref{1already}), the first term in (\ref{1streee}) is
994: cancelled. The second term in (\ref{1streee}) is proportional
995: to the free field equation of $\l$ and can be canceled by adding
996: a new term to the transformation law of $\l , \delta \l = 2 h
997: \dot\varphi \epsilon$, because then $\delta ({i \over 2}\l
998: \dot\l)=-i\dot\l (2h\dot\varphi \epsilon)$ cancels the second
999: term in (\ref{1streee}).
1000:
1001: The new transformation law $\delta \l = 2 h \dot\varphi \epsilon$
1002: produces a new variation in the Noether action (\ref{1already})
1003: \eqa
1004: \delta L\,(\text{Noether due to} \; \delta \l = 2 h
1005: \dot\varphi \epsilon) =- i \psi \dot\varphi (2 h \dot\varphi
1006: \epsilon).
1007: \label{1Noether}
1008: \eqae
1009: Since this term is proportional to $\dot\varphi \dot\varphi$ it can
1010: be canceled by a final extra term in
1011: $\delta h$, namely $\delta h = 2i h \epsilon \psi$. Then the new variation of
1012: $h$ used in (\ref{1however}) cancels (\ref{1Noether}).
1013:
1014: Because each time when we replace an $\dot\epsilon$ in a variation by $\psi$ we loose a time derivative,
1015: this process of adding further terms to the action and transformation laws is guaranteed to stop.
1016: Of course it is not guaranteed that an invariant action exists.
1017: Examples are known where this process does not yield an invariant action, for example adding a cosmological constant to supergravity in $10 + 1$ dimensions.
1018:
1019: We have now canceled all variations for the case $x=0$, hence we have
1020: constructed a locally susy action.
1021: The final results read
1022: \begin{align} \label{1thirtythree}
1023: L & = {1 \over 2} \dot\varphi^2 + {i \over 2} \l \dot\l - i \psi
1024: \dot\varphi \l - h
1025: \dot\varphi^2, \nonumber \\
1026: \delta \varphi & = i \epsilon \l, \quad \delta \l = -\dot\varphi \epsilon
1027: + i \psi \l \epsilon +2 h
1028: \dot\varphi \epsilon, \nonumber\\
1029: \delta \psi & = \dot\epsilon - 2 h \dot\epsilon,\quad \delta h = -i
1030: \epsilon \psi +2 i h \epsilon \psi.
1031: \end{align}
1032:
1033: Before going on, we make three comments.
1034:
1035: \medskip
1036: \noindent 1) There is no gauge action for gravity or local
1037: supersymmetry in one dimension as one might expect, since the scalar
1038: curvature $R$ and its linearization, the Fierz-Pauli action,%
1039: \footnote{The linearized form of the Einstein-Hilbert action $e R$ is called the
1040: Fierz-Pauli action and is given in n-dimensions by
1041: \[
1042: L = - {1 \over 2} \varphi^2_{\mu\nu , \l} + \varphi_\mu^2 -
1043: \varphi_\mu \varphi_{, \mu} + {1 \over 2} \varphi_{, \mu}^2,
1044: \]
1045: where $\varphi_\mu = \del^\nu \varphi_{\mu\nu}$, $\varphi=\eta^{\mu\nu}
1046: \varphi_{\mu\nu}$ and the metric $g_{\mu\nu} \equiv \eta_{\mu\nu} +
1047: \kappa h_{\mu\nu}$ is related to $\varphi_{\mu\nu}$ by
1048: $(\sqrt{-g} g^{\mu\nu})_{\rm lin} - \eta^{\mu\nu} =- \kappa
1049: (h^{\mu\nu}- {1 \over 2} \eta^{\mu\nu} h)=\varphi^{\mu\nu}$.
1050: In one dimension the first term in $L$ cancels the last term, and the second term cancels the third term.}
1051: vanishes in one dimension. (Also the gravitino gauge action vanishes in one and $1+1$ dimensions.
1052: The gravitational action is a total derivative in $1+1$ dimensions, where it yields the Euler invariant). A gauge action for supergravity in one dimension
1053: would have to start with $L = {1 \over 2} \dot{h} \dot{h} + {i \over 2}
1054: \psi \dot\psi$ and it is indeed
1055: invariant under the rigid symmetries $\delta h = i
1056: \epsilon \psi, \delta \psi = -\dot{h} \epsilon$, see (\ref{1one}) and
1057: (\ref{1five}). However, for local $\epsilon (t)$ the
1058: rules were already fixed by the matter coupling, see
1059: (\ref{1thirtythree}), and these rules do not leave this action
1060: invariant.
1061:
1062: \medskip
1063: \noindent 2) The term $\int g \varphi^n dt$ with $g$ a coupling
1064: constant cannot be made supersymmetric. Yukawa couplings
1065: do not exist in this model because $\l \varphi \l$ vanishes.
1066: However, one can make $\l$ a complex Dirac spinor and then
1067: supersymmetric interactions exist. One can also supersymmetrize a term $f
1068: ( \varphi ) \dot\varphi
1069: \dot\varphi$; the action becomes $f (\varphi ) ( \dot\varphi
1070: \dot\varphi + i \l \dot\l)$ and is called a
1071: susy nonlinear $\s$ model because $f(\varphi)$ can be nonlinear, for
1072: example $\exp \varphi$.
1073:
1074: \medskip
1075: \noindent 3) One can also couple the first-order action
1076: in~(\ref{achamil}) to supergravity. Denoting the graviton and
1077: gravitino fields by $H$ and $\Psi$, the result is \eqa &&
1078: L=\dot\varphi p + \dot\l \pi - {1 \over 2}p^2 - i \Psi \left( {1
1079: \over 2}p\l + i p \pi \right) - Hp^2, \nn && \delta p=0,\quad \delta
1080: \pi={i \over 2} p \epsilon,\quad \delta \l=-p \epsilon, \nn && \delta
1081: \varphi={i \over 2}\epsilon \l - \epsilon \pi,\quad \delta
1082: \Psi=\dot\epsilon,\quad \delta H=-i\epsilon \Psi. \label{1Gaussian}
1083: \eqae Note that $\delta (\pi+{i \over 2}\l)=0$ agrees with the
1084: constraint $\pi+{i \over 2}\l=0$. So we may replace $\pi$ by $-{i
1085: \over 2} \l$ in the action and transformation rules. Furthermore
1086: we can eliminate $p$ by integrating in the path integral over a
1087: Gaussian with $p^2$ \cite{rfeynman}. The result of these
1088: manipulations is the following action \eqa L = {1 \over 2} {1
1089: \over 1+2H} \dot\varphi^2 - {i \over 1+2H} \dot\varphi \Psi \l +
1090: {i \over 2} \l \dot\l. \eqae Comparison with the second-order
1091: action in (\ref{1thirtythree}) (the action without conjugate
1092: momenta) we can read off how $H$ is related to $h$, and $\Psi$ to
1093: $\psi$. \eqa\label{dvatr} H= {h \over 1-2h} , \;\; \Psi \; = {1
1094: \over 1-2h} \psi , \;\; {\rm or} \;\; h = {H \over 1+2H} , \;\; \psi
1095: \; = {1 \over 1+2H} \Psi. \eqae The Jacobian for the change of
1096: variables from $(H,\Psi)$ to $(h,\psi)$ is ${1 \over
1097: 1-2h}=1+2H$. (One needs a super Jacobian, in particular $\del
1098: \delta \psi / \del \psi$ is equal to $1-2 h$, and not simply
1099: equal to $(1-2 h)^{-1}$). The transformation rules in
1100: (\ref{1Gaussian}) go over into (\ref{1thirtythree}) if one uses
1101: these redefinitions.
1102:
1103: For physicists: The Jacobian $J = 1/1-2h$ can be exponentiated using a new kind of ghosts, introduced by Bastianelli and the author and playing the same role as the Faddeev-Popov ghosts. In order that the theories with $h$ and $\psi$ and $H$ and $\Psi$ are equivalent at the quantum level (by which we mean that they should give the same Feynman graphs) one needs those new ghosts. The propagators of $h$ and $\psi$ come from the gauge fixing terms.
1104:
1105: We now return to the model in (\ref{1thirtythree}) and evaluate the local susy algebra.
1106: On $\varphi$ one finds
1107: \eqa
1108: [ \delta (\epsilon_2), \delta (\epsilon_1) ] \varphi &=&
1109: i \epsilon_1 (-\dot\varphi \epsilon_2 + i \psi \l \epsilon_2 + 2 h \dot\varphi \epsilon_2) -
1110: i \epsilon_2 (-\dot\varphi \epsilon_1 + i \psi \l \epsilon_1 + 2 h \dot\varphi \epsilon_1)
1111: % 1 \leftrightarrow 2
1112: \nn
1113: &=& [ 2 i (1-2 h) \epsilon_2 \epsilon_1 ] \dot\varphi + i [-2 i
1114: \epsilon_2 \epsilon_1 \psi ] \l.
1115: \eqae
1116: The right-hand side contains a general coordinate transformation
1117: $\delta \varphi = \hat\xi \dot\varphi$ with
1118: $\hat\xi = 2i (1-2h) \epsilon_2 \epsilon_1$; this is clearly the
1119: gravitational extension of the
1120: nongravitational rigid translation with parameter $\xi = 2i \epsilon_2 \epsilon_1$ which we
1121: found in the rigid susy commutator.
1122: The second term is a local susy transformation $i \hat{\epsilon} \l$ of $\varphi$ with
1123: parameter $\hat\epsilon = - 2 i \epsilon_2 \epsilon_1 \psi$. Note that the composite parameters $\hat\xi$ and $\hat\epsilon$ are field- dependent. The structure constants are no longer constants! This has led to a new development in group theory.
1124:
1125: On $\l$ one finds after somewhat lengthy algebra
1126: \eqa
1127: [ \delta ( \epsilon_2 ), \delta (\epsilon_1) ] \l = \hat\xi \dot\l -
1128: \dot\varphi \hat\epsilon +2 h
1129: \dot\varphi \hat\epsilon.
1130: \eqae
1131: and the terms with $\hat\epsilon$ agree with (\ref{1thirtythree}) (because $i \psi \l \hat\epsilon =0$ due to $\psi \psi =0$).
1132: The term with $\hat\xi$ constitutes a general coordinate transformation on $\l$.
1133: Clearly, the same algebra is found on $\l$ as on $\psi$!
1134:
1135: On $\psi$ one finds
1136: \eqa
1137: [ \delta (\epsilon_2), \delta (\epsilon_1) ] \psi &=&
1138: - 2 (-i \epsilon_2 \psi + 2i h \epsilon_2 \psi ) \dot\epsilon_1 + 2 (-i \epsilon_1 \psi + 2i h \epsilon_1 \psi )\dot\epsilon_2
1139: %- 1 \leftrightarrow 2
1140: \nn
1141: &=& -2i \left[ {d \over dt} (\epsilon_2 \epsilon_1 ) \right] \psi +
1142: 4 i h \left[ {d \over dt}
1143: (\epsilon_2 \epsilon_1 ) \right] \psi \nn
1144: &=& {d \over dt} \hat\epsilon + 2i \epsilon_2 \epsilon_1 \dot\psi
1145: -2h {d \over dt}
1146: \hat\epsilon - 4 ih \epsilon_2 \epsilon_1 \dot\psi \nn
1147: &=& {d \over dt} \hat\epsilon - 2h {d \over dt} \hat\epsilon +
1148: \hat\xi \dot\psi.
1149: \eqae
1150: So also on $\psi$ the same local algebra is realized.
1151:
1152: Finally also on $h$ the same algebra is realized%
1153: \footnote{Here and below ``$1 \leftrightarrow 2$'' means ``switch indices 1 and 2 in the preceding expression.''}
1154: \eqa
1155: && [ \delta ( \epsilon_2), \delta (\epsilon_1)] h =
1156: - i \epsilon_1 ( \dot\epsilon_2 - 2h \dot\epsilon_2 ) - 2i [ \delta (\epsilon_2) h \psi ] \epsilon_1
1157: % + (i \epsilon_2 ( \dot\epsilon_1 - 2h \dot\epsilon_1 ) - 2i [ \delta (\epsilon_1) h \psi ] \epsilon_2)
1158: - (1 \leftrightarrow 2) \nn
1159: && = - i \epsilon_1 (\dot\epsilon_2 - 2h \dot\epsilon_2) - 2i (-i \epsilon_2 \psi +2 i h \epsilon_2 \psi) \psi \epsilon_1 - 2i h
1160: (\dot\epsilon_2 - 2h \dot\epsilon_2) \epsilon_1
1161: % && + ( i \epsilon_2 (\dot\epsilon_1 - 2h \dot\epsilon_1) - 2i (i \epsilon_1 \psi +2 i h \epsilon_1 \psi) \psi \epsilon_2 - 2i h
1162: % (\dot\epsilon_1 - 2h \dot\epsilon_1) \epsilon_2 )
1163: - ( 1 \leftrightarrow 2 )
1164: \nn
1165: && = - i \hat\epsilon \psi +2i h \hat\epsilon \psi + \hat\xi \dot{h} -
1166: \dot{\hat\xi} h + {1 \over 2} \dot{\hat\xi}.
1167: \label{1thirtyseven}
1168: \eqae
1169: The terms with $\hat\epsilon$ clearly agree with (\ref{1thirtythree}).
1170: The terms with $\hat\xi$ in (\ref{1thirtyseven}) yield a general
1171: coordinate transformation of $h$ as we shall
1172: discuss below. Hence, the local susy algebra closes on all fields uniformly.
1173: We can write this as
1174: \begin{equation}\label{moya}
1175: [\delta_s(\epsilon_2),\delta_s(\epsilon_1)]=\delta_{s}(-2i\epsilon_2
1176: \epsilon_1\psi)+\delta_g((1-2h)2i\epsilon_2\epsilon_1).
1177: \end{equation}
1178: The commutator of $\delta_s$ with $\delta_g$, and with itself,
1179: close (they are proportional to $\delta_s$ and $\delta_g$).
1180:
1181: We can understand the closure of the local supersymmetry algebra by using
1182: the same argument as used for the rigid supersymmetry algebra. There is
1183: one bosonic gauge field component ($h$) and one fermionic gauge
1184: field component ($\psi$), hence no auxiliary fields in the gauge sector are needed.
1185:
1186: We now consider the case $x \not= 0$. Here a simple argument
1187: suffices. Rescaling \eqa \l = ( 1 +2 h x)^{1/2} \tilde\l,
1188: \label{1thirtyeight} \eqae we obtain as action from (3.10) \eqa \cl = {1
1189: \over 2} \dot\varphi^2 + {i \over 2} (1 + 2 hx) \tilde\l {d \over
1190: dt} \tilde\l -i \tilde\psi \dot\varphi \tilde\l - h \dot\varphi^2,
1191: \label{action} \eqae where we also rescaled $\psi$ according to
1192: \eqa \psi (1+2 hx)^{1/2} = \tilde\psi. \label{1forty} \eqae We
1193: have produced the action with $x \not= 0$ from the action with
1194: $x=0$ by a simple rescaling of the fields $\l$ and $\psi$. It
1195: follows that this action is also locally susy. The precise susy
1196: transformation rules follow from this rescaling \eqa && \delta \l
1197: = (1+2 hx)^{1/2} \delta \tilde\l + (1+2 hx)^{-1/2} x \delta h
1198: \tilde\l \nn &&\,\,\,\,\,\,\,\, = - \dot\varphi \epsilon + i
1199: \tilde\psi \tilde\l \epsilon +2 h \dot\varphi \epsilon.
