0712.0615/ssm.tex
1: \documentclass[11pt]{article}
2: \usepackage{amsmath}
3: \usepackage{latexsym}
4: \usepackage{graphicx}
5: \usepackage{bbm}
6: \usepackage{color}
7: \usepackage{euscript}
8: \usepackage{amsfonts}
9: \usepackage{cite}
10: 
11: \textwidth=6.0in
12: \hoffset=-0.55in
13: %\def\baselinestretch{1.4}
14: 
15: \setlength{\parskip}{10pt plus 1pt minus 1pt}
16: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
17: 
18: %%%%% number equations by section %%%%%%%%
19: \makeatletter
20: \@addtoreset{equation}{section}
21: \makeatother
22: \renewcommand{\theequation}{\thesection.\arabic{equation}}
23: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
24: \def \AP {{\it Annals Phys.}, }
25: \def \APP {{\it Astropart. Phys.}, }
26: \def \ATMP {{\it Adv. Theor. Math. Phys.}, }
27: \def \CMP {{\it Commun. Math. Phys.}, }
28: \def \CQG {{\it Class. Quant. Grav.}, }
29: \def \IJMP {{\it Int. J. Mod. Phys.}, }
30: \def \JCAP {{\it JCAP}, }
31: \def \JGP {{\it J. Geom. Phys.}, }
32: \def \JHEP {{\it JHEP}, }
33: \def \JMP {{\it J. Math. Phys.}, }
34: \def \MPL {{\it Mod. Phys. Lett.}, }
35: \def \NC {{\it Nuovo Cim.}, }
36: \def \NCL {{\it Nuovo Cim. Lett.}, }
37: \def \NP {{\it Nucl. Phys.}, }
38: \def \NPPS {{\it Nucl. Phys. Proc. Suppl.}, }
39: \def \PL {{\it Phys. Lett.}, }
40: \def \PR {{\it Phys. Rev.}, }
41: \def \PREP {{\it Phys. Reports}, }
42: \def \PRL {{\it Phys. Rev. Lett.}, }
43: \def \PTP {{\it Prog. Theor. Phys.}, }
44: \def \RMP {{\it Rev. Mod. Phys.}, }
45: \def \SHEP {{\it Surveys High Energ. Phys.}, }
46: \def \SPTP {{\it Suppl. Prog. Theor. Phys.}, }
47: 
48: 
49: \def\beq{\begin{equation}}
50: \def\eeq{\end{equation}}
51: \def\be{\begin{equation}}
52: \def\ee{\end{equation}}
53: 
54: \def\half{{1 \over 2}}
55: \def\bea{\begin{eqnarray}}
56: \def\eea{\end{eqnarray}}
57: \def\ft#1#2{{\textstyle{\frac{\scriptstyle #1}{\scriptstyle #2}}}}
58: \def\fft#1#2{\frac{#1}{#2}}
59: \def\half{{\ft{1}{2}}}
60: \let\a=\alpha
61: \let\b=\beta
62: \let\g=\gamma 
63: \let\d=\delta 
64: \let\ep=\epsilon
65: \let\vep=\varepsilon
66: \let\z=\zeta 
67: \let\et=\eta 
68: \let\th=\theta
69: \let\vt=\vartheta 
70: \let\i=\iota 
71: \let\k=\kappa
72: \let\vk=\varkappa
73: \let\l=\lambda 
74: \let\m=\mu 
75: \let\n=\nu  
76: \let\r=\rho
77: \let\vr=\varrho
78: \let\s=\sigma 
79: \let\vs=\varsigma
80: \let\t=\tau 
81: \let\u=\upsilon 
82: \let\ph=\phi
83: \let\vph=\varphi 
84: \let\ch=\chi 
85: \let\ps=\psi
86: \let\w=\omega      
87: \let\G=\Gamma 
88: \let\D=\Delta 
89: \let\T=\Theta 
90: \let\L=\Lambda
91: \let\X=\Xi 
92: \let\P=\Pi 
93: \let\S=\Sigma 
94: \let\U=\Upsilon 
95: \let\Ph=\Phi 
96: \let\Ps=\Psi
97: \let\C=\Chi 
98: \let\W=\Omega     
99: \let\la=\label \let\ci=\cite \let\re=\ref
100: \let\fr=\frac \let\bl=\bigl \let\br=\bigr
101: \let\Br=\Bigr \let\Bl=\Bigl 
102: \let\bm=\bibitem
103: \let\na=\nabla
104: \newcommand{\tr}{{\rm tr} }
105: \newcommand{\Tr}{{\rm Tr} } 
106: \def\im{{{\rm i}}}
107: \def\jm{{{\rm j}}}
108: \def\km{{{\rm k}}}
109: \def\R{{\mathbbm R}}
110: \def\C{{{\mathbbm C}}}
111: \def\H{{{\mathbbm H}}}
112: \def\RP{{{\mathbbm R}{\mathbbm P}}} 