1200: \label{1fortyone} \eqae By dividing (\ref{1fortyone}) by $(1+2
1201: hx)^{1/2}$ and using $\delta h = -i \epsilon \psi +2i h \epsilon
1202: \psi$ we find $\delta \tilde\l$. It reads \eqa && \delta \tilde\l
1203: =- \dot\varphi \tilde\epsilon + i \tilde\psi \tilde\l
1204: \tilde\epsilon +2 h \dot\varphi \tilde\epsilon - x (1 + 2
1205: hx)^{-1} (-i \tilde\epsilon \tilde\psi +2i h \tilde\epsilon
1206: \tilde\psi ) \tilde\l, \eqae where we also rescaled $\epsilon$
1207: according to $(1+2 hx)^{-1/2} \epsilon = \tilde\epsilon$. (We
1208: used $\epsilon \psi = \tilde\epsilon \tilde\psi$ in $\delta h$).
1209: By rearranging these terms as \eqa \delta \tilde\l &=& -(1-2h)
1210: \dot\varphi \tilde\epsilon +i \left( 1+ {x-2hx \over 1+2 hx}
1211: \right) \tilde\psi \tilde\l \tilde\epsilon \nn &= & -(1-2h)
1212: \dot\varphi \tilde\epsilon +i \left( {1 +x \over 1+2 hx} \right)
1213: \tilde\psi \tilde\l \tilde\epsilon, \eqae we find a polynomial
1214: result for $x=-1$ \eqa \delta \tilde\l = -(1-2h) \dot\varphi
1215: \tilde\epsilon,\quad \delta \varphi =i \tilde\epsilon (1-2h) \tilde\l
1216: \label{delta}. \eqae For $\delta h$ and $\delta \tilde\psi$ we
1217: find in a similar manner \eqa \delta h &=& -i \tilde\epsilon
1218: \tilde\psi +2i h \tilde\epsilon \tilde\psi, \delta \tilde\psi =
1219: \sqrt{(1-2h)} \delta \psi = \sqrt{(1-2 h)} (1-2h) \dot\epsilon \nn
1220: &=& (1-2 h) [(1-2 h) \dot{\tilde\epsilon} - \dot{h}
1221: \tilde\epsilon ]. \label{polynomial} \eqae We used that $\delta h
1222: \psi$ vanishes. Thus for $x=-1$ all laws become again polynomial.
1223:
1224: We can, in fact, apply the Noether method also directly to gravity.
1225: This will explain the $\hat\xi$
1226: terms in (\ref{1thirtyseven}). Starting with
1227: \eqa
1228: L \; {\rm (rigid)} \; = {1 \over 2} \dot\varphi^2 + {i \over 2} \l \dot\l,
1229: \eqae
1230: we find from the translation symmetry rules $\delta \varphi = \xi
1231: \dot\varphi$ and $\delta \l = \xi
1232: \dot\l$ for local $\xi$
1233: \eqa
1234: \delta S \; {\rm (rigid)} \; = \int \left[ \dot\varphi {d \over dt}
1235: (\xi \dot\varphi) + {i \over 2} \l
1236: {d \over dt} ( \xi \dot\l) + {i \over 2} \xi \dot\l \dot\l \right] dt.
1237: \eqae
1238: The third term vanishes as $\dot\l \dot\l =0$, and the second term
1239: vanishes after partial integration,
1240: whereas the first term yields after partial integration of one-half
1241: of this term (this ${1 \over 2} - {1
1242: \over 2}$ trick cancels the $\ddot\varphi$ terms)
1243: \eqa
1244: \delta S \; {\rm (rigid)} \; = \int \left[ {1 \over 2} \dot\xi
1245: \dot\varphi \dot\varphi + {d \over dt}
1246: \left( {1 \over 2} \xi \dot\varphi \dot\varphi + {i \over 2} \xi \l
1247: \dot\l \right) \right] dt.
1248: \eqae
1249: We see that $\dot\varphi \dot\varphi$ is the Noether current for
1250: translations for the $(\varphi, \l)$ system. (Also the other way to construct the Noether current gives the same result).
1251:
1252: Introducing the gauge field $h$ for gravity and defining $\delta h = {1 \over 2} \dot\xi + \cdots$ we
1253: obtain
1254: \eqa
1255: L \; {\rm (Noether)} \; =- h \dot\varphi \dot\varphi, \quad \delta h
1256: = {1 \over 2} \dot\xi.
1257: \eqae
1258: Note that there is no term of the form $h \l \dot\l$ in $L$
1259: (Noether). (However
1260: the field redefinition (\ref{1thirtyeight}) produces a $h \l \dot\l$ term
1261: in the action in (\ref{action})). As Noether current for $\l$ one might have expected $\pi \dot\l$ where $\pi$ is the conjugate momentum of $\l$, but $\dot\l$ vanishes on-shell, and anyhow time-derivatives of fields should not appear in the Hamiltonian formalism.
1262:
1263: Variation of $\dot\varphi$ in $L$ (Noether) yields
1264: \eqa
1265: \delta S \; {\rm (Noether)} \; = \int - 2 h \dot\varphi {d \over dt}
1266: (\xi \dot\varphi) dt,
1267: \eqae
1268: which can be canceled, using the $\tfrac{1}{2} - \tfrac{1}{2}$ trick, by $\delta h =
1269: \xi \dot{h} - h \dot\xi$. We could also
1270: have proceeded in another way: if we would partially integrate we
1271: would obtain $2 \dot{h} \dot\varphi \xi
1272: \dot\varphi + 2 h \ddot\varphi \xi
1273: \dot\varphi$, and the first variation could be removed by $\delta h =
1274: 2 \xi \dot{h}$ while the second
1275: variation could be eliminated by a suitable $\delta \varphi = 2 \xi h
1276: \dot\varphi$ in ${1 \over 2} \dot\varphi^2$.
1277: However, the variation $\delta \varphi =2 \xi h \dot\varphi$ produces
1278: in $S$ (Noether) with the $\tfrac{1}{2} - \tfrac{1}{2}$ trick a new
1279: variation $-2 \dot\xi h \dot\varphi h \dot\varphi$, which can be
1280: canceled in two ways, etc. All these ambiguities (and more)
1281: are equivalent to field redefinitions.
1282:
1283: We have thus shown that
1284: \eqa
1285: L = {1 \over 2} \dot\varphi^2 + {i \over 2} \l \dot\l - h \dot\varphi^2
1286: \label{over}
1287: \eqae
1288: is invariant under
1289: \eqa
1290: \delta \varphi = \xi \dot\varphi , \delta \l = \xi \dot\l , \delta h
1291: = {1 \over 2} \dot\xi + \xi \dot{h}
1292: - \dot\xi h.
1293: \label{varphi}
1294: \eqae
1295: In particular the result for $\delta h$ shows that the $\hat\xi$
1296: dependent terms in (\ref{1thirtyseven}) are
1297: indeed a general coordinate transformation.
1298:
1299: In higher dimension the coupling of a scalar field to gravity is given by
1300: \[
1301: L=-{1 \over 2}(-\det g)^{1 \over 2}g^{\mu\nu}\del_\mu \varphi
1302: \del_\nu\varphi,
1303: \]
1304: so in one dimension $e_m^\mu$ corresponds to $1-2 h$. Then (\ref{varphi}) agrees with the usual transformation rules of general relativity.
1305:
1306: Adding the coupling to the gravitino $-i \psi \dot\varphi \l$, we
1307: obtain invariance under $\xi$
1308: transformations, provided we choose $\delta \psi$ appropriately,
1309: \eqa
1310: \delta (-i \psi \dot\varphi \l) =- i \delta \psi \dot\varphi \l-i
1311: \psi \left[ {d \over dt} (\xi \dot\varphi ) \right]
1312: \l -i \psi \dot\varphi \xi \dot\l.
1313: \eqae
1314: Partially integrating the last two terms, we obtain
1315: \eqa
1316: -i \delta \psi \dot\varphi \l + i \dot\psi \xi \dot\varphi \l,
1317: \eqae
1318: and these two terms can be canceled by choosing $\delta \psi = \xi
1319: \dot\psi$. In general, Lagrangian
1320: densities in general relativity transform as coordinate densities,
1321: which means that
1322: \eqa
1323: \delta (-i \psi \dot\varphi \l) = {d \over dt} (-i \xi \psi \dot\varphi \l).
1324: \eqae
1325: This is indeed achieved if $\psi$ transforms like
1326: \eqa
1327: \delta \psi = \xi \dot\psi.
1328: \label{dot}
1329: \eqae
1330:
1331: Thus the model in (\ref{1thirtythree}) has now also been shown to be invariant under
1332: general coordinate transformation. The result in (\ref{dot}) can be explained
1333: by noting that the field $\psi$ in the Noether coupling in (\ref{1already}) corresponds to
1334: $(\det g)^{1 \over 2} g^{\mu\nu} \psi_\nu \sim e_m{}^\mu \psi_\mu$ in one
1335: dimension. This field $e^\mu_m \psi_\mu$ is a coordinate scalar, in agreement with (\ref{dot}).
1336:
1337: Let us again rescale $\l = \sqrt{1+2 hx} \tilde\l$ to obtain an
1338: action with a term $h \tilde\l \dot{\tilde\l}$. We obtain from (\ref{over})
1339: \eqa
1340: && L = {1 \over 2} \dot\varphi^2 + {i \over 2} \tilde\l
1341: \dot{\tilde\l} -h \left( \dot\varphi^2 - i x \tilde\l {d
1342: \over dt} \tilde\l \right), \nn
1343: && \delta \tilde\l = {\delta \l \over \sqrt{1+2 h x}} - {\delta h x
1344: \l \over (1+2 hx)^{3/2}} = \xi {d \over dt} \tilde\l -
1345: {1 \over 2} x \dot\xi \left( {1-2h \over 1+2 hx} \right) \tilde\l.
1346: \eqae
1347: For $x =-1$ we find again a polynomial result $\delta \tilde\l = \xi
1348: \dot{\tilde\l} + {1 \over 2} \dot\xi \tilde\l$. The second term indicates
1349: that $\tilde\l$ is a half-density, in agreement with our earlier
1350: observation that $1-2h=e_m{}^\mu=\det e_m{}^\mu$ is a density, and $\tilde\l = \sqrt{1-2 h} \l$ is a half-density.
1351:
1352: We already saw that it is natural to rescale $\psi$ as in
1353: (\ref{1forty}). After the rescaling $\psi \sqrt{1+2
1354: hx} = \tilde\psi$ according to which $-i \psi \dot\varphi \l = -i
1355: \tilde\psi \dot\varphi \tilde\l$, the field $\tilde\psi$ transforms
1356: for $x =-1$ as
1357: \eqa
1358: \delta \tilde\psi = \xi {d \over dt} \tilde\psi - {1 \over 2} \dot\xi
1359: \tilde\psi.
1360: \eqae
1361:
1362: Collecting all results for $x=-1$, and dropping the tildas, we have
1363: found the following action and
1364: transformation rules for general coordinate transformations
1365: \eqa
1366: && L = {1 \over 2} \dot\varphi^2 + {i \over 2} \l \dot\l -h
1367: (\dot\varphi^2 + i \l \dot\l)- i \psi \dot\varphi \l,
1368: \nn
1369: && \delta \varphi = \xi \dot\varphi , \quad \delta \l = \xi \dot\l + {1
1370: \over 2} \dot\xi \l,\quad \delta h = {1
1371: \over 2} \dot\xi + \xi \dot{h} - \dot\xi h , \quad \delta \psi = \xi
1372: \dot\psi - {1 \over
1373: 2} \dot\xi \psi.
1374: \label{general}
1375: \eqae
1376: The susy transformation rules for this model were given in (\ref{delta}) and (\ref{polynomial}),
1377: \eqa
1378: && \delta \varphi = i \epsilon (1-2 h) \l,\quad \delta \l = -(1-2 h)
1379: \dot\varphi \epsilon,\nn
1380: && \delta \psi = (1-2h) [ (1-2 h) \dot\epsilon - \dot{h} \epsilon ], \quad
1381: \delta h = -(1-2 h) i \epsilon \psi.
1382: \label{1sixty}
1383: \eqae
1384:
1385: Hence we have found two polynomial formulations of this supergravity model,
1386: one without a coupling $h \l \dot\l$ in (\ref{1thirtythree}), and one with it in (\ref{general}).
1387: If one varies
1388: the susy Noether current $\dot\varphi \l$ in flat
1389: space under rigid susy, one finds
1390: \eqa
1391: \delta ( \epsilon ) \dot\varphi \l = {d \over dt} (i \epsilon \l) \l
1392: - \dot\varphi \dot\varphi \epsilon =
1393: -(\dot\varphi \dot\varphi + i \l \dot\l ) \epsilon.
1394: \label{1sixtyone}
1395: \eqae
1396: This is the current which couples to $h$ in the model with a $h \l
1397: \dot\l$ coupling. The action of this
1398: model can suggestively be written as
1399: \eqa
1400: L = {1 \over 2} (1-2 h) (\dot\varphi^2 +i \l \dot\l ) - i \psi \dot\varphi
1401: \l.
1402: \label{1sixtyfoour}
1403: \eqae
1404: It is this formulation of the model which is easiest to write in superspace.
1405:
1406: \mathversion{bold}
1407: \section{Rigid and local $N=1$ superspace}
1408: \mathversion{normal}
1409:
1410: We now turn to the superspace description.
1411: The coordinates of the superspace for our toy model are $t$ and $\theta$; both are real, and $\theta$ is a Grassmann variable.
1412: A superfield $\Phi$ depends on $t$ and $\theta$, but a Taylor expansion of $\Phi$ in terms of $\theta$ contains only two terms because $\theta \theta =0$. We begin with
1413: \eqa
1414: \Phi = \varphi + i \theta \l,
1415: \eqae
1416: where $\varphi$ and $\l$ are arbitrary functions of $t$.
1417: Since $\varphi$ is real, also $\Phi$ is real, and this requires the factor $i$.
1418: Susy transformations in flat space are generated by the hermitian%
1419: \footnote{ To show that $Q$ is hermitian, note that from the anticommutator $\{ {\del \over \delta \theta} , \theta \} =1$
1420: if follows that $\del / \del \theta$ is hermitian.
1421: Likewise, it follows from the commutator $[ {\del \over \del t} , t ] =1$ that $\del / \del t$ is antihermitian.}
1422: generator $Q$
1423: \eqa
1424: Q= {\del \over \del \theta} + i \theta {\del \over \del t},
1425: \eqae
1426: because
1427: \eqa
1428: \epsilon Q \Phi &=& \epsilon \left( {\del \over \del \theta} + i
1429: \theta {\del \over \del t} \right) (
1430: \varphi + i \theta \l) = i \epsilon \l + i \theta (-\epsilon \dot\varphi)\nn
1431: &=& \delta \varphi + i \theta \delta \l.
1432: \eqae
1433: The results for $\delta \varphi$ and $\delta \l$ agree with (\ref{1five}).
1434:
1435: The susy covariant derivative, by definition, anticommutes
1436: with $Q, \{ Q,D \} =0$, and this determines $D= {\del \over \del \theta} - i \theta
1437: {\del \over \del t}$. The action can be written as
1438: \eqa
1439: S &=& {i \over 2} \int dtd \theta (\del_t \Phi ) (D \Phi) \nn
1440: &=& {i \over 2} \int dtd \theta (\dot\varphi + i \theta \dot\l ) ( i
1441: \l - i \theta \dot\varphi) = {1
1442: \over 2} \int [ \dot\varphi \dot\varphi + i \l \dot\l ] dt.
1443: \eqae
1444: We used $\int d \theta \theta =1$ and $\int d \theta c =0$ if $c$ is
1445: independent of $\theta$. This definition of integration is called the
1446: Berezin integral \cite{Berezinfa} and it follows from translational invariance in $\theta$.
1447: Namely requiring that $\int d \theta f(\theta) = \int d \theta
1448: f(\theta+c)$, and using that $f(\theta)=a + b \theta$ because $\theta
1449: \theta=0 $, one finds $\int d \theta c =0$. We normalize $\theta$ such
1450: that $\int d \theta \theta = 1$.
1451:
1452: The susy invariance of the action follows from $\delta \Phi =[\epsilon
1453: Q, \Phi]$, $\del_t \epsilon Q = \epsilon Q \del_t$ and $D \epsilon Q =
1454: \epsilon Q D$. One gets then for the variation of the action an expression
1455: of the form $\delta S = {i \over 2} \int d t d \theta Q[...]$. The
1456: ${\del \over \del\theta}$ in $Q$ yields zero because $\del_\theta [\dots]$
1457: contains no $\theta$ and $\int d \theta c = 0$. The $i \theta {\del
1458: \over \del t}$ in $Q$ yields zero because of the total t-derivative. This shows that any action built from superfields and derivatives ${\del \over \del t}$ and $D$ is always supersymmetric.
1459:
1460: After performing the $\theta$ integration, we have obtained the correct $t$-space action. We can also show before the $\theta$ integration that $\partial_t \phi D \phi $ is the correct Lagrangian density by checking that it has the correct dimension: the dimension of $\phi$ is that of $\varphi$, so $[ \phi ] =- 1/2$, and $[ dt ] =-1$ and $d \theta = 1/2$. The action should be dimensionless, so we need derivatives acting on $\phi$ with dimension $+3/2$. The only derivatives we have available are ${\del \over \del t}$ and $D$, so the Lagrangian density is unique.
1461: If one considers generalized unitary group elements $e^{\theta Q + i t H}$, one finds by
1462: left-multiplication the vielbein fields
1463: \eqa
1464: && e^{\epsilon Q +iaH} e^{\theta Q +itH}= e^{\epsilon Q +iaH +\theta Q +itH
1465: +{1 \over 2} [ \epsilon Q, \theta Q ]} \nn
1466: && \equiv \exp [ \{ z^\Lambda + dz^M E_M{}^\Lambda (Z) \} ] T_\Lambda.
1467: \label{equiv}
1468: \eqae
1469: We used the Baker-Campbell-Hansdorff formua, and $z^\Lambda = (t, \theta ), T_\Lambda = (H,Q)$ and $dz^M = (a, \e)$.
1470: Using $[\epsilon Q, \theta Q]=- \epsilon \{Q,Q\} \theta= - 2\epsilon H
1471: \theta$ we find in the exponent $(\theta + \epsilon)Q + i (t + a +
1472: i\epsilon \theta)H$. Thus $\delta (\epsilon) \theta = \epsilon $ and $\delta (\epsilon) t = i
1473: \epsilon \theta$. This is a nonlinear representation of the susy algebra
1474: \eqa
1475: \{Q,Q\}=2H {\bf ,} \;\; [Q,H]=0, \;\; {\rm and } \;\; [H,H]=0.
1476: \eqae
1477: in terms of coordinates. The field representation in (\ref{1five}) is induced by
1478: this coordinate representation, namely $\Phi ' (t',\theta ') = \Phi
1479: (t,\theta)$.%
1480: \footnote{Actually, one finds in this way minus the result of (2.6). To get (2.6), we would have replaced $\epsilon$ (and $a$) in (4.5) by $-\epsilon$ (and $-a$). In general, coordinates transform contragradiently (opposite) to fields.}
1481:
1482: One can repeat the same procedure as in (\ref{equiv}) but now using right multiplication.
1483: Right multiplication yields $D$, and now we understand why $\epsilon_1 D$ and $\epsilon_2 Q$ commute: because left and
1484: right multiplications commute.
1485:
1486: In 4 dimensions it is better to define $x$-space components by $D$
1487: derivatives of $\Phi$ at $\theta
1488: =0$. Here this makes no difference since there are only two components in $\Phi$
1489: \eqa
1490: \Phi_{\big| \theta =0} = \varphi (t), \quad D \Phi_{\big| \theta =0} = i \l (t).
1491: \eqae
1492: Since $\int dt \int d \theta$ can be replaced
1493: by $\int dt D$ (recall that $\int d \theta = {\del \over \del \theta}$
1494: and $\int dt i \theta {\del \over
1495: \del t} L =0$ because it is a total $t$-derivative) we get
1496: \eqa
1497: S &=& {i \over 2} \int dt [ (\del_t D \Phi) D \Phi + \del_t \Phi DD \Phi
1498: ]\nn
1499: &=& {i \over 2} \int dt [ (i \dot\l ) (i \l) - \dot\varphi i \dot\varphi ].
1500: \eqae
1501:
1502: The extended susy algebra of $Q, D$ and $H = i {\del \over \del t}$ reads
1503: \eqa
1504: && \{ Q, Q \} = 2 H,\; \; \{ D,D \} =- 2 H, \; \; \{ Q, D \} =0, \nn
1505: && [H, Q] = [ H,D ] =0.
1506: \eqae
1507:
1508: To formulate also the supergravity model in superspace, the $x$-space
1509: action of (3.43)
1510: \eqa
1511: L = {1 \over 2} (1-2 h) (\dot\varphi^2 +i \l \dot\l) -i \psi \dot\varphi\l,
1512: \eqae
1513: suggests to introduce as superfield for the gauge fields of supergravity
1514: \eqa
1515: H = ( 1-2 h) +2i \theta \psi,
1516: \label{2seventytwo}
1517: \eqae
1518: and to write
1519: \eqa
1520: L = { i \over 2} \int d \theta H (\del_t \Phi) D \Phi.
1521: \label{1seventyfour}
1522: \eqae
1523: This is the action in the so-called prepotential approach.
1524: This action is not manifestly covariant w.r.t.\ general coordinate
1525: transformations in superspace, but we
1526: shall soon give another formulation which is manifestly covariant and
1527: equivalent, giving the same
1528: $t$-space action. Let us first check that the expression in (\ref{1seventyfour}) yields
1529: indeed the $t$-space action
1530: \eqa
1531: L &=& {i \over 2} \int d \theta [ (1-2h) +2i \theta \psi ] (
1532: \dot\varphi + i \theta \dot\l )(i \l - i \theta
1533: \dot\varphi) \nn
1534: &=& {1 \over 2} (1-2h) (\dot\varphi \dot\varphi + i \l \dot\l ) - i
1535: \psi \dot\varphi \l.
1536: \eqae
1537: This agrees with (\ref{1sixtyfoour}).
1538:
1539: We now turn to the covariant approach. One introduces ``flat
1540: covariant derivatives"
1541: \eqa
1542: \cd_{\bar{t}} &=& E_{\bar{t}}{}^\theta D_\theta + E_{\bar{t}}{}^t \del_t ,\nn
1543: \cd_{\bar\theta} &=& E_{\bar\theta}{}^\theta D_\theta +
1544: E_{\bar\theta}{}^t \del_t.
1545: %\cd_\theta &\equiv& \cd.
1546: \eqae
1547: The superfields $E_{\bar{t}}{}^\theta$ etc are the inverse super vielbeins.
1548: The bars on $\bar\theta$ and $\bar{t}$ denote that these are flat
1549: indices, while $\theta$ and $t$ are curved
1550: indices. Because in rigid susy the natural derivatives are $\del_t$
1551: and $D_\theta$ rather than $\del_t$ and
1552: $\del_\theta$, one introduces vielbeins and parameters on this basis.
1553: We introduce the notation $D_\L = \{ D_\theta ,
1554: \del_t \}$ and $\cd_M = \{ \cd_{\bar\theta}, \cd_{\bar{t}} \}$ where
1555: $\L = \{ \theta, t \}$ and $M = \{
1556: \bar\theta, \bar{t} \}$. Then the super-vielbeins on the basis
1557: $\del_\Lambda =(\del_\theta , \del_t)$ are related to the vielbeins on the
1558: basis $\cd_\Lambda$ as follows
1559: \eqa
1560: && \tilde{E}_M{}^\L \del_\L = E_M{}^\L D_\L,\;\; \tilde{E}_M{}^\theta =
1561: E_M{}^\theta, \;\; {\rm but} \;\;
1562: \tilde{E}_M{}^t = E_M{}^t - E_M{}^\theta i \theta, \nn
1563: && \tilde\Xi^\L \del_\L = \Xi^\L D_\L,\;\; \tilde\Xi^\theta = \Xi^\theta,
1564: \;\; {\rm but}\; \; \tilde\Xi^t =
1565: \Xi^t - \Xi^\theta i \theta.
1566: \label{1seventyseven}
1567: \eqae
1568:
1569: A supercoordinate transformation of $\Phi$ becomes $\delta \Phi =
1570: \tilde\Xi^\L \del_\L \Phi = \Xi^\L D_\L
1571: \Phi$. The (inverse) super vielbein $\tilde{E}_M{}^\L$ transforms as
1572: in ordinary general relativity.
1573: \eqa
1574: \delta \tilde{E}_M{}^\L &=& \tilde\Xi^\Pi (\del_\Pi \tilde{E}_M{}^\L) -
1575: \tilde{E}_M{}^\Pi (\del_\Pi
1576: \tilde\Xi^\L) \nn
1577: &=& \Xi^\Pi (D_\Pi \tilde{E}_M {}^\L) - E_M{}^\Pi (D_\Pi \tilde\Xi^\L).
1578: \eqae
1579: Contracting with $\del_\L$ on both sides of the equation and then
1580: using (\ref{1seventyseven}) one obtains
1581: \eqa
1582: \delta E_M{}^\S = \Xi^\Pi D_\Pi E_M{}^\S + E_M{}^\Pi D_\Pi \Xi^\S +
1583: E_M{}^\Pi \Xi^\L T^{(0)}_{\L \Pi}{}^\S,
1584: \eqae
1585: where $[D_\L , D_\Pi \} = T^{(0)}_{\L \Pi}{}^\S D_\S$. In our
1586: case only the $\theta\theta$ component of the supertorsion is
1587: nonvanishing, $T^{(0)\; t}_{\theta\theta} = -2i$ (see (4.9)), and we get
1588: \eqa
1589: \delta E_M{}^t &=& \Xi^\Pi D_\Pi E_M{}^t - E_M{}^\Pi (D_\Pi \Xi^t +
1590: 2i \Xi^\theta \delta_\Pi{}^\theta), \nn
1591: \delta E_M{}^\theta &=& \Xi^\Pi D_\Pi E_M{}^\theta - E_M{}^\Pi D_\Pi \Xi^\theta.
1592: \eqae
1593:
1594: In $x$-space we only have the field content of $H$. To reduce the extra
1595: superfields we shall now impose a constraint on the super-vielbein and
1596: choose a gauge. Afterward we shall come back to the general super
1597: coordinate transformations in the presence of this constraint and gauge.
1598:
1599: To avoid cumbersome notation we write $\cd$ for $\cd_{\bar\theta}$ and
1600: parametrize $\cd$ as follows
1601: \eqa
1602: \cd = E D + X i \del_t, \quad {\rm with} \quad D=D_{\theta} = {\del \over \del \theta}
1603: - i \theta {\del \over \del t}.
1604: \eqae
1605: Thus $E_{\bar\theta}{}^\theta$ is denoted by $E$, and
1606: $E_{\bar\theta}{}^t$ by $i X$. The fields in
1607: $\cd_{\bar{t}}$ are determined in terms of the fields in $\cd$ by imposing
1608: the following \emph{constraint}.
1609: \eqa
1610: {1 \over 2} \{ \cd , \cd \} = -i \cd_{\bar{t}}.
1611: \label{1eightyone}
1612: \eqae
1613: This is the curved space extension of $\{ D, D \} =- 2 i \del_t$. It
1614: is a so-called conventional constraint; in 4 dimensions
1615: it expresses the bosonic Lorentz connection in terms of the fermionic
1616: Lorentz connections, but here no connection is present,
1617: and the constraint expresses the super vielbeins with flat index
1618: $\bar{t}$ in terms of those with flat index $\bar\theta$.
1619:
1620: The anticommutator in (\ref{1eightyone}) yields
1621: \begin{gather}
1622: {1 \over 2} \{ ED + X i \del_t, ED + Xi \del_t \} \nonumber \\
1623: = [-E^2 + E (D X)
1624: + i X \dot{X} ] i \del_t + [E (DE) + i X \dot{E} ]D \nonumber \\
1625: \equiv -i [E_{\bar{t}}{}^{\theta} D + E_{\bar{t}}{}^{t} \del_{t}].
1626: \end{gather}
1627: Hence
1628: \eqa
1629: E_{\bar{t}}{}^t &=& E^2 - E (DX) - i X \dot{X}, \nn
1630: E_{\bar{t}}{}^\theta &=& i E (DE) - X \dot{E}.
1631: \eqae
1632: As a matrix, the inverse super vielbein is thus given by
1633: \eqa
1634: E_M{}^\L = \left( \begin{array}{ccc} E^2 - E (DX) - i X \dot{X}
1635: &\qquad & i E (DE) - X \dot{E} \\ i X & & E
1636: \end{array} \right).
1637: \eqae
1638: The superdeterminant of $\left( \begin{array}{cc} A & B \\ C & D
1639: \end{array} \right)$ being given by%
1640: \footnote{ Use $\left(
1641: \begin{array}{cc} A & B \\ C & D \end{array} \right) = \left(
1642: \begin{array}{cc} A-BD^{-1} C
1643: & BD^{-1} \\ 0 & 1 \end{array} \right) \left( \begin{array}{cc} 1 &
1644: 0 \\ C & D \end{array} \right)$ and take the product of
1645: the \\two superdeterminants.}
1646: \eqa
1647: s \det \left( \begin{array}{cc} A & B \\ C & D \end{array} \right) =
1648: {1 \over D} (A - B D^{-1} C),
1649: \eqae
1650: and we find
1651: \eqa
1652: s \det E_M{}^\L &=& {1 \over E} \left\{ E^2 - E (DX) - i X \dot{X} +
1653: [E (DE) + i X \dot{E} ] {1 \over E} X
1654: \right\}
1655: \nn &=& E - (DX) - i {X \dot{X} \over E} + {DE \over E} X,
1656: \eqae
1657: where we used that $XX=0$.
1658:
1659: We now choose \emph{the gauge which sets $i X = E_{\bar\theta}{}^t$ to zero}. This will fix some
1660: of the gauge symmetries in superspace as we discuss
1661: later. In the covariant formalism, an invariant action (an action
1662: invariant under super general coordinate
1663: transformation, there are no Lorentz transformations) is given by ($s
1664: \det E_\L{}^M$) times covariant
1665: derivatives on tensors (scalars in our case) with flat indices. We
1666: find then for the covariant action in
1667: superspace, which generalizes (4.12),
1668: \eqa
1669: S &=& {i \over 2} \int dt d \theta ( s \det E_\L{}^M )
1670: (\cd_{\bar{t}} \Phi ) (\cd_{\bar{\theta}} \Phi ) \nn
1671: &=& \int dt d \theta \left( {i \over 2E} \right) (E^2 \del_t \Phi )
1672: (ED \Phi ).
1673: \eqae
1674: (A term $i E(DE) (D\Phi)$ in $\cd_{\bar{t}} \Phi$ cancels out).
1675: Identifying $E^2 =H$ we find the same action as in (\ref{1seventyfour}).
1676:
1677: Let us now discuss the local symmetries in superspace. On $\Phi$ a
1678: general supercoordinate transformation would read
1679: \eqa
1680: \delta \Phi = \tilde\Xi^t \del_t \Phi + \tilde\Xi^\theta \del_\theta \Phi.
1681: \eqae
1682: To incorporate rigid susy as the flat superspace limit $E=1, X=0$, we
1683: expand instead in terms of $\del_t$ and $D_\theta$.
1684: Then
1685: \eqa
1686: \delta \Phi = \Xi^t \del_t \Phi + \Xi^\theta D_\theta \Phi = \Xi \Phi\;
1687: \; {\rm with} \; \;\Xi = \Xi^t \del_t + \Xi^\theta D_\theta.
1688: \eqae
1689: We used again (4.15). But $\Xi^\theta$ is not a free super parameter as we now show. We can also write $\Xi \Phi$ as the operator equation $[ \Xi, \Phi ] = 0$.
1690:
1691: The transformation rules of the super vielbein follow from
1692: \eqa
1693: \delta \cd_M = [ \Xi , \cd_M ].
1694: \eqae
1695: In particular for $\cd_{\bar{\theta}} = \cd = ED + X i \del_t$ we find
1696: \eqa
1697: \delta [ED +X i \del_t ] = [ \Xi^t \del_t + \Xi^\theta D, E D + X i \del_t].