113: \def\CP{{{\mathbbm C}{\mathbbm P}}} 
114: \def\HP{{{\mathbbm H}{\mathbbm P}}}
115: \def\Z{{\mathbbm Z}} 
116: 
117: \def\cK{{\cal K}}
118: \def\cG{{\cal G}}
119: \def\cH{{\cal H}}
120: \def\cV{{\cal V}}
121: \def\cI{{\cal I}}
122: \def\cJ{{\cal J}}
123: \def\bG{{\bf G}}
124: \def\bH{{\bf H}}
125: \def\bK{{\bf K}}
126: \def\del{{\partial}}
127: \def\bm{\bibitem}
128: \def\sst#1{{\scriptscriptstyle #1}}
129: \def\n{{\sst (n)}}
130: \def\m{{\sst (m)}}
131: \def\p{{\sst (p)}}
132: \def\nn{\nonumber}
133: \def\td{\tilde}
134: \def\wtd{\widetilde}
135: \def\bep{{\bar\ep{\,}}}
136: \def\ie{{\it i.e.\ }} 
137: 
138: \thispagestyle{empty}
139: \newcount\hour \newcount\minute
140: \hour=\time  \divide \hour by 60
141: \minute=\time
142: \loop \ifnum \minute > 59 \advance \minute by -60 \repeat
143: \def\nowtwelve{\ifnum \hour<13 \number\hour:% 		% supresses leading 0's
144:                       \ifnum \minute<10 0\fi%		% so add it it
145:                       \number\minute
146:                       \ifnum \hour<12 \ A.M.\else \ P.M.\fi
147: 	 \else \advance \hour by -12 \number\hour:% 	% supresses leading 0's
148:                       \ifnum \minute<10 0\fi%		% add it in
149:                       \number\minute \ P.M.\fi}
150: \def\nowtwentyfour{\ifnum \hour<10 0\fi% 		% need a leading 0 
151: 		\number\hour:% 				% supresses leading 0's
152:          	\ifnum \minute<10 0\fi% 		% add it in
153:          	\number\minute}
154: \def\now{\nowtwelve}
155: 
156: 
157: 
158: \begin{document}
159: 
160: 
161: 
162: \voffset=0.3truein
163: \hfuzz=100pt
164: 
165: 
166: \title{Infinite-Dimensional Symmetries of Two-Dimensional Coset Models
167:  Coupled to Gravity}
168: \author{\Large H. L\"u,$^1$ Malcolm J. Perry$^2$ and
169: C.N. Pope$^{1,2}$
170: \\ \\ \\
171: ${}^1$ George P. \& Cynthia W. Mitchell Institute for Fundamental Physics,\\
172:          Texas A\&M University,
173:          College Station,
174:          TX 77843-4242,
175:          USA.\\ \\
176: ${}^2$ DAMTP, Centre for Mathematical Sciences,\\
177:          University of Cambridge,
178:          Wilberforce Road,
179:          Cambridge CB3 0WA,
180:          England.\\ \\ \\  \\} 
181: 
182: \date{\empty}
183: \maketitle
184: 
185: 
186: %\includegraphics[scale = 0.05, bb= 0 -8500 100 -8600 ]{arms}
187: %\includegraphics[scale = 0.17, bb= -2000 -2500 -2030 -2530 ]{tamuseal}
188: 
189: \includegraphics[scale = 0.05, bb= 0 -6800 100 -8600 ]{arms}
190: \includegraphics[scale = 0.17, bb= -1900 -2000 0 -1800 ]{tamuseal}
191: 
192: \vskip -5.5truein
193: 
194: \centerline{DAMTP-2007-115\hskip 1.5truein MIFP-07-31}
195: \bigskip\bigskip
196: \centerline{\bf{ArXiv:0712.0615}}
197: 
198: \vskip 3.5truein
199: 
200: 
201: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
202: %%%%%%Comment out the next line in the final version%%
203: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
204: %\centerline{\now}
205: 
206: 
207: 
208: 
209: 
210: 
211: 
212: \begin{abstract} 
213: 
214:   In an earlier paper we studied the infinite-dimensional symmetries
215: of symmetric-space sigma models (SSMs) in a flat two-dimensional spacetime.
216: Here, we extend our investigation to the case of two-dimensional SSMs
217: coupled to gravity.  These theories arise from the toroidal reduction
218: of higher-dimensional gravity and supergravities to two dimensions.
219: We construct explicit expressions for the symmetry transformations under
220: the affine Kac-Moody extension $\hat\cG$ that arises when starting from
221: a $\cG/\cH$ coset model.  We also construct further explicit symmetry
222: transformations that correspond to the modes $L_n$ of a Virasoro subalgebra
223: with $n\ge -1$.
224: 
225: 
226: 
227: \end{abstract}
228: 
229: \thispagestyle{empty}
230: 
231: \pagebreak
232: \voffset=0pt
233: \setcounter{page}{1}
234: 
235: \tableofcontents
236: 
237: \addtocontents{toc}{\protect\setcounter{tocdepth}{2}}
238: 
239: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
240: %%%%Document specific macros go here%%%%
241: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
242: \def\V{\mathcal{V}}
243: \def\hV{\hat{\mathcal{V}}}
244: 
245: \newpage
246: 
247: \section{Introduction}
248: 
249:    The study of supergravity theories, and their symmetries, have played
250: a very important r\^ole in uncovering the underlying structures of
251: string theory.  Especially significant are the U-duality symmetries of
252: the string, which have their origin in the classical global symmetries
253: exhibited by eleven-dimensional supergravity and type IIA and IIB supergravity
254: after toroidal dimensional reduction.  For example, if one reduces
255: eleven-dimensional supergravity on an $n$-torus, for $n\le8$, the
256: resulting $D=(11-n)$-dimensional theory exhibits a global $E_{n}$ symmetry
257: \cite{crejul,cj2,cjlp}.  In the cases $n\ge3$ this symmetry arises in quite a
258: subtle way, involving an interplay between the original eleven-dimensional
259: metric and the 3-form potential.  The global symmetry can be understood 
260: most simply by first focusing attention on the scalar-field sector of the
261: dimensionally-reduced theory.  The scalars are described by a non-linear
262: sigma model, and in fact the cosets that arise are always symmetric spaces. 
263: The case of reduction to three dimensions is in many ways the most elegant,
264: because all of the bosonic fields in the theory (aside from the metric) 
265: are now scalars, and so one has just an $E_8/O(16)$ symmetric-space 
266: sigma model in a gravity background.
267: 
268:    In view of the large $E_8$ symmetry that one finds after reduction to
269: three dimensions, it is natural to push further and investigate the symmetries
270: after further reduction to two dimensions, and even beyond.  It turns out that
271: the analysis for a reduction to two dimensions is considerably more
272: complicated than the higher-dimensional ones.  There are two striking new
273: features that lead to this complexity.  The first is that, unlike a
274: reduction to $D\ge3$ dimensions, one can no longer use a reduction
275: scheme in which the metric is reduced from an Einstein-frame metric in the
276: higher dimension to an Einstein-frame metric in the lower dimension.  (In
277: the Einstein conformal frame, the Lagrangian for gravity itself takes the
278: form ${\cal L} \sim \sqrt{-g} R$, with no scalar conformal factor.)  The
279: inability to reach the Einstein conformal frame in two dimensions is
280: intimately connected to the fact that $\sqrt{-g} R$ is a conformal
281: invariant in two dimensions.   