1698: \eqae
1699: If the gauge $X=0$ has been reached, the terms with $\del_t$ on the
1700: right-hand side should vanish because they are absent on the left-hand
1701: side. This yields
1702: the following relation between $\Xi^\theta$ and $\Xi^t$
1703: \eqa
1704: -\Xi^\theta E 2i \del_t - E (D \Xi^t ) \del_t =0 \,\Rightarrow\,
1705: \Xi^\theta = {i \over 2} (D \Xi^t).
1706: \eqae
1707: Expanding
1708: \eqa
1709: \Xi^t = \xi (t) -2 i \theta \epsilon (t),
1710: \eqae
1711: we find
1712: \eqa
1713: \Xi^\theta = {i \over 2} (-2i \epsilon (t) - i \theta \dot\xi ) =
1714: \epsilon (t) + {1 \over 2} \theta \dot\xi.
1715: \eqae
1716:
1717: The transformation of $\Phi$ becomes then
1718: \eqa
1719: \delta \Phi = (\xi - 2 i \theta \epsilon ) \dot\Phi + \left( \epsilon
1720: + {1 \over 2} \theta \dot\xi \right)
1721: D \Phi,
1722: \eqae
1723: which reads in components
1724: \eqa
1725: \delta \varphi &=& \xi \dot\varphi + i \epsilon \l, \nn
1726: \delta \l &=& \xi \dot\l + {1 \over 2} \dot\xi \l - 2 \epsilon
1727: \dot\varphi + \epsilon \dot\varphi.
1728: \eqae
1729: These are the correct $x$-space results of (3.38) and (\ref{1sixty})
1730: after defining $\epsilon = \epsilon (x- {\rm space}) (1 -2 h)$.
1731:
1732: To obtain the transformation rules of the supergravity gauge fields
1733: in $E$, we return to
1734: \eqa
1735: \delta \cd_M = [ \Xi, \cd_M ] \; {\rm with} \; \Xi = \Xi^t \del_t +
1736: \Xi^\theta D.
1737: \eqae
1738: From this relation $\delta E$ follows if one takes $\cd_{\bar{\theta}} =
1739: \cd = ED$
1740: and collects all terms with $D$,
1741: \eqa
1742: ( \delta E) D &=& ( \Xi E) D -E (D \Xi^\theta )D, \nn
1743: \delta E &=& \Xi^t \dot{E} + \Xi^\theta (DE) - E(D \Xi^\theta),\nn
1744: \delta H &=& 2 E \delta E = \Xi^t \dot{H} + \Xi^\theta DH - 2H (D \Xi^\theta).
1745: \label{2ninety7}
1746: \eqae
1747: Substituting the component expressions of (\ref{2seventytwo}) leads to
1748: \eqa
1749: && - 2 \delta h + 2i \theta \delta \psi = ( \xi -2 i \theta \epsilon)
1750: (-2 \dot{h} +2i \theta \dot\psi ) + \left(
1751: \epsilon + {1
1752: \over 2} \theta \dot\xi \right) (2 i \psi +2i \theta \dot{h} )\nn
1753: && \quad - 2 [ (1-2 h) +2i \theta \psi ] \left( {1 \over 2} \dot\xi -
1754: i \theta \dot\epsilon \right).
1755: \eqae
1756: This leads to
1757: \eqa
1758: \delta (-2h) &=& -2 \xi \dot{h} +2i \epsilon \psi - (1-2 h) \dot\xi, \nn
1759: \delta (2 \psi) &=& 4 \epsilon \dot{h} +2 \xi \dot\psi + \dot\xi
1760: \psi - 2 \dot{h} \epsilon \nn
1761: && - 2 \dot\xi \psi +2 (1-2 h) \dot\epsilon.
1762: \label{doot}
1763: \eqae
1764: These results agree with (3.41) using again the identification $\epsilon =
1765: \epsilon (x- \; {\rm space} \; ) (1-2 h)$.
1766:
1767: Instead of choosing a gauge we can also find a redefinition of the
1768: fields which leads to the same result. We could demonstrate
1769: this in the most general case, but to simplify the analysis we still
1770: impose the constraint ${1 \over 2} \{ \cd , \cd \} = -i
1771: \cd_{\bar{t}}$. Hence $E_{\bar{t}}{}^t$ and $E_{\bar{t}}{}^\theta$
1772: are given in terms of $E$ and $X$. The action in the gauge
1773: $X=0$ is given by
1774: \eqa
1775: L= \int d \theta \left( {i \over 2} \right) E^2 (\del_t \Phi ) (D \Phi ).
1776: \label{2one01}
1777: \eqae
1778: whereas the action without this gauge choice reads
1779: \begin{gather} \label{2one02}
1780: L = {i \over 2} \int d \theta \left( E-DX - i X \dot{X} /E + {DE
1781: \over E} X \right)^{-1} \nonumber \\
1782: [ (E^2 - E(DX) -i X \dot{X} ) \dot\Phi + ( i EDE - X
1783: \dot{E} ) D \Phi ) ] [ ED \Phi + i X \dot\Phi ].
1784: \end{gather}
1785: The question we ask is this: which field redefinition (redefinition
1786: of $E, X, \Phi$) leads from (4.41) to (\ref{2one01})? The
1787: solution of this question is most easily found by first making a
1788: general symmetry transformation on the fields in (4.41) and
1789: then requiring that the new field $X'$ vanishes. This will express
1790: the symmetry parameter in terms of $E$ and $X$, and provide an
1791: explicit expression for the new field $E'$ in terms of $E$ such that
1792: the action takes the form in (\ref{2one01}).
1793:
1794: The symmetries of the theory are in this case general supercoordinate
1795: transformations. In classical general relativity a contravariant supervector $F^\L$
1796: transforms, by definition, as follows:
1797: $(F^\L )' (Z') = F^\Pi (Z) {\del \over \del Z^\Pi} (Z')^\L$. Since
1798: field redefinitions do not change $Z$ into $Z'$, we
1799: need a way to write general supercoordinate transformations as
1800: relations between $(F^\L)' (Z)$ and $F^\L (Z)$ \emph{at the
1801: same point $Z$}. Such a formalism exists and we now explain it and
1802: then will use it. One may call it the ``operator approach to diffeomorphisms''.
1803:
1804: For a scalar superfield (a scalar with respect to general
1805: supercoordinate transformations) we define, as it is customary for internal symmetries in
1806: Yang-Mills theory,
1807: \eqa
1808: \phi' (Z) = e^{\hat{X} (Z) D} \phi (Z).
1809: \label{2one03}
1810: \eqae
1811: Here $Z= \{ \theta, t\}$ and $\hat{X} (Z)$ is an arbitrary superfield
1812: parameter which we will later write as a complicated
1813: expression in terms of $E$ and $X$. The most general transformation would
1814: be
1815: \eqa
1816: \phi' (Z) = e^{\hat{X} D + \hat{Y} \del_t} \phi (Z).
1817: \eqae
1818: but we will reach our aim with $\hat{Y} =0$. There will then still
1819: remain super-diffeomorphisms with one superfield parameter
1820: which keep $X' =0$. They correspond to $\epsilon$ and $\xi$
1821: transformations in $x$-space.
1822:
1823: Usually one writes a general supercoordinate transformation as
1824: \eqa
1825: \phi' (Z') = \phi (Z),\quad Z' = Z' (Z).
1826: \label{2one05}
1827: \eqae
1828: The relation between (\ref{2one05}) and (\ref{2one03}) becomes clear
1829: if one writes the latter as
1830: \eqa
1831: e^{-\hat{X} (Z) D} \phi' (Z) = \phi' (Z').
1832: \label{2one06}
1833: \eqae
1834: A particular case is $e^{- \hat{X} (Z)D} Z= Z'$. This yields the relation between
1835: $Z'(Z)$ and $\hat{X}(Z)$. To work this out we expand
1836: \eqa
1837: && e^{\hat{X} (Z) D} = 1+ \hat{X} D + {1 \over 2} \hat{X} D \hat{X} D
1838: + {1 \over 3!} \hat{X} D \hat{X} D \hat{X} D \nn
1839: && = 1 + \hat{X} D + {1 \over 2} \hat{X} (D \hat{X} ) D + {1 \over
1840: 3!} \hat{X} (D \hat{X}) (D \hat{X} ) D + \cdots \nn
1841: && = 1 + \hat{X} \left( {e^{(D \hat{X})} -1 \over (D \hat{X})}
1842: \right) D \equiv 1 + f \hat{X} D,
1843: \eqae
1844: where $f = f ((D \hat{X})) = (e^{(D \hat{X})} -1) / (D \hat{X})$. We
1845: used that $\hat{X} \hat{X} =0$, because $\hat{X}$ is anticommuting. It follows that
1846: \eqa
1847: Z' = e^{- \hat{X} D} Z= (1- \bar{f} \hat{X} D) Z=Z- \bar{f} \hat{X} (DZ),
1848: \eqae
1849: where $\bar{f} \equiv f (-(D \hat{X})) = 1- {1 \over 2} (D \hat{X} )
1850: + \cdots$. More explicitly,
1851: \eqa
1852: \theta' = \theta - \bar{f} \hat{X} (D \theta) = \theta - \bar{f}
1853: \hat{X},\quad t' =t - \bar{f} \hat{X} (D t) = t + \bar{f} \hat{X} (i
1854: \theta).
1855: \eqae
1856: Then we find indeed (\ref{2one06})
1857: \eqa
1858: e^{- \hat{X} D} \phi' (Z) &=& (1- \bar{f} \hat{X}D ) \phi ' = \phi '
1859: - \bar{f} \hat{X} \left( {\del \over \del \theta} - i
1860: \theta {\del \over \del t} \right) \phi ' \nn
1861: & = &\phi' ( \theta - \bar{f} \hat{X} , t +\bar{f} \hat{X} i \theta )
1862: = \phi' (Z').
1863: \eqae
1864:
1865: Consider now a transformation with $\hat{X}$ which is such that the
1866: new field $X'$ vanishes. In terms of
1867: \eqa
1868: \cd = ED + iX \del_t,
1869: \eqae
1870: this means that
1871: \eqa
1872: \cd' = e^{\hat{X} D} \cd e^{- \hat{X} D} = E' (Z) D.
1873: \eqae
1874: Thus in $D'$ all terms with free derivatives $i \del_t$ and $i
1875: \del_t D$ must cancel, and the coefficient of $D$ is by
1876: definition $E' (Z)$. Requiring that the coefficients of $i \del_t$
1877: and $i \del_t D$ vanish fixes $\hat{X}$ as a function of
1878: $X$ and $E$.
1879:
1880: The details are as follows. We begin with
1881: \eqa
1882: e^{\hat{X} D} \cd e^{- \hat{X} D} &=& (1 + f \hat{X} D) (ED +X i
1883: \del_t)(1- \bar{f} \hat{X} D)\nn
1884: &=& E' (Z) D + X' (Z) i \del_t.
1885: \label{2one12}
1886: \eqae
1887: Somewhat laborious algebra using $DD = - i \del_t$ yields the following
1888: result for this expression,
1889: \eqa
1890: && [ \{ E+ f \hat{X} (DE) \} D+ \{ X - f \hat{X} E+ f \hat{X} (DX) \}
1891: i \del_t \nn
1892: && - f \hat{X} X i \del_t D] (1- \bar{f} \hat{X} D) \nn
1893: && = [ E + f \hat{X} (DE - ED ( \bar{f} \hat{X} ) - f \hat{X} (DE)
1894: \bar{f} (D \hat{X}) \nn
1895: && - X i \del_t (\bar{f} \hat{X} ) - f \hat{X} E \bar{f} (i \del_t
1896: \hat{X} ) - f \hat{X} (DX) \bar{f} (i \del_t \hat{X} ) ] D
1897: \nn
1898: && + [ X - f \hat{X} E + f \hat{X} (DX) - E \bar{f} \hat{X} ] i \del _t \nn
1899: && + [ - f \hat{X} X - X \bar{f} \hat{X} +f\hat{X}X\bar{f}(D\hat{X})] D i \del_t.
1900: \eqae
1901: The terms with $Di\del_t$ should vanish identically, and one can check that this is the
1902: case. The vanishing of the terms with $i \del_t$ shows that $X$ is
1903: proportional to $\hat{X}$, and then the terms with $D i \del_t$
1904: also vanish. Hence
1905: \eqa
1906: && X = f \hat{X} E - f \hat{X} (DX) + E \bar{f} \hat{X}, \nn
1907: && (1 + f \hat{X} D) X =(f + \bar{f}) E \hat{X} ,\nn
1908: && e^{\hat{X} D} X = (f + \bar{f}) E \hat{X}, \nn
1909: && X = e^{- \hat{X} D} [(f + \bar{f} ) \hat{X} E].
1910: \eqae
1911: This relation expresses $X$ in terms of $\hat{X}$ and $E$, but we
1912: need $\hat{X}$ in terms of $X$ and $E$. We can invert by
1913: expanding the right-hand side and then solve for $\hat{X}$ iteratively
1914: \eqa
1915: X &=& - (f + \bar{f} ) \hat{X} E + \hat{X} D [(f + \bar{f} ) \hat{X}
1916: E ] + \cdots \nn
1917: &=& - \left(2 + {1 \over 3} (D \hat{X})^2 + {1 \over 60} (D
1918: \hat{X})^4 + \cdots \right) \hat{X} E \nn
1919: && + \hat{X} D \left[ 2 \hat{X} E + {1 \over 3} (D \hat{X} )^2
1920: \hat{X} E + {1 \over 60} (D \hat{X} )^4 \hat{X}
1921: E + \cdots \right] \nn
1922: &=& -2 \hat{X} E + 2 \hat{X} (D \hat{X}) E \; .
1923: \eqae
1924: Inversion of $\hat{X} - \hat{X} (X \hat{X}) = - {1 \over 2} X/E
1925: \equiv - \tilde{X}$ yields
1926: \eqa
1927: && \hat{X} = - \tilde{X} + \tilde{X} (D \tilde{X} ) - 2 \tilde{X} (D
1928: \tilde{X})^2 \nn
1929: && \quad 5 \tilde{X} (D \tilde{X} )^3 + 14 \tilde{X} (D \tilde{X})^4
1930: + \cdots \quad {\rm with} \quad \tilde{X} \equiv {1 \over 2} {X \over E}.
1931: \eqae
1932:
1933: This expression for $\hat{X}$ can then be used to find the new field $E'$
1934: \eqa
1935: E' &=& E + f \hat{X} (DE) - E D (\bar{f} \hat{X} ) + \cdots \nn
1936: &=& E- {X \over E} (DE) + E D ({X \over E} ).
1937: \label{2one17}
1938: \eqae
1939: If one then substitutes this $E'$ and the $\Phi'$ in (4.42) into (4.40), the result is
1940: equal to (4.41).
1941:
1942: It may be helpful for becoming more familiar with the operator
1943: approach to diffeomorphisms to check that one gets the same
1944: results as from the usual approach. In the usual approach one begins
1945: with supervielbeins (supervectors in supercoordinate space)
1946: $\tilde{E}_M{}^\L$ and general coordinate transformations $Z^\L
1947: \rightarrow Z^\L + \Xi^\L$ where we take $\Xi^\L$ as infinitesimal.
1948: Then
1949: \eqa
1950: \delta \tilde{E}_M{}^\L &=& \tilde\Xi^\Pi \del_\Pi \tilde{E}_M{}^\L -
1951: \tilde{E}_M{}^\Pi \del_\Pi \tilde\Xi^\L \nn
1952: &=& \Xi^\Pi D_\Pi \tilde{E}_M{}^\L - E_M{}^\Pi D_\Pi \tilde\Xi^\L,
1953: \eqae
1954: with $\del_\Pi = \left( {\del \over \del \theta} , {\del \over \del
1955: t} \right)$ and $D_\Pi = \left( \tfrac{\del}{\del \theta} - i
1956: \theta {\del \over \del t} ,{\del \over \del t} \right)$. From (4.15)
1957: \eqa
1958: && \tilde{E}_M{}^\theta = E_M{}^\theta,\quad \tilde{E}_M{}^t = E_M{}^t -
1959: E_M{}^\theta i \theta\nn
1960: && \tilde\Xi^\theta = \Xi^\theta,\quad \tilde\Xi^t = \Xi^t -\Xi^\theta i \theta.