282: 
283:    The second striking new feature is that an axionic
284: scalar field (\ie a scalar appearing everywhere covered by a derivative)
285: can be dualised to give another axionic scalar field in the
286: special case of two dimensions.  This has the remarkable consequence that
287: the global symmetry group actually becomes infinite in dimension.  This
288: was seen long ago by Geroch, in the context of four-dimensional gravity
289: reduced to two.   There are degrees of freedom
290: in two dimensions that are described by the sigma model $SL(2,\R)/O(2)$,
291: and under dualisation this yields another $SL(2,\R)/O(2)$ sigma model.
292: Geroch showed that the two associated global $SL(2,\R)$ symmetries do
293: not commute, and that if one takes repeated commutators of the two
294: sets of transformations, an infinite-dimensional algebra results \cite{geroch}.
295: The precise nature of this symmetry, now known as the {\it Geroch Group},
296: was not uncovered in \cite{geroch}.
297: 
298:    It is of considerable interest, therefore, to study the general
299: case of symmetric-space sigma model in a three-dimensional gravitational
300: background, after performing a further circle reduction to two dimensions.
301: (We shall use the acronym SSM to denote a symmetric-space sigma model.)
302: One arrives at a system in two dimensions that comprises an SSM coupled to 
303: gravity, together with an additional scalar field which can be thought
304: of as the Kaluza-Klein scalar for the reduction from three to two dimensions.
305: 
306:    In a previous paper \cite{lupepo}, we studied the global symmetries
307: of the simpler situation where one has an SSM in a purely flat two-dimensional
308: background.  Our goal in the present paper is to extend our analysis to the
309: full gravitationally-coupled system that arises when one reduces a 
310: gravity-coupled SSM from three to two dimensions.  
311: 
312:    There is quite a considerable earlier literature on the subject of 
313: the infinite
314: dimensional symmetries of two-dimensional symmetric-space sigma models,
315: both in the flat and the gravity-coupled cases (see, for example,
316: \cite{pol,lus,polus,kinchi,belzak,mais1,breitmais,dolan,devfair,yswu,jul,%
317: nicolai,nicjul},
318: some of which considers also principal chiral models).  In our previous 
319: work on the flat-space SSM case in
320: \cite{lupepo}, we found the work of Schwarz in \cite{schwarz1} to be
321: notably accessible, and we took that as our starting point.  Starting
322: from a two-dimensional flat-space SSM based on the coset $\cG/\cH$, 
323: Schwarz gave an explicit construction of symmetry transformations whose 
324: commutators generated a certain subalgebra $\hat\cG_H$ of the affine Kac-Moody 
325: extension $\hat\cG$ of $\cG$.  He also constructed additional symmetry 
326: transformations that formed a subalgebra of the centreless Virasoro algebra. 
327: We were then able to extend Schwarz's results, and obtain explicit
328: results exhibiting the full $\hat\cG$ Kac-Moody symmetry.
329: 
330:    In the present paper, we again take as our starting point the work
331: of Schwarz, who had extended his flat-space analysis to the gravitationally
332: coupled case in \cite{schwarz2}.  We develop Schwarz's construction by 
333: first improving on the removal of certain singularities in the transformation
334: rules, and then in addition we are able again to extend Schwarz's 
335: construction of explicit $\hat G_H$ transformations to the full set
336: of transformations for the entire Kac-Moody algebra $\hat\cG$.  Schwarz
337: did not succeed in generalising his construction of  Virasoro-type 
338: transformations to the gravitationally-couple case.  We succeed in doing
339: this too.  An essential feature now is that the Kaluza-Klein scalar coming
340: from the descent from three dimensions transforms under the Virasoro-type
341: symmetries, even though it is inert under the Kac-Moody transformations.
342: 
343:    We also discuss, in an appendix, how one can take a decoupling limit
344: of our results, to recover our earlier results in \cite{lupepo} on the
345: infinite-dimensional symmetries of two-dimensional sigma models in the
346: absence of gravity.
347: 
348: 
349: \section{Lax Equation and Infinite-Dimensional Symmetries}
350: 
351: \subsection{Reduction from three dimensions}
352: 
353:    We take as our starting point a symmetric-space non-linear sigma
354: model defined on the coset manifold $\bK=\bG/\bH$, and coupled to gravity 
355: in three spacetime dimensions.  The commutation relations
356: for the corresponding generators of the Lie algebra $\cG$ take the form
357: %%%%%
358: \be
359: [\cH,\cH]=\cH\,,\qquad [\cH,\cK]=\cK\,,\qquad
360:    [\cK,\cK]=\cH\,.\label{coset}
361: \ee
362: %%%%%
363: The condition that $\bK$ is a symmetric space is reflected in the absence of
364: $\cK$ generators on the right-hand side of the last commutation relation.
365: The symmetric-space algebra implies that there is an involution $\sharp$ 
366: under which
367: %%%%%
368: \be
369: \cK^\sharp = \cK\,,\qquad \cH^\sharp=-\cH\,.
370: \label{invol1}
371: \ee
372: %%%%%
373: In many cases, such as when $\bG=SL(n,\R)/O(n)$, the involution map is given
374: by Hermitean conjugation,
375: %%%%%
376: \be
377: \cK^\dagger=\cK\,,\qquad \cH^\dagger
378:  = -\cH\,.\label{invol2}
379: \ee
380: %%%%%
381: In some cases,
382: such as $\bG=E_{(8,8)}$, $\cH=O(16)$, the involution $\sharp$ is more involved. 
383: 
384:    The fields of
385: the sigma model $\bG/\bH$ may be parameterised by a coset representative
386: $\cV$, in terms of which we may define
387: %%%%%
388: \be
389: M=\cV^\sharp\, \cV\,,\qquad A= M^{-1} dM\,.\label{MAdef}
390: \ee
391: %%%%%
392: Under transformations 
393: %%%%%
394: \be
395: \cV\longrightarrow h\cV g\,,\label{gtrans}
396: \ee
397: %%%%%
398: where $g$ is a global element
399: in the group $\bG$ and $h$ is a 
400: local element in the denominator subgroup $\bH$, we have shall have
401: %%%%%
402: \be
403: M\longrightarrow g^\sharp M g\,,\qquad 
404:                    A\longrightarrow g^{-1} A g\,,\label{Mtrans}
405: \ee
406: %%%%%
407: since it follows from $\cH^\sharp =-\cH$ that $h^\sharp= h^{-1}$.
408: Henceforth, we shall consider for simplicity cases where the involution 
409: $\sharp$ is just Hermitean conjugation.  For the general case, all
410: occurrences of $\dagger$ should be replaced by $\sharp$.
411: 
412:    The Cartan-Maurer equation $d(M^{-1} dM)= -(M^{-1} dM)\wedge (M^{-1} dM)$
413: implies that the field strength for $A$ vanishes:
414: %%%%%
415: \be
416: F\equiv dA + A\wedge A=0\,.\label{cartmau0}
417: \ee
418: %%%%%
419: The Lagrangian for the three-dimensional model may be written as 
420: %%%%%
421: \be
422: {\cal L}_3 = \sqrt{-\hat g} \, 
423: \big(\hat R - \ft14\hat g^{MN}\, \tr(A_M A_N)\big)\,,
424: \label{d3lag}
425: \ee
426: %%%%%
427: where $\hat g_{MN}$ is the three-dimensional spacetime metric tensor.  The
428: sigma-model equations of motion are therefore given by
429: %%%%%
430: \be
431: \hat\nabla^M A_M=0\,.
432: \ee
433: %%%%%
434: 
435:    We now assume that the metric and sigma-model fields are all 
436: independent of one of the coordinates, which we shall denote by $z$.  We
437: may reduce the metric according to the Kaluza-Klein ansatz
438: %%%%%
439: \be
440: d\hat s_3^2 = e^\psi ds_2^2 + \rho^2\, dz^2\,.\label{metred}
441: \ee
442: %%%%%
443: Note that the two-dimensional field $\psi$ is redundant, in the sense
444: that it could be absorbed into a conformal rescaling of the two-dimensional
445: metric $ds_2^2$.  However, it it useful to retain it since it will allow
446: us later to take $ds_2^2$ to be just the Minkowski metric.  It is
447: straightforward to see that the three-dimensional Lagrangian (\ref{d3lag})
448: reduces to give
449: %%%%%
450: \be
451: {\cal L}_2 = \sqrt{-g}\, \rho\, \big(R-\ft14 g^{\mu\nu} \tr(A_\mu A_\nu) +
452:  \rho^{-1}\, g^{\mu\nu}  \del_\mu\rho\, \del_\nu\psi\big)\,,\label{d2lag}
453: \ee
454: %%%%
455: (after an integration by parts).