1961: \eqae
1962: One finds by straightforward substitution
1963: \eqa
1964: \delta E_M{}^\theta &=& \delta \tilde{E}_M{}^\theta = \Xi^\Pi D_\Pi
1965: \tilde{E}_M{}^\theta - E_M{}^\Pi D_\Pi \tilde\Xi^\theta \nn
1966: &=& \Xi^\Pi D_\Pi E_M{}^\theta - E_M{}^\Pi D_\Pi \Xi^\theta.
1967: \eqae
1968: This agrees with (4.36).
1969: However, in $\delta E_M{}^t$ there are extra terms beyond those which
1970: are present in the corresponding relation for $\tilde{E}_M{}^t$.
1971: \eqa
1972: && \delta E_M{}^t = \delta ( \tilde{E}_M{}^t + E_M{}^\theta i \theta
1973: ) = \Xi^\Pi D_\Pi \tilde{E}_M{}^t -E_M{}^\Pi D_\Pi (
1974: \Xi^t - \Xi^\theta i \theta) \nn
1975: && + ( \Xi^\Pi D_\Pi E_M{}^\theta - E_M{}^\Pi D_\Pi \Xi^\theta ) i
1976: \theta. %- E_M{}^\theta i \Xi^\theta
1977: \eqae
1978: %where we used that in the last term that $\delta \theta =
1979: %\Xi^\theta$.
1980: Various terms cancel and one obtains
1981: \eqa
1982: \delta E_M{}^t = \Xi^\Pi D_\Pi E_M{}^t - E_M{}^\Pi D_\Pi \Xi^t - 2 i
1983: E_M{}^\theta \Xi^\theta.
1984: \eqae
1985: The last term is due to the rigid torsion $T_{\theta \theta}{}^t =
1986: -2i$. Expanding the finite result for $E'$ and $X'$ in
1987: (\ref{2one12}) to first order in $\hat{X} = \Xi^\theta$ we find agreement.
1988:
1989: One may also check that
1990: \eqa
1991: \delta \cd &=& ( e^{\hat{X} D} E_{\bar{\theta}}{}^\L e^{-\hat{X} D} ) D_\L +
1992: E_{\bar{\theta}}{}^\L (e^{\hat{X} D} D_\L e^{- \hat{X} D} )
1993: \nn
1994: &=& \delta E_{\bar{\theta}}{}^\L D_\L,
1995: \eqae
1996: but note that $e^{\hat{X} D} E_{\bar{\theta}}{}^\L e^{- \tilde{X} D}$ only
1997: yields the transport terms $E_{\bar{\theta}}{}^\L (Z')$, while the
1998: terms with $e^{\hat{X} D} D_\L e^{-\hat{X} D}$ yield the terms which
1999: are due to the index $\L$ of $E_{\bar{\theta}}{}^\L$.
2000:
2001: \section{Extended rigid supersymmetries}
2002:
2003: We can also construct models with extended $(N>1)$ susy. We restrict ourselves to rigid supersymmetry, but $N>1$ supergravities can also be constructed. We perform the analysis in $x$-space but one could also use $N>1$ superspace.
2004:
2005: Consider the action
2006: \eqa
2007: L = {1 \over 2} \dot\phi \dot\phi + {i \over 2} \sum^N_{j=1} \l^j \dot\l^j.
2008: \eqae
2009: It is invariant under the following rigid susy transformations
2010: \eqa
2011: \delta \phi = i \sum^N_{j=1} \epsilon^j \l^j , \delta \l^j = -\dot\phi
2012: \epsilon^j.
2013: \eqae
2014: The proof is as before: $\delta L = \dot\phi {d \over dt} (i
2015: \epsilon^j \l^j) + i \l^j {d \over dt} (\dot\phi \epsilon^j)$
2016: vanishes after partial integration.
2017:
2018: There is something unusal about this action: it has more fermions ($N$)
2019: than bosons (one). Nevertheless it has the same number
2020: of fermionic and bosonic states because after quantization there are
2021: only zero modes $(\hat{x}_0$ and
2022: $\hat{\l}^j_0)$ and their conjugate momenta ($\hat{p}$ and again
2023: $\hat{\l}^j_0$). The $\hat\l^j_0$ can be combined into $[{N \over 2}]$
2024: annihilation and creation operators $(\hat\l^1_0 \pm i
2025: \hat\l^2_0) / \sqrt{2}$, ... If $N$ is odd, the last $\hat\l^N_0$
2026: yields projection operators $P_{\pm}={1 \over 2}(1 \pm \hat\l^N_0)$ and
2027: the Hilbert space ${\mathcal H}$ split into two spaces ${\mathcal H}=P_{\pm}
2028: {\mathcal H}$. No operator can bring one from a state in ${\mathcal
2029: H}_+$ to a
2030: state in ${\mathcal H}_-$ because operators either have no $\hat\l^N_0$, or if
2031: they have one $\hat\l^N_0$ then still $\hat\l^N_0 {\mathcal H}_+ ={\mathcal H}_+$
2032: because $\hat\l^N_0 {1 \over 2}(1+ \hat\l^N_0)= {1 \over 2} (1+
2033: \hat\l_0^N)$.
2034: The bosonic states are $e^{ip\hat{x}_0} \mid 0 \rangle = \mid
2035: p \rangle$ but also $\hat\l^1_0 \hat\l^2_0 \mid p \rangle$ etc., and
2036: the fermionic states are $\hat\l^j_0 \mid p \rangle$ but also
2037: $\hat\l^1_0 \hat\l^2_0 \hat\l^3_0 \mid p \rangle$ etc. In fact there
2038: are $2^{N-1}$ bosonic states (with an even number of
2039: $\hat\l^j_0$) and $2^{N-1}$ fermionic states (with an odd number of
2040: $\l_0^j$). In higher dimensions there are also oscillators,
2041: and then one can count the number of bosonic and fermionic states by
2042: looking at the number of propagating fields.%
2043: \footnote{ In $d=(1,1)$ one can divide the real scalar into a left-moving piece
2044: $\varphi (x+t)$ and a right-moving piece $\varphi (x-t)$,
2045: and susy requires then a two-component spinor $\l = {\l_+\choose\l_-}$ with $\l_+ (x+y)$ right-moving
2046: on-shell and $\l_- (x-t)$ left-moving. In the zero mode sector there
2047: are always the same number of bosonic and
2048: fermionic states but in the nonzero mode sector one may drop $\l_+$
2049: (or $\l_-$) and still have susy. Then $\l_-$ transforms
2050: into $\varphi$ and vice-versa, while $\l_+$ is inert.} For
2051: example, in $d=(3,1)$ one has the Wess-Zumino model with one Majorana
2052: spinor $\l^\a$ (2 states) and two real propagating massless
2053: scalars (again 2 states per oscillator).
2054:
2055: In our model similar things happen. To close the algebra off-shell
2056: we need an equal number of bosonic and fermionic field
2057: components. Thus for $N=2$ we need one real auxiliary field, which
2058: we call $F$. The action reads
2059: \eqa
2060: L = {1 \over 2} \dot\varphi \dot\varphi - {i \over 2} (\l^1 \dot\l^1
2061: + \l^2 \dot\l^2) + {1 \over 2} F^2.
2062: \eqae
2063: One can view $(\varphi, \l^1)$ as one multiplet, and $(\l^2, F)$ as
2064: another. The action for these multiplets reads in
2065: superspace
2066: \eqa
2067: L (\varphi, \l^1) &=& {1 \over 2} \dot\varphi \dot\varphi - {i \over
2068: 2} \l^1 \dot\l^1 \sim \int d \theta D \Phi \dot\Phi,\quad \Phi =
2069: \varphi + i \theta \l^1, \nn
2070: L (\l^2, F) &=& - {i \over 2} \l^2 \dot\l^2 + {1 \over 2} F^2 \sim
2071: \int d \theta \Psi D \Psi, \quad \Psi = \l^2 + \theta F.
2072: \eqae
2073: Clearly $\delta \varphi = i \epsilon^1 \l^1$ and $\delta \l^1 =
2074: \dot\varphi \epsilon^1$ form a closed susy algebra, $\{ Q, Q \}
2075: \sim P$. Also $\delta \l^2 = F \epsilon^2$ and $\delta F = i
2076: \epsilon^2 \dot\l^2$ forms the same closed algebra. However, in
2077: $x$-space we can write down more general rules
2078: \eqa
2079: \delta \varphi &=& \sum^2_{j=1} i \epsilon ^j \l^j , \delta \l^j =
2080: \dot\varphi \epsilon^j + \a^{jk} \epsilon^k F, \nn
2081: \delta F &=& i \epsilon^j \beta^{jk} \dot\l^k,
2082: \eqae
2083: where $\a$ and $\beta$ are arbitrary real matrices. Invariance of
2084: the action requires
2085: \eqa
2086: \a^T = \beta.
2087: \eqae
2088: The commutator algebra on $\varphi$ yields
2089: \eqa
2090: && [ \delta (\epsilon_1), \delta (\epsilon_2) ] \varphi = \sum i
2091: \epsilon^j_2 (\dot\varphi \epsilon_1^j + \a^{jk} \epsilon_1^k
2092: F) - (1 \leftrightarrow 2) \nn
2093: && = (2 i \epsilon^j_2 \epsilon^j_1 ) \dot\varphi + [ i \epsilon_2
2094: (\a + \a^T) \epsilon_1 ] F.
2095: \eqae
2096: Clearly, for $\a$ antisymmetric (hence $\a^{jk} \sim \epsilon^{jk}$)
2097: we find only a translation of $\varphi$. For $F$ one finds
2098: \eqa
2099: && [ \delta (\epsilon_1), \delta (\epsilon_2) ] F = i \epsilon^j_2
2100: \beta^{jk} {d \over dt} (\dot\varphi \epsilon^k_1 + \a^{kl}
2101: \epsilon^l_1 F) - (1 \leftrightarrow 2) \nn
2102: && = [ i \epsilon_2 ( \a+ \a^T) \epsilon_1 ] \ddot{\varphi} + [2i
2103: \epsilon_2 \a^T \a \epsilon_1 ] \dot{F}.
2104: \eqae
2105: We now recognize various symmetries.
2106: \medskip
2107:
2108: \noindent (i) $\delta \varphi = z F$ and $\delta F = z
2109: \ddot{\varphi}$ with $z= i \epsilon_2 (\a + \a^T) \epsilon_1$. This
2110: is an
2111: equation of motion symmetry. (These are symmetries of pairs of
2112: fields $f$ and $g$ of the form $\delta f = A {\delta S \over \del
2113: g}$ and $\delta g= - A {\delta S \over \delta f}$. Obviously they
2114: leave the action invariant).\\
2115: \noindent (ii) The usual translation. \\
2116: \noindent (iii) An extra symmetry $\delta F = \s \dot{F}$ where $\s$
2117: is $2 i \epsilon_2 \a^T \a \epsilon_1 - 2 i \epsilon_2
2118: \epsilon_1$. This is clearly a symmetry of $L = {1 \over 2} F^2$.
2119:
2120: For the fermions the susy commutator yields
2121: \eqa
2122: && [ \delta (\epsilon_1), \delta (\epsilon_2) ] \l^j = \delta
2123: (\epsilon_1) [ \dot\varphi \epsilon^j_2 + \a^{jk} \epsilon^k_2
2124: F] - (1 \leftrightarrow 2) \nn
2125: && = - i (\epsilon^l_1 \epsilon^j_2 - \epsilon^l_2 \epsilon^j_1 )
2126: \dot\l^l + [ (\a^{jk} \epsilon^k_2 ) ( i \epsilon^l_1
2127: \beta^{lm} \dot\l^m ) - (1 \leftrightarrow 2) ] \nn
2128: && = [2 i \epsilon_2^l \epsilon_1^l) \dot\l^j + i (\epsilon^l_2
2129: \epsilon^j_1 \dot\l^l - \epsilon^l_2 \epsilon^l_1 \dot\l^j -
2130: (1 \leftrightarrow 2) ) + [ \cdots ].
2131: \eqae
2132: (iv) $\delta \l^j = s^{jl} \dot\l^l$ with symmetric and real
2133: $s^{jl}$. This is again an equation of motion symmetry.
2134:
2135: The extra symmetries do not form a closed algebra. Their commutators
2136: generate new equation of motion symmetries, etc.,
2137: etc. However, the minimal extension of the susy rules with $F$ forms
2138: a closed algebra
2139: \eqa
2140: && \delta \varphi = i (\epsilon^1 \l^1 + \epsilon^2 \l^2 ) ,\quad \delta
2141: \l^1 = \dot\varphi \epsilon^1 - F \epsilon^2 , \quad\delta \l^2
2142: = \dot\varphi \epsilon^2 + F \epsilon^1, \nn
2143: && \delta F = i \epsilon^1 \dot\l^2 - i \epsilon^2 \dot\l^1.
2144: \eqae
2145:
2146: We can try to construct and $N=2$ superspace by introducing
2147: \eqa
2148: && \theta = {\theta^1 + i \theta^2 \over \sqrt{2}},\quad \bar\theta =
2149: {\theta^1 - i \theta^2 \over \sqrt{2}} , \quad\bar{D} = {\del \over \del
2150: \theta} +
2151: i \bar\theta \del_t, \quad D = {\del \over \del \bar\theta} + i \theta \del_t, \nn
2152: && \bar{Q} = {\del \over \del \theta} - i \bar\theta \del_t, \quad Q =
2153: {\del \over \del \bar\theta} -i \theta \del_t , \quad \l = {\l^1 +
2154: i \l^2 \over \sqrt{2}}, \quad \bar\l = {\l^1 - i \l^2 \over \sqrt{2}} ,\nn
2155: && L = {1 \over 2} \dot\varphi \dot\varphi - i \bar\l \dot\l + {1
2156: \over 2} F^2 ,\quad \Phi = \varphi + i \bar\theta \l + i \theta \bar\l + F \bar\theta \theta, \nn
2157: && L = - \int d \theta d \bar\theta \bar{D} \Phi D \Phi = - \int d
2158: \theta d \bar\theta [ i \bar\l - F
2159: \bar\theta + i
2160: \bar\theta \dot\varphi - \bar\theta \theta \dot{\bar\l} ] [ i \l + F \theta + i \theta \dot\varphi - \theta \bar\theta
2161: \dot\l ] \nn
2162: && \quad = -i \bar\l \dot\l + i \dot{\bar\l} \l +
2163: \dot\varphi \dot\varphi + F^2.
2164: \eqae
2165: The $t$-space components of $\Phi$ are given by $\Phi_\mid = \varphi,
2166: D \Phi_\mid = i \l, \bar{D} \Phi_\mid = i \bar\l$ and ${1 \over 2} [
2167: \bar{D}, D] =F$. The susy rules are generated by $(\bar\epsilon D +
2168: \epsilon \bar{D})$ acting on these components, using the $D$ algebra.
2169:
2170: The $N=2$ model has a mass term and a Yukawa coupling
2171: \eqa
2172: && L = \left[ {1 \over 2} \dot\varphi^2 - {i \over 2} \l_1 \dot\l_1 -
2173: {i \over 2} \l_2 \dot\l_2 + {1 \over 2} F^2 \right] \nn
2174: && - m ( F \varphi + i \l_1 \l_2 ) + g \left( {1 \over 2} F \varphi^2
2175: + i \l_1 \l_2 \varphi \right).