456: 
457:    The equations of motion that follow from varying $\cV$, $\psi$\, $\rho$
458: and $g_{\mu\nu}$ are, respectively,
459: %%%%%
460: \bea
461: \nabla^\mu(\rho\, A_\mu) &=&0\,,\label{Veom}\\
462: %%
463: \square\rho &=&0\,,\label{rhoeq}\\
464: %%
465: \square\psi &=& R -\ft14 g^{\mu\nu}\, \tr(A_\mu A_\nu)\,,\label{psieq}\\\
466: %%
467: 0&=& \del_\mu\rho\, \del_\nu\psi + \del_\nu\rho\, \del_\mu\psi -g_{\mu\nu}
468:   \del_\sigma\rho\del^\sigma\psi - 2\nabla_\mu\nabla_\nu\rho\nn\\
469: %% 
470: && -
471:   \ft12\tr \big(A_\mu A_\nu - 
472:    \ft12 g_{\mu\nu} \tr(A_\sigma A^\sigma)\big)\,,\label{einst}
473: \eea
474: %%%%%
475: (after using (\ref{rhoeq}) to simplify (\ref{einst}).
476: 
477:    Using general coordinate transformations, any two-dimensional metric
478: can be written, locally, as a conformal factor times the Minkowski metric.
479: Thus we may now take $ds_2^2$ to be the Minkowski metric, with $e^\psi$
480: as the required conformal factor.  It is convenient, furthermore, to
481: introduce light-cone coordinates $x^\pm$, so that we have 
482: $ds_2^2=2dx^+ dx^-$.  Note that the $++$ and $--$ components of the 
483: Einstein equation (\ref{einst}) can be used to solve for $\psi$, since
484: it gives
485: %%%%% 
486: \be
487: (\del_+\rho)\, \del_+\psi= \del_+^2\rho + \ft14 \tr(A_+^2)\,,\qquad
488: (\del_-\rho)\, \del_-\psi= \del_-^2\rho + \ft14 \tr(A_-^2)\,.\label{psisol}
489: \ee
490: %%%%%
491: Equation (\ref{psieq}) can now be seen to be a consequence of (\ref{Veom}),
492: (\ref{rhoeq}), (\ref{psisol}) and (\ref{cartmau0}), so the two-dimensional
493: equations reduce to solving
494: %%%%%
495: \bea
496: d(\rho{*A}) &=&0\,,\label{eom}\\
497: %%
498: \qquad dA+ A\wedge A&=&0\,,\label{cartmau}\\
499: %% 
500: \square\rho&=&0\,,\label{rhoeom}
501: \eea
502: %%%%%
503: with $\psi$ then being found using (\ref{psisol}).
504: 
505: 
506: \subsection{The Lax equation}\label{formsec}
507: 
508:    The use of the Lax equation in this context was discussed in \cite{pol},
509: and in a somewhat different, but related way, in \cite{nicolai}.