2176: \eqae
2177: The corresponding action in $N=1$ superspace is
2178: \eqa
2179: && S^{(0)} = \left( {-i \over 2} \right) \int dt d \theta [ (D_\theta
2180: \phi ) \del_t \phi + i \psi D_\theta \psi ], \nn
2181: && \phi = \varphi + i \theta \l_1,\quad \psi = \l_2 + \theta F,
2182: \eqae
2183: \eqa
2184: S (m) &=& - \int dt d \theta m \phi \psi, \nn
2185: S (g) &=& \int dt d \theta {1 \over 2} g \phi \phi \psi.
2186: \eqae
2187: One can write more generally a superpotential term as
2188: \eqa
2189: S (W) = \int dt d \theta W (\phi) \psi = \int dt (WF + i \l_1 W' \l_2 ),
2190: \eqae
2191: where $W (\phi) = - m \phi + {1 \over 2} g \phi^2 + \cdots$.
2192:
2193: One can also write down nonlinear sigma models for the $N=1$ model
2194: \eqa
2195: && \quad S \, {\rm (nonl)} = {-i \over 2} \int dt d \theta G(\phi)
2196: D_\theta \phi \del_t \phi \nn
2197: &&= \int dt \left[ G (\varphi) \left( {1 \over 2} \dot\varphi^2 - {i
2198: \over 2} \l_1 \dot\l_1 \right) + {i \over 2} \l_1 G' (\varphi) i \l_1
2199: \dot\varphi \right] \nn
2200: && \quad = \int dt G (\varphi) \left( {1 \over 2} \dot\varphi^2 - {i
2201: \over 2} \l_1 \dot\l_1 \right).
2202: \eqae
2203: Similarly, one can find nonlinear $\s$ models with $N=2$ susy in
2204: $N=1$ superspace. For all these $N=2$ models in $N=1$ superspace
2205: there is also a $N=2$ superspace formulation.
2206:
2207: The action in Hamiltonian form for the $N=2$ model reads
2208: \eqa
2209: && L = \dot\varphi p + \dot\l^1 \pi^1 + \dot\l^2 \pi^2 + \dot{F} p_F
2210: - {1 \over 2} F^2 - {1 \over 2} p^2,\nn
2211: && \delta \varphi = {i \over 2} \epsilon^j \left( \l^j + {2 \over i}
2212: \pi^j \right), \quad \delta p =0,\quad \delta \pi^j = {i \over 2} p
2213: \epsilon^j ,\nn
2214: && \delta \l^1 = p^1 \epsilon^1 + F \epsilon^2,\quad \delta \l^2 = p^2
2215: \epsilon^2 - F \epsilon^1, \nn
2216: && \delta p_F = \epsilon^1 \pi^2 - \epsilon^2 \pi^1, \quad \delta F =0.
2217: \eqae
2218: As in the $N=1$ case, there are primary constraints $\pi^j - {i \over
2219: 2} \l^j =0$ and $p_F =0$, and secondary constraints
2220: $F=0$. Since $\delta F \sim \dot\l$ in the Lagrangian formulation,
2221: but $\dot\l =0$ is the full field equation for $\l$, we
2222: cannot express $\dot\l$ in terms of $\del_x \l$ by means of its field
2223: equation, and this explains why $\delta F =0$ and
2224: $\delta p =0$. The susy
2225: Noether charges in the Hamiltonian approach are given by
2226: \eqa
2227: Q^1_{H} &=& \left( i \pi^1 - {1 \over 2} \l^1 \right) p- iF \pi^2, \nn
2228: Q^2_{H} &=& \left( i \pi^2 - {1 \over 2} \l^2 \right) p + i F \pi^1.
2229: \eqae
2230: These changes reproduce the Hamiltonian susy laws \emph{ exactly} if
2231: one uses ordinary Poisson brackets. The reason is that in
2232: the Hamiltonian action there are no constraints. In the action of
2233: the Lagrangian formulation one finds Dirac brackets and the
2234: following hermitian susy charges
2235: \eqa
2236: Q_L^j = - \l^j p.
2237: \eqae
2238: The algebra reads in both cases
2239: \eqa
2240: \{ Q^i_H , Q^j_H \}_P &=& \delta^{ij} (-p^2) =- 2 \delta^{ij} H, \nn
2241: \{ Q^i_L, Q^j_L \}_D &=& \delta^{ij} (-p^2) =- 2 \delta^{ij} H.
2242: \eqae
2243: Thus $F$ and $p_F$ do not transform under $Q_L$, but $p_F$ transforms
2244: under $Q_H$. However, the algebra on all fields is the
2245: same. For example, $\epsilon Q_H \l \sim p \epsilon + F \epsilon
2246: =0$ and $Q_L p = \epsilon$, in agreement
2247: with $H \l =0$ for
2248: $H= H_L = H_H$.
2249:
2250: The $N=2$ model can be used as a toy model for instanton physics. In
2251: Minkowski time $t$ the action is
2252: \eqa
2253: && L = {1 \over 2} \dot\varphi^2 - i \bar\l \dot\l - \bar\l \l (m - g
2254: \varphi) + {1 \over 2} \left( F - m \varphi + {1 \over
2255: 2} g \varphi^2 \right)^2 \nn &&\quad\quad - {1 \over 8} g^2 \left[ \left( \varphi - {m
2256: \over g} \right)^2 - \left( {m \over g} \right)^2
2257: \right]^2.
2258: \eqae
2259: This action is hermitian, and $\bar\l = (\l)^\dagger$. In Euclidean
2260: time $\tau, \l$ and $\bar\l$ become independent complex
2261: spinors. The Wick rotation is a complex Lorentz rotation (a $(U(1)$
2262: rotation) in the $(t, \tau)$ plane (see the joint paper with Waldron)
2263: \eqa
2264: t^1_\theta = e^{i \theta} t, \;\; t^1_{\theta = \pi/2} \equiv \tau = i t.
2265: \eqae
2266: The spinor $\l$ transforms then half as fast
2267: \eqa
2268: \l^1_\theta = e^{i \theta/2} \l, \l_{\theta = \pi/2} \equiv \l_E = \sqrt{i} \l.
2269: \eqae
2270: Making these substitutions, one automatically obtains a
2271: supersymmetric action for the Euclidean case.
2272:
2273: It is a pleasure to thank Martin Ro\v{c}ek for discussions about the
2274: covariant approach to superspace supergravity and the field redefinition in (\ref{2one17}).
2275:
2276: \section{BRST quantization in a Hamiltonian approach}
2277: So far we have been discussing classical actions. The classical
2278: supergravity action was a gauge action, an action with two local
2279: symmetries in out toy model: diffeomorphisms and local supersymmetry. In the
2280: quantum theory, one uses a quantum action which is obtained by
2281: adding two more terms: a gauge fixing term and a ghost action.
2282: The local symmetry is then broken, but a rigid residual symmetry
2283: remains, the so-called BRST symmetry (due to Becchi, Rouet, Stora
2284: and Tyutin). The crucial property of BRST transformations is that
2285: they are nilpotent (see below). An infinitesimal BRST
2286: transformation has as parameter a purely imaginary anticommuting
2287: constant $\L$. It is not the supersymmetry parameter (which is
2288: also anticommuting). Physicists use (at least ) two ways to
2289: formulate the BRST formalism: a Lagrangian approach and a
2290: Hamiltonian approach. The infinitesimal BRST transformation rules
2291: in the former are written as $\d_B\Phi$ (where $\Phi$ denotes any
2292: field), but in the Hamiltonian approach one uses operators and
2293: brackets. For example, the BRST transformations are generated by
2294: a BRST operatorial charge $Q$, and for every field there is a
2295: canonically conjugate field (called momentum by physicists) which
2296: satisfy equal-time canonical commutation rules or
2297: anticommutation rules. Then one has $\d_B\Phi\sim\{\Phi,Q\}$.
2298: Nilpotency means $\d_B\d_B\Phi=0$ in the Lagrangian approach, and
2299: $\{Q,Q\}=0$ in the Hamiltonian approach. In this section we apply
2300: this general formalism to our quantum mechanical toy model. One
2301: word about terminology: we use the words real and hermitian (and
2302: purely imaginary and antihermitian) as equivalent.
2303:
2304: To quantize the supergravity action using covariant quantization
2305: in the Lagrangian approach to BRST symmetry, we should add a gauge
2306: fixing term and a corresponding ghost action. We begin with the
2307: classical action in~(\ref{1thirtythree}) although we could also
2308: start with the classical action in~(\ref{1sixtyfoour}). We fix the
2309: gauge of general coordinate transformations by $h=0$, and the
2310: gauge of local supersymmetry by $\psi =0$. The corresponding
2311: gauge fixing terms in the action are then given by
2312: \begin{equation}\label{pod} L_{\rm fix} = dh+\Delta\psi.
2313: \end{equation}
2314: The fields $d$ and $\Delta$ are Lagrange multipliers which fix
2315: the gauges according to $h=0$ and $\psi=0$. Hermiticity of the
2316: action requires that $d$ be real and $\Delta$ antihermitian. The
2317: ghost action is then given by \eqa L {\rm (ghost)} \L = b \d_B h
2318: + \beta \d_B \psi, \eqae where $\d_B$ are the BRST transformations
2319: and $\L$ is the constant anticommuting imaginary BRST parameter.
2320: The fields $b$ and $\b$ are the antighosts; $b$ and $\b$ are
2321: antihermitian. (Mathematical physicists remove $\L$ on both sides
2322: of this equation and call it then a derivation. It is usually
2323: denoted by $s$, and the relation to $\delta_B$ is
2324: $\delta_B\phi=\Lambda s\phi$ for any field $\phi$). We obtain
2325: thus for the full quantum action \eqa\label{brsttran} && L=L {\rm
2326: (class)} + L {\rm (fix)} + L {\rm ghost)}, \nonumber \\ && L {\rm
2327: (class)} = {1 \over 2} \dot\varphi^2 + {i \over 2} \l \dot\l -h
2328: \dot\varphi^2 -i \psi \dot\varphi \l, \nonumber\\ && L {\rm (fix)}
2329: = dh + \Delta \psi,\quad L {\rm (ghost)} \L = b \d_B h + \beta \d_B
2330: \psi. \eqae
2331:
2332: For the classical fields ($h$ and $\psi$ and $\varphi$ and $\l$)
2333: the BRST transformations are just gauge transformations with a
2334: special choice of the parameters: $\xi = c\L$ and $\epsilon =- i
2335: \g \L$ where $c$ and $\g$ are real ghosts ($\L$ is imaginary, and
2336: $\xi$ and $\epsilon$ are real). Because $\xi$ is commuting, $c$ is anticommuting,
2337: and because $\epsilon$ is anticommuting, $\gamma$ is commuting.
2338: The BRST transformation rules for
2339: the classical fields follow from~(\ref{1thirtythree}),
2340: (\ref{varphi}) and~(\ref{dot}) \eqa && \d_B h = [(1-2 h) \psi \g
2341: + {1 \over 2} (1-2 h) \dot{c} + \dot{h} c ] \L, \nn && \d_B \psi =
2342: [ -i (1-2 h) \dot\g + \dot\psi c ] \L, \nn && \d_B \varphi = (- \l
2343: \g + \dot\varphi c) \L,\quad \d_B \l = [ i (1-2 h) \dot\varphi \g+
2344: \psi\l\g + \dot\l c ] \L .\label{class}\eqae
2345: The ghost action then becomes
2346: \begin{gather}
2347: L(\mathrm{ghost})=b[(1-2h)\psi\gamma +\frac{1}{2}(1-2h)\dot{c} +\dot{h}c]
2348: + \beta[-i(1-2h)\dot{\gamma}+ \dot{\psi}c].
2349: \end{gather}
2350: It is clearly hermitian.
2351:
2352: The BRST transformation rules of the ghosts follow from the
2353: structure constants of the classical
2354: gauge algebra or from the nilpotency of BRST transformations. One has uniformly on all classical
2355: fields ($\varphi , \l, h$ and $\psi$), as discussed
2356: in (3.18)
2357: \eqa && [ \d_s (\e_2) ,
2358: \d_s (\e_1) ] = \d_g (2i (1-2 h) \e_2 \e_1 ) + \d_s (-2 i \psi
2359: \e_2 \e_1 ), \nn && [ \d_g ( \xi_2 ), \d_g (\xi_1) ] = \d_g (-
2360: \xi_2 \dot{\xi_1} + \xi_1 \dot\xi_2 ), \nn && [ \d_g (\xi) , \d_s
2361: (\e) ] = \d_s (- \xi \dot\e ). \eqae Thus the classical local
2362: gauge algebra ``closes": the commutator of two local symmetries
2363: is a linear combination of local symmetries. New from a
2364: mathematical point of view is the appearance of fields ($h$ and
2365: $\psi$) in the composite parameters, and thus in the structure
2366: constants. (One should thus rather speak of ``structure
2367: functions"). One obtains \eqa \d_B c &=& -c \dot{c} \L +i (1-2
2368: h) \g \g \L ,\nn \d_B \g &=& c \dot\g \L + \psi \g \g \L .\eqae An
2369: easier way to obtain these results is to use that the BRST
2370: variation of~(\ref{class}) should vanish. One may check that all
2371: BRST transformation rules preserve the reality properties of the
2372: fields.
2373:
2374: The antighosts $b$ and $\beta$ transform into the auxiliary
2375: fields $d$ and $\Delta$, and the auxiliary fields are BRST
2376: invariant (``contractible pairs") \eqa\label{contr} \d_B b= \L d
2377: , \quad \d_B \beta =\L \Delta , \quad \d_B d=0 ,\quad \d_B \Delta =0. \eqae All
2378: BRST transformation laws are nilpotent, and they leave the action
2379: $S=\int L dt$ invariant.
2380:
2381: In the Lagrangian approach the BRST charge $Q$ is the Noether
2382: charge for rigid BRST transformations but one does not use
2383: brackets. To obtain $Q$, one lets $\L$ become local (t-dependent),
2384: and one collects all terms in the variation of the action
2385: proportional to $\dot\L$. (We used this procedure before to
2386: construct the supersymmetry charge). The BRST transformation
2387: rules of the fields themselves for local $\L$ do not contain by
2388: definition any $\dot\L$; for example, $\d_B \psi = (1-2 h) (-i
2389: \dot\g) \L (t)+\dots$, and not $\d_B \psi = (1-2 h) {d \over dt}
2390: (-i \g \L (t))+\dots$. The classical action yields the following
2391: terms proportional to $\dot\L$ \eqa \d S {\rm (class)} = \int [ L
2392: {\rm (class)} c \dot\L - (1-2 h) \l \dot\varphi \g \dot\L ] dt.
2393: \eqae The term with $c \dot\L$ is expected from the result $\d
2394: \cl = \del_\a ( \xi^\a \cl )$ in general relativity, while the
2395: term with $\g \dot\L$ comes from the variation with ${d \over dt}
2396: \e$ in $i \l {d \over dt} \d \l$ which is canceled by the Noether
2397: term. The gauge fixing term produces no terms with $\dot\L$
2398: because $\d_B h$ and $\d_B \psi$ do not contain $\dot\L$ terms by
2399: definition, but the ghost action
2400: yields further terms proportional to $\dot\L$
2401: \eqa && \d S {\rm (ghost)} = b [ {1 \over 2} (1-2 h) ( -c
2402: \dot{c} +i (1-2 h) \g \g ) \dot\L \nn && \qquad + c ( (1-2h )
2403: \psi \g + {1 \over 2} (1-2 h) \dot{c} + \dot{h} c ) \dot\L ] \nn
2404: && + \beta [ (-i) (1-2h) (c \dot\g + \psi \g \g ) \dot\L + i c
2405: (1-2h) \dot\g \dot\L ] \eqae Several terms cancel in this
2406: expression.
2407:
2408: The BRST charge in the Lagrangian approach is thus \eqa Q &=& c L
2409: {\rm (class)} - (1-2 h) \l \dot\varphi \g \nn &+& b [ {i \over
2410: 2} (1-2 h)^2 \g \g + (1-2 h) c \psi \g ] \nn &-& i\beta (1-2 h)
2411: \psi \g \g. \label{is} \eqae
2412:
2413: In the Hamiltonian approach the BRST charge should contain terms
2414: of the form (we discuss this in more detail later) \eqa Q = c^\a
2415: \phi_\a-\frac12c^\b c^\a f_{\a\b}{}^\g b_\g(-)^\b, \eqae with
2416: ghosts $c^\a = (c, \g)$, antighosts $b_\a = (b, \beta)$, and
2417: first class constraints $\phi_\a = (T, J)$ where $T$ is the
2418: generator of diffeomorphisms and $J$ the generator of
2419: supersymmetry. The sign $(-)^\beta$ is equal to $+1$ if the
2420: corresponding symmetry has a commuting parameter (and thus an
2421: anticommuting ghost); when the ghost is commuting $(-)^\beta$
2422: equals $-1$. However, because the structure constants contain
2423: $\dot\xi$ and $\dot \epsilon$ we expect in $Q$ terms with
2424: derivatives of $c$, namely $bc \dot{c}$ and $\b \dot\g c$ terms.
2425: On the other hand, in a truly Hamiltonian approach no time
2426: derivatives of fields are allowed in the charges. This suggests that the $b$
2427: and $\beta$ field equations have been used to eliminate $\dot{c}$
2428: \eqa \dot{c} = - 2 \psi \g - {2 \over 1-2 h} \dot{h} c;\qquad
2429: \dot\g = -\frac{i\dot\psi c}{1-2h}. \eqae However, then one
2430: obtains time derivatives of $h$ and $\psi$. Another problem is
2431: that we seem to have too many factors of $(1-2 h)$ but this could
2432: be repaired by redefining fields. To resolve these issues we
2433: first construct $Q$ by Hamiltonian methods. This is a very
2434: general approach which only uses as input the first-class
2435: constraints of the classical theory, and which provides a quantum
2436: action in phase space and a Hamiltonian BRST charge with in
2437: general many more fields than in the usual (Lagrangian)
2438: formulation. Eliminating nonpropagating fields we should regain
2439: the Lagrangian BRST charge $Q$ in~(\ref{is}). Let's see how this
2440: works out.
2441:
2442: In the Hamiltonian framework of Fradkin and Vilkovisky (and
2443: others) the quantum action is of the form \eqa L = \dot{q}^i p_i
2444: - H + \{ Q_H, \psi_g \}, \label{Ham} \eqae where $q^i$ denotes all
2445: fields: classical fields (including the Lagrange multipliers $h$
2446: and $\psi$), ghosts and antighosts. The $p_i$ are canonical
2447: momenta \emph{for all of them}. (So, for example, there are
2448: canonical momenta for the ghosts, and separate canonical momenta
2449: for the antighosts). The BRST charge $Q_H$ should be nilpotent
2450: \eqa \{ Q_H , Q_H \} =0. \eqae The full quantum Hamiltonian $H$ is
2451: constructed from Dirac's Hamiltonian $H_D$ such that
2452: \begin{equation}
2453: [H, Q_H] =0.
2454: \end{equation}
2455: Neither $H_D$ nor $H$ should depend on Lagrange multipliers, and
2456: will be constructed below. The BRST charge is in general given by
2457: \eqa\label{ingen} Q_H = c^\a \varphi_\a + p^\mu (B) \pi_\mu (\l)
2458: - {1 \over 2} c^\beta c^\a f_{\a\b}{}^\g p_\g (-)^\beta ,\eqae
2459: where $\varphi_\a$ are the first-class constraints,
2460: $f_{\a\b}{}^\g$ the structure functions defined by
2461: $\{\varphi_\a,\varphi_\b\}=f_{\a\b}{}^\g\varphi_\g$, and $\l^\mu$
2462: are the Lagrange multipliers (classical fields which appears in
2463: the classical action without time derivatives), and $B$ denotes
2464: the antighosts. The structure functions should only depend on
2465: $p_i$ and $q^i$ but not on $\l^\mu$. Usually one has to take
2466: suitable linear combinations of local symmetries and add
2467: so-called equation-of-motion symmetries to achieve this. We shall
2468: demonstrate this in our model.
2469:
2470: The BRST invariance of the action in~(\ref{Ham}) is almost
2471: obvious: each of the 3 terms is separately invariant due to the
2472: relations $\int \frac{d}{dt}Q_H dt =0$, $[Q_H,H]=0$ and
2473: $\{Q_H,Q_H\}=0$.
2474:
2475: We start again from the classical gauge invariant action
2476: in~(\ref{1thirtythree})
2477: \begin{equation}
2478: L=\frac12 (1-2h){\dot\varphi}^2+\frac{i}{2}\lambda {\dot \lambda} -
2479: i\psi{\dot \varphi}\lambda.
2480: \end{equation}
2481: The primary constraints are $p_h=0$, $\pi_{\psi}=0$ and
2482: $\pi_{\lambda}+ \frac{i}{2}\lambda=0$, and the naive Hamiltonian
2483: with all primary constraints added, reads
2484: \begin{equation}
2485: H_{\rm naive}=\frac{1}{1-2h}\frac12\left( p+i\psi\lambda\right)^2+
2486: ap_h+\alpha\pi_{\psi}+\eta(\pi_{\lambda}+\frac{i}{2}\lambda).
2487: \end{equation}
2488: The functions $a(t)$, $\alpha(t)$ and $\eta (t)$ are arbitrary
2489: Lagrange multipliers which enforce the primary constraints. By a
2490: redefinition of $\eta$ we can replace $\lambda$ in the first term
2491: by the field $\frac12 \lambda+i\pi_{\lambda}$ which anticommutes
2492: with the constraint $\pi_{\lambda} +\frac{i}{2}\lambda $. This
2493: will simplify the analysis.
2494:
2495: Conservation of the 3 primary constraints yields two secondary
2496: constraints and fixes one Lagrange multiplier
2497: \begin{equation}
2498: p^2=0,\quad p(\pi_{\lambda} -\frac{i}{2}\lambda)=0,
2499: \quad \eta (t) =0.
2500: \end{equation}
2501: Both secondary constraints are preserved in time
2502: because $\pi_{\lambda} -\frac{i}{2}\lambda$ anticommutes with the constraint
2503: $\pi_{\lambda} +\frac{i}{2}\lambda$. The Hamiltonian can be rewritten as
2504: \begin{equation}
2505: H_{\rm naive}=(1+2H)\frac12 p^2+i\Psi(\frac{\lambda}{2} + i\pi_{\lambda})
2506: p+ap_h+\alpha\pi_{\psi}.
2507: \end{equation}
2508: Here
2509: \begin{equation}\frac{1}{1-2h}=1+2H \quad\text{and}\quad
2510: \frac{\psi}{1-2h}=\Psi,
2511: \end{equation}
2512: so we encounter again the fields $H$ and $\Psi$
2513: of~(\ref{dvatr}). From now on we use $G\equiv1+2H$ as
2514: gravitational field, and $\Psi$ as gravitino. Hence, the Dirac
2515: Hamiltonian (the Hamiltonian on the constraint surface) vanishes
2516: \begin{equation}
2517: H_{D}=0.
2518: \end{equation}
2519: and there are 5 constraints: one second class constraint
2520: \begin{equation}\label{secclass}
2521: \pi_{\lambda} +\frac{i}{2}\lambda=0,
2522: \end{equation}
2523: and four first-class constraints
2524: \begin{equation}\label{four1}
2525: p_h=\pi_{\psi}=p^2=p(\pi_{\lambda} -\frac{i}{2}\lambda)=0.
2526: \end{equation}
2527: Thus the Dirac brackets in (2.19) remain valid for supergravity.
2528:
2529: The constraint $p^2=0$ generates diffeomorphisms. One might expect
2530: that they are given by \eqa \delta\varphi=\xi{\dot \varphi}, \quad
2531: \delta p=\xi{\dot p}, \quad \delta\lambda=\xi{\dot \lambda},\quad
2532: \delta\pi=\xi{\dot \pi}, \nonumber \\
2533: \delta H= \frac{d}{dt}\left[ \frac12(1+2H)\xi\right],\quad
2534: \delta\Psi = \frac{d}{dt}(\xi\Psi),\label{delhal} \eqae because
2535: these transformations leave~(\ref{1Gaussian}) invariant. However
2536: in the Hamiltonian tranformation laws, no time derivatives are
2537: allowed. Moreover one expects that $p^2$ can only act on
2538: $\varphi$ (and perhaps $p$ and $H$), but not an $\l$, $\pi$ and
2539: $\psi$. We now perform a series of modifications of the
2540: transformation rules which cast~(\ref{delhal}) into the expected
2541: form.
2542:
2543: By adding terms proportional to the $p$ and $\varphi$ field
2544: equations to $\d\varphi$ and $\d p$, respectively (so-called
2545: equation of motion symmetries), namely, $\delta\phi=-\xi\tfrac{\del S}{\del p}$
2546: and $\delta p=\xi\tfrac{\del S}{\del \phi}$, one obtains \eqa
2547: &&\delta\varphi = (1+2H)\xi p - \xi \Psi (\pi - \frac{i}{2}\lambda),\nonumber \\
2548: &&\delta p =0. \eqae Similarly, adding terms proportional to the $\pi$ and $\l$ field
2549: equations to $\delta \lambda$ and $\delta \pi$ yields
2550: \begin{equation}
2551: \delta\lambda = - \xi p \Psi, \quad \delta\pi_\l = \frac{i}{2}\xi
2552: p \Psi.
2553: \end{equation}
2554: As a check of these last two results note that the constraint
2555: $\pi_\l + \frac{i}{2}\lambda$ is invariant. Finally we can remove
2556: the complicated last term in $\delta\varphi$ adding a local susy
2557: transformation in~(\ref{1Gaussian}) with parameter $\epsilon =-
2558: \xi\Psi$ to {\bf all} fields. This yields \eqa\label{pdv}
2559: \delta\varphi={\hat\xi}p, \quad \delta p=0, \quad
2560: \delta\lambda=0,\quad
2561: \delta\pi_\l=0, \nonumber\\
2562: \delta\Psi=0,\quad \delta (1+2H)= \frac{d}{dt}{\hat \xi},\quad
2563: {\hat \xi} = (1+2H)\xi. \eqae These are the transformations of the
2564: matter fields generated by $\frac{1}{2}p^2$. The classical gauge
2565: fields are $(1+2H)$ and $\Psi$ and they transform in general as
2566: \begin{equation}\label{AAA}
2567: \delta h^A = \frac{d}{dt} \epsilon^A + f^A_{\ BC} h_{\mu}^B \epsilon^C,
2568: \end{equation}
2569: where $\epsilon^A$ and $h^A_\mu$ correspond to $\hat \xi$ and
2570: $1+2H$ in the case of diffeomorphisms. The same results should be
2571: obtained by taking the brackets with $Q_H$.
2572:
2573: Next we consider the local susy generator in~(6.20) which
2574: we already multiply with the classical susy gauge parameter
2575: $\epsilon (t)$, hence $\epsilon (\pi_{\lambda}
2576: -\frac{i}{2}\lambda)p$. It generates the following classical transformation
2577: laws, obtained using the Dirac brackets,
2578: \begin{equation}\label{pdv2}
2579: \delta\varphi =- \epsilon (\pi_{\lambda} -\frac{i}{2}\lambda),\ \ \
2580: \delta p = 0, \ \ \
2581: \delta\lambda=-p\epsilon,\ \ \
2582: \delta\pi_{\lambda}=\frac{i}{2}p\epsilon.
2583: \end{equation}
2584: These are the transformation laws of~(\ref{1Gaussian}).
2585: The gauge fields should transform according to~(\ref{AAA})
2586: \begin{equation}\label{pdv3}
2587: \delta\Psi = {\dot \epsilon },\qquad \delta (1+2H)=
2588: -2i\epsilon\Psi.
2589: \end{equation}
2590: (The factor 2 in $-2i\epsilon\Psi$ comes from the two terms in
2591: $f^{A}_{\ \ BC}h_{\mu}^B\epsilon^C$ with $h_{\mu}^B=\Psi$ and
2592: $\epsilon^C=\epsilon$, or vice-versa). Also these rules agree
2593: with~(\ref{1Gaussian}). (Note that the local classical gauge
2594: algebra based on~(\ref{pdv}), (\ref{pdv2}) and (\ref{pdv3})
2595: closes, and that only the commutator of two local supersymmetry
2596: transformations is nonzero).
2597:
2598: The local gauge algebra of the transformation in~(\ref{pdv})
2599: and~(\ref{pdv2}),~(\ref{pdv3}) has now structure functions which
2600: are independent of the Lagrange multipliers $h$ and $\psi$ (or
2601: $H$ and $\Psi$), just as required for a Hamiltonian treatment.
2602: Having shown that the two first class constraints $\tfrac{1}{2}p^2$ and
2603: $p(\pi_{\lambda} -\frac{i}{2}\lambda)$ generate indeed the local
2604: symmetries of the classical phase space action
2605: in~(\ref{1Gaussian}), we now proceed with the construction of the
2606: BRST charge $Q_H$ and the quantum action.
2607:
2608: The BRST charge is given by \eqa\label{pdv4} Q_H=\frac12c p^2 -
2609: i\gamma p(\pi_{\lambda} -\frac{i}{2}\lambda)
2610: +p_bp_G+\pi_{\beta}\pi_{\Psi} - i\pi_c\gamma\gamma, \eqae where
2611: $G=1+2H$. We denote antihermitian conjugate momenta by $\pi$ as
2612: in $\pi_{\lambda}$, $\pi_{\Psi}$, $\pi_c$, $\pi_{\beta}$, but
2613: hermitian momenta by $p$ as in $p_G$, $p_b$, $p_\g$. By $p$
2614: in~(\ref{pdv4}) we mean $p_{\varphi}$, as before. (Recall that all
2615: ghosts are real, but the antighosts $b$ and $\b$ are
2616: antihermitian). The BRST charge is real and anticommuting. The
2617: first four terms contain the four first-class constraints, and
2618: the last term in~(\ref{pdv4}) comes from the last term
2619: in~(\ref{ingen}). The coefficients of the terms with $p_b$ and
2620: $\pi_\b$ are not fixed by nilpotency. It is easiest to fix the
2621: coefficients and signs by working out the transformation rules
2622: and fitting to the results obtained earlier, although in
2623: principle we need not follow this path since all terms are well
2624: defined. One can prove the nilpotency of the BRST laws by
2625: directly evaluating $\{Q_H,Q_H\}$, using the equal-time canonical
2626: (anti)commutations relations.
2627:
2628: Defining BRST transformations by
2629: \begin{equation}\label{ptr}
2630: \delta_{B}\Phi = -i[\Lambda Q_H,\Phi]=-i[\Phi,Q_H\L],
2631: \end{equation}
2632: for any field $\Phi$, we find the following results
2633: \begin{equation}\label{delb}
2634: \begin{array}{l}
2635: \delta_B\varphi = c p \Lambda - i (\pi_{\lambda} -\frac{i}{2}\lambda)\gamma\Lambda, \\
2636: \delta_{B}\lambda = i\gamma p \Lambda, \\
2637: \delta_{B}G = p_b \Lambda,\\
2638: \delta_{B} c = i\gamma \gamma \Lambda,\\
2639: \delta_{B}\gamma=0,\\
2640: \d_B\pi_c =-\frac12 p^2 \L\\
2641: \d_B p_G = \d_B\pi_\psi = \d_Bp_b=\d_B\pi_\b = 0,
2642: \end{array}
2643: \begin{array}{l}
2644: \delta_{B}p=0,\\
2645: \delta_{B}\pi_{\lambda} =\frac12 \g p \L, \\
2646: \delta_B{\Psi}=-\pi_{\beta} \Lambda, \\
2647: \delta_{B}b=-{\Lambda} p_G, \\
2648: \delta_{B}\beta=\pi_\psi\Lambda, \\
2649: \d_B p_\g
2650: =2i\pi_c\g\L+ip(\pi_\l-\frac{i}{2}\l).