510: We shall begin by following Schwarz's discussion of the Lax-pair formulation 
511: in \cite{schwarz2}, except that we prefer
512: to use differential forms where possible, rather than light-cone
513: coordinates.  The equations (\ref{eom}) 
514: and (\ref{cartmau}) 
515: can both be derived from the integrability condition for a solution $X$ of
516: the Lax equation
517: %%%%%
518: \be
519: \tau (d+A) X= {*dX}\,.\label{Lax2}
520: \ee
521: %%%%%
522: By taking the appropriate linear
523: combination of this and its dual, we obtain\footnote{Note that $*^2=+1$ when
524: acting on 1-forms in signature $(1,1)$ spacetimes.  A summary of some further
525: useful properties of forms in two dimensions can be found in \cite{lupepo}.}
526: %%%%%
527: \be
528: dX X^{-1} = \fft{\tau}{1-\tau^2} \, {*A} + 
529:        \fft{\tau^2}{1-\tau^2} \, A\,.\label{Lax3}
530: \ee
531: %%%%%
532: Unlike the flat-space case where the spectral parameter $\tau$ is a constant,
533: here it must be allowed to have a specific dependence on the 
534: two-dimensional spacetime 
535: coordinates. It is convenient in what follows to parameterise $\tau$ in terms
536: of $\theta$ according to
537: %%%%%
538: \be
539: \tau=\tanh\ft12\theta\,,\label{tautheta}
540: \ee
541: %%%%%
542: which implies we shall also have
543: %%%%%
544: \be
545: s\equiv\sinh\theta = \fft{2\tau}{1-\tau^2}\,,\qquad
546:     c\equiv\cosh\theta = \fft{1+\tau^2}{1-\tau^2}\,.\label{scdef}
547: \ee
548: %%%%%
549: 
550:    In fact, one finds that the $x^\mu$ dependence of $\tau$ should
551: occur through the function $\rho(x^\mu)$, as follows:
552: %%%%%
553: \be
554: d\tau = \fft{\tau}{\rho}\, (c d\rho + s {*d}\rho)\,.\label{dtau}
555: \ee
556: %%%%%
557:  Equation (\ref{tautheta}) then implies
558: %%%%%
559: \be
560: d\theta = \fft{s}{\rho}\, (c d\rho + s {*d}\rho)\,.\label{dtheta}
561: \ee
562: %%%%%
563: With these preliminaries, it is now an elementary calculation to see
564: that from (\ref{Lax3}) that the Cartan-Maurer equation $d(dX X^{-1})=
565:   (dX X^{-1})\wedge (dX X^{-1})$ implies
566: %%%%%
567: \be
568: \tau\, (dA + A\wedge A) + \fft1{\rho}\, d(\rho {*A})=0\,,
569: \ee
570: %%%
571: from which the Bianchi identity (\ref{cartmau}) and the equation of
572: motion (\ref{eom}) indeed follow.
573: 
574:    A simple calculation also shows that the integrability condition 
575: that follows by taking the exterior derivative of (\ref{dtheta}) implies
576: %%%%%
577: \be
578: d{*d}\rho=0\,.\label{boxrho}
579: \ee
580: %%%%%
581: This is indeed the correct equation of motion (\ref{rhoeom}) 
582: for $\rho$.  If we now
583: introduce light-cone coordinates for the two-dimensional spacetime,
584: and work in the gauge where the metric is flat, $ds^2=2dx^+ dx^-$, then
585: on any function $f$ we shall have
586: %%%%%
587: \be
588: df=\del_+ f \, dx^+ + \del_-f\, dx^-\,,\qquad
589: {*d}f =\del_+ f \, dx^+ - \del_-f\, dx^-\,.
590: \ee
591: %%%%%
592: Thus (\ref{boxrho}) becomes $\del_+\del_-\rho=0$, with the general
593: solution
594: %%%%%
595: \be
596: \rho = \rho_+(x^+) + \rho_-(x^-)\,.\label{rhosol}
597: \ee
598: %%%%% 
599: 
600:    The equation (\ref{dtheta}) that governs the $x^\mu$ dependence of
601: $\theta$ becomes, using the light-cone coordinates,
602: %%%%%
603: \be
604: \del_+\theta = \fft{1}{2\rho}\, (e^{2\theta} -1)\, \del_+\rho\,,\qquad
605:   \del_-\theta = \fft{1}{2\rho}\, (1-e^{-2\theta})\, \del_-\rho\,.
606: \ee
607: %%%%%
608: These equations may be integrated to give
609: %%%%%
610: \be
611: 1-e^{-2\theta} = \rho\, f_-(x^-)\,,\qquad 
612: e^{2\theta} -1 = \rho\, f_+(x^+)\,,\label{thetasol}
613: \ee
614: %%%%%
615: where $f_\pm$ are arbitrary functions of their respective arguments.  
616: Eliminating $\theta$ between the two expressions implies
617: %%%%%
618: \be
619: \rho= \fft1{f_-(x^-)} - \fft1{f_+(x^+)}\,.
620: \ee
621: %%%%%
622: Comparing with (\ref{rhosol}), we can write
623: %%%%%
624: \be
625: \fft1{f_-(x^-)} = \rho_-(x^-) + \fft1{2\lambda} \,,\qquad
626: \fft1{f_+(x^+)} = -\rho_+(x^+) + \fft1{2\lambda}\,,
627: \ee
628: %%%%%
629: where $\lambda$ is a constant.  From (\ref{rhosol}) and (\ref{thetasol}),
630: the solution for $\theta$ is given by
631: %%%%%
632: \be
633: e^{2\theta} = \fft{1+2\lambda\, \rho_-(x^-)}{1-2\lambda\, \rho_+(x^+)}\,.
634: \label{thetasol2}
635: \ee
636: %%%%%
637: It then follows from (\ref{tautheta}) that
638: %%%%%
639: \bea
640: \tau &=& \fft1{\lambda\rho}\, \Big[ 1- \lambda(\rho_+ -\rho_-) -
641:   \sqrt{(1+2\lambda\rho_-)(1-2\lambda\rho_+)}\Big] \,,\label{taulambda}\\
642: %%
643: &=& \ft12\lambda \rho + \ft12\lambda^2\rho(\rho_+-\rho_-) +
644:    \ft18\lambda^3 \rho[\rho^2 + 4(\rho_+-\rho_-)^2] +\cdots\,.\nn
645: \eea
646: %%%%%
647: Note that we have chosen the negative sign in front of the square root,
648: so that $\tau$ is analytic in $\lambda$ in the neighbourhood of
649: $\lambda=0$.  Furthermore, $\tau=0$ when $\lambda=0$.
650: 
651:  In what follows, the constant parameter $\lambda$ will play the r\^ole of the
652: spectral parameter for the Lax equation.  The solution $X$ of the Lax 
653: equation (\ref{Lax2}) depends on the spectral function $\tau$, and since 
654: $\tau$ is given by (\ref{taulambda}) it follows that we may characterise
655: $X$ by the value of the arbitrary constant spectral parameter $\lambda$.
656: Thus when we consider a solution $X$ of the Lax equation, we shall 
657: denote it by $X(\lambda)$.  
658: (Of course, $X$ also depends on the two-dimensional
659: spacetime coordinates $x^\mu$, but we shall suppress the explicit 
660: indication of this dependence.) 
661: We shall choose the solution so that $X(0)=1$;
662: this is clearly consistent with the Lax equation (\ref{Lax3}), since
663: $\tau$ vanishes at $\lambda=0$. 
664: 
665: 
666: \section{The Kac-Moody Symmetries}\label{kac-moodysec}
667: 
668: \subsection{The Kac-Moody transformations}
669: 
670:    The symmetries of the two-dimensional system are described by 
671: field transformations that preserve the equations of motion (\ref{eom})
672: and (\ref{rhoeom}).  The full set of symmetries will involve transformations
673: of the coset representative $\cV$ that are expressed using the 
674: solution $X$ of the Lax equation (\ref{Lax2}).  It is therefore necessary
675: also to find how $X$ itself transforms; this is determined by requiring in
676: addition that the Lax equation (\ref{Lax2}) be left invariant.
677: 
678:   We begin, following \cite{schwarz1,schwarz2}, by constructing transformations
679: of $\cV$.  These are of the form 
680: %%%%%
681: \be
682: \delta \cV = w \cV \eta + \delta h \cV\,,\qquad \eta\equiv 
683:    X(\lambda)\ep X(\lambda)^{-1}\,,\label{delcV0}
684: \ee
685: %%%%%
686: where $\ep$ an infinitesimal constant parameter taking values in the
687: Lie algebra $\cG$, and $\delta h$ is a $\lambda$-dependent and 
688: field-dependent compensating transformation in $\cH$ that restores the original
689: gauge.  The quantity $w$ is a function that is a singlet under
690: the Lie-algebra, and will be determined shortly. 
691: The meaning of (\ref{delcV0}) is as follows.  The
692: coset representative $\cV$ itself is, of course, independent of
693: the spectral parameter $\lambda$.  The transformation $\delta$ is
694: $\lambda$-dependent, and in fact we expand it as a power series
695: in $\lambda$:
696: %%%%%
697: \be
698: \delta(\ep,\lambda) = \sum_{n\ge0