2651: \end{array}
2652: \end{equation}
2653: On the classical fields these rules agree with the classical gauge
2654: transformations with $\xi=c\L$ and $\epsilon=-i\g\L$. In
2655: principle one should use Dirac brackets to obtain these
2656: transformation rules. These Dirac brackets can be constructed
2657: from the second class constraint in~(\ref{secclass}), but because
2658: the second class constraint commutes with the first class
2659: constraints, the results are the same\footnote{
2660: Only in the exceptional case that the Poisson bracket of a first
2661: and a second class constraint gives a square of a second class
2662: constraint, one would get a different answer, and in that case
2663: one would need to use Dirac brackets. This situation does not
2664: occur in our model.}.
2665:
2666: The rules in~(\ref{delb}) are nilpotent. In fact, we could have
2667: used nilpotency of $Q_H$ to derive $\delta_B c$ and
2668: $\delta_B\gamma$ from $\delta_B\varphi$, namely as follows \eqa
2669: &&\delta_B^2\varphi =0 = \delta_B\left[ cp-i\gamma(\pi_{\lambda} -
2670: \frac{i}{2}\lambda)\right]\nonumber\\
2671: &&\qquad =(\delta_B c)p - i(\delta_B\gamma)(\pi_{\lambda}
2672: -\frac{i}{2}\lambda)- i\gamma (\gamma
2673: p\Lambda)\;\Rightarrow\;\delta_B c =i\gamma\gamma\Lambda,\ \
2674: \delta_B\gamma=0. \eqae The result $\delta_B\gamma=0$ can also be
2675: derived from $\delta_B^2\lambda=0$.
2676:
2677: Comparison of $\d_Bb$ and $\d_B\b$ in~(\ref{delb})
2678: and~(\ref{contr}) reveals that the BRST auxiliary fields are the
2679: canonical momenta of the Lagrange multipliers
2680: \begin{equation}
2681: p_G=-d,\qquad \pi_{\Psi}=-\Delta.
2682: \end{equation}
2683:
2684: Finally we construct the quantum action in the Hamiltonian approach
2685: \begin{equation}
2686: L={\dot q}^ip_i-H+\{ Q_H,\psi_g\}.
2687: \end{equation}
2688: As ``gauge-fixing fermion'' we take the following hermitian anticommuting expression
2689: \begin{equation}\label{gaugefix}
2690: \psi_g = -i(b(G-1)+G\pi_c)+(-i\beta\Psi-\Psi p_{\gamma}).
2691: \end{equation}
2692:
2693: The quantum Hamiltonian $H_H$ which commutes with $Q_H$ vanishes in our case
2694: (and in any gravitational theory) because $H_D=0$ (see (6.23)). We denote the gravitational field
2695: $1+2H$ by $G$, and find then for the quantum Hamiltonian
2696: \begin{equation}\label{acti}
2697: L={\dot \varphi}p +{\dot \lambda}\pi_{\lambda}+{\dot G}p_G + {\dot
2698: \Psi}\pi_{\Psi} +{\dot c}\pi_c +{\dot b}p_b+{\dot
2699: \gamma}p_{\gamma} + {\dot \beta}\pi_{\beta}+\{Q_H, \psi_g\}.
2700: \end{equation}
2701: The evaluation of $\{Q_H,\psi_g\}$ is tedious but
2702: straightforward. For any pair of canonically conjugate variables
2703: we have $[p,q]=\frac1i$ or $\{\pi ,q\}=\frac1i$. This yields
2704: \eqa\label{esse} \{Q_H,-
2705: i(b+\pi_c)G+(-i\beta-p_{\gamma})\Psi\}&&\!\!\!\!\!\!\!\!=
2706: -p_G(G-1)-\frac12 Gp^2+(b+\pi_c)p_b-\pi_\Psi\Psi
2707: \nonumber\\
2708: &&\!\!\!\!\!\!\!\!- p(\pi_{\lambda} -\frac{i}{2}\lambda)\Psi -
2709: 2\pi_c\gamma\Psi+\pi_\b(-\b+ip_\g). \eqae The transformation laws
2710: $\d_BG=p_b\L$ and $\d_B\Psi=-\pi_\b\L$ which follow
2711: from~(\ref{ptr}) agree with the rules $\d_BG=({\dot
2712: c}+2\g\Psi)\L$ and $\d\Psi=-i{\dot \gamma}\L$ which we found from
2713: the classical transformation laws by substituting $\xi=c\L$ and
2714: $\epsilon=-i\g\L$, provided
2715: \begin{equation}\label{pigam}
2716: -p_b+{\dot c}+2\g\Psi=0, \qquad i\pi_\b+{\dot \g} =0.
2717: \end{equation}
2718: These should be algebraic field equations, and indeed they are
2719: the field equations of $\pi_c$ and $p_\g$. The relevant terms in
2720: the action are
2721: \begin{equation}
2722: L=(-p_b+{\dot c}+2\g\Psi)\pi_c + ({\dot \g}+i\pi_\b)p_\g.
2723: \end{equation}
2724: Integrating out $\pi_c$, $p_b$, $p_\g$ and $\pi_\b$
2725: imposes~(\ref{pigam}).
2726:
2727: The terms $-p_G(G-1)$ and $-\pi_{\Psi}\Psi$ are the gauge fixing
2728: terms. Thus to compare with the action as obtained from the
2729: Lagrangian BRST formalism we should use $b(G-1)+\Delta\Psi$
2730: in~(\ref{brsttran}) as gauge fixing term . However, we now got
2731: terms ${\dot G}p_G-p_GG$ and ${\dot \Psi}\pi_\Psi-\pi_\Psi\Psi$
2732: in the action. In Yang-Mills gauge theories one usually takes
2733: $\psi_g=b_a\chi^a+p_{c,a}A_0^a$ where $A_0^a$ is the
2734: time-component of the classical gauge field, and
2735: $\chi^a=\partial^kA_k^a$. Then one makes a change of integration
2736: variables $p(A_0)=kp(A_0)'$ and $b_a=kb_a'$ where $k$ is a
2737: constant. The superjacobian is unity, and taking the limit
2738: $k\rightarrow 0$ one arrives at the quantum action with
2739: relativistic unweighted gauge
2740: $\partial^{\mu}A_{\mu}^a=0$~\cite{Henne}. In our case the gauge
2741: choices were $G-1=0$ and $\Psi=0$. We used the same $\psi_g$ but
2742: with both $\chi^a$ and $A^a_0$ equal to $2H$ and $\Psi$. This led
2743: to~(\ref{gaugefix}), but it is clear that we cannot rescale
2744: $A_0^a\sim(G$ and $\Psi)$ but keep fixed $\chi\sim(G$ and
2745: $\Psi)$. In fact, we could have noticed before that there is
2746: something wrong with $\psi_g$ in~(\ref{gaugefix}): the terms have
2747: different dimensions.
2748:
2749: The resolution is also clear: drop the terms with $b$ and $\b$
2750: in~(\ref{gaugefix}). This removes the offending terms
2751: $-p_GG-\pi_\Psi\Psi$ and also the terms $bp_b$ and
2752: $-\pi_\beta\beta$ in~(\ref{esse}). The action now becomes after
2753: eliminating $\pi_c$, $p_b$, $\pi_\g$ and $\pi_b$ \eqa
2754: \label{theac} L&\!\!=&\!\!{\dot \varphi} p + {\dot \l} \pi_\l
2755: -\frac12
2756: Gp^2-p(\pi_\l-\frac{i}{2} \l)\Psi\nonumber \\
2757: &+&\!\! {\dot G}p_G+{\dot \Psi}\pi_\Psi\\
2758: &+&\!\! {\dot b} ({\dot c}+2\g\Psi)+{\dot\b}(i{\dot \g}).
2759: \nonumber\eqae The first line is~(\ref{1Gaussian}), the second
2760: line is the gauge fixing term, and the third line the ghost
2761: action. So the action in~(\ref{acti}) from the Hamiltonian BRST
2762: formalism indeed agrees with the action from the Lagrangian BRST
2763: formalism if one chooses ${\dot G}$ and ${\dot \Psi}$ instead of
2764: $G$ and $\Psi$ as gauge fixing terms.
2765:
2766: We have seen that there is a Lagrangian and a Hamiltonian
2767: approach. The latter contains conjugate momenta for all
2768: variables, hence brackets can be defined and charges constructed.
2769: We saw that the gauges $G-1=0$ and $\Psi=0$ in the Hamiltonian
2770: approach led to the same results after solving the algebraic
2771: field equations for the canonical momenta as the gauge ${\dot
2772: G}=0$ and ${\dot \Psi}=0$ in the Lagrangian approach. (We recall
2773: that $G=1+2H=\frac{1}{1-2h}$ and $\Psi=\frac{\psi}{1-2h}$.) One
2774: might ask whether the Hamiltonian approach can also lead to the
2775: gauge $G-1=0$ and $\Psi=0$ in the Lagrangian approach
2776: (corresponding to $h=0$ and $\psi =0$ in~(\ref{pod})). This is
2777: indeed possible. One chooses as gauge fermion $\psi_g=0$. The
2778: action becomes then very simple \begin{equation}\label{simpac}
2779: L=\dot{\varphi}p+\dot{\l}\pi_\l+\dot{G}p_G+\dot{\Psi}p_{\Psi}+
2780: \dot{c}\pi_{c}+\dot{b}p_b+\dot{\g}p_\g + \dot{\b}\pi_\b.
2781: \end{equation}
2782: Next one factorizes the path integral into a minimal part with the
2783: fields $(p,\varphi)$, $(\l,\pi)$, $(c,\pi_c)$ and $(\g,p_\g)$,
2784: and a nonminimal part with the pairs $(G,p_G)$, $(\Psi,\pi_\Psi)$,
2785: $(b,p_b)$ and $(\b,\pi_\b)$. Finally one discards the latter, and
2786: \emph{reinterprets} the momenta $\pi_c$ and $p_\g$ as the
2787: Lagrangian antighosts $b$ and $\b$, respectively. This yields
2788: then the action in~(\ref{brsttran}) with $d$, $\Delta$, $h$ and
2789: $\psi$ integrated out (removed by their algebraic field
2790: equations).
2791:
2792: This concludes our discussion of BRST formalism applied to our
2793: simple model. We obtained one result which is a bit surprising
2794: (or interesting): the differentiated gauge choices
2795: $\dot{G}=\dot{\Psi}=0$ in the Lagrangian approach correspond
2796: directly to the gauge choices $G-1=\Psi=0$ in the Hamiltonian
2797: approach. On the other hand, the gauge choices $G-1=\Psi=0$ in
2798: the Lagrangian approach did not correspond in a direct way to the
2799: Hamiltonian approach (we had to discard a sector). String theory
2800: uses Lagrangian gauge choices corresponding to $G-1=\Psi=0$.
2801: Perhaps the corresponding differentiated gauge choices have
2802: advantages in certain respects.
2803:
2804: \begin{thebibliography}{99}
2805: \bibitem{Berezinfa} F.A. Berezin, The method of second quantization, Academic Press 1966. In this book
2806: manipulations with Grassmann variables are defined. See also F.A. Berezin, Introduction to Super Analysis,
2807: Reidel, 1987.
2808: \bibitem{rfeynman} R. Feynman and A. Hibbs, Quantum mechanics and path integrals, McGraw Hill 1965.
2809: Feynman pioneered the path integral in physics, and this is the standard reference.
2810:
2811: \bibitem{twoi} Two introductory texts which the reader may try to tackle after reading these lectures are J. Bagger and J. Wess, Supersymmetry and Supergravity, Princeton University Press,
2812: 1992, and M.F. Sohnius, Introducing supersymmetry, {\it Phys.\ Rep.} {\bf 128} (1985) 39.
2813: More advanced introductions are
2814: P. van Nieuwenhuizen, Supergravity, {\it Phys.\ Rep.} {\bf 68} (1981) 189, and
2815: S.J. Gates, M.T. Grisaru, M. Ro\v{c}ek and W. Siegel, Superspace, One thousand and one lessons in supersymmetry.
2816: \bibitem{Henne} For a discussion of the Hamiltonian form of BRST symmetry, see
2817: M. Henneaux, {\it Phys. Rep.} {\bf 126} (1985) 1. A good
2818: introduction for mathematicians to BRST symmetry and general
2819: properties of gauge systems is M. Henneaux and C. Teitelboim,
2820: Quantization of Gauge Systems, Princeton University Press 1992.
2821: We used sections 16.5.6, 16.54b, 19.1.7 and 19.1.8 for discussions
2822: at the very end of chapter~6.
2823: \bibitem{spinning} Spinning point particles have been discussed in L. Brink, S. Deser,
2824: B. Zumino, P. Di Vecchia and P. Howe, {\it Phys.\ Lett.} {\bf 64B} (1976) 435; see also
2825: F.A. Berezin and M.S. Marinov, {\it Ann. \ Phys. (N.Y.)} {\bf 104} (1977) 336. A Hamiltonian
2826: path integral treatment of relativistic supersymmetric particles is given in M. Henneaux and
2827: C. Teitelboim, {\it Ann. \ Phys. (N.Y.)} {\bf 143} (1982) 127.
2828: \end{thebibliography}
2829: \end{document}
2830:
2831: \begin{array}{lllll}
2832: && [ G_a , G_b ] = f_{ab}{}^c G_c & G_a = (p (h), p (\psi) ; T, J) \nn
2833: && [ \varphi_a , \varphi_b ] = f_{ab}{}^c G_c & \eta^a = (p (b) , - p(\beta) ; c, \g ) \nn
2834: && [ p_\a , \varphi_b ] = f_{ab}{}^c G_c & p_a = ( b, \beta ; p(c) , p(\g) ) \end{array} \nn
2835: && f_{AB}{}^c G_c = \left( \begin{array}{ccccc} 0 & 0 & {2 \over 1-2 h}T & {2J \over 1-2 h} \\ 0 & 0 & 0 & 0 \\ & & 0 & 0 \end{array} \right) \nn
2836: && \d_\varphi = {2 pc \over 1-2 h} = 2c \left( \dot\varphi + {i \psi \l \over 1-2 h} \right)
2837: \eqae
2838:
2839: %%%%%%%%%%%%%%%5
2840:
2841: One is left with
2842: \begin{equation}
2843: L={\dot \varphi}p +{\dot \lambda}\pi_{\lambda}+ {\dot c}\pi_c +
2844: {\dot \gamma}p_{\gamma} -p_HG+i\pi_{\Psi}\Psi+\frac12Gp^2-p
2845: (\pi_{\lambda} -\frac{i}{2}\lambda)\Psi
2846: +\pi_cp_b-2\pi_c\gamma\Psi+i\pi_{\beta} p_{\gamma}
2847: \end{equation}
2848: Integration over $\pi_c,p_b$ and over $\pi_{\beta},p_{\gamma}$
2849: yields for the ghost sector \eqa
2850: {\dot c}\pi_c +{\dot \gamma}p_{\gamma}+\pi_cp_b- 2\pi_c\gamma\Psi+i\pi_{\beta} p_{\gamma}
2851: =(\pi_c)(p_b-{\dot c}-2\gamma\Psi)+(i\pi_{\beta}+{\dot
2852: \gamma})p_{\gamma} \nonumber\\+ {\dot \gamma}({\dot
2853: c}-2\gamma\Psi)-{\dot \gamma}p_{\gamma} \eqae
2854:
2855:
2856: %\eqa
2857: %Q_H = \left( c {p^2 \over 1-2 h} + i \g {\l p \over 1-2 h} \right) + p (b) p(h) -
2858: %p (\beta) p ( \psi)
2859: %\eqae
2860:
2861: One should now go on, and work out the explicit form of the
2862: quantum Hamiltonian in (\ref{Ham}). Then one could eliminate
2863: several nonpropagating fields (fields without derivatives) and
2864: show that the results agree with the Lagrangian results in
2865: (\ref{is}) and (\ref{class}). We leave these steps as
2866: exercises. The reader who has come to this point should be able
2867: to read the physics literature~\cite{twoi}.
2868:
2869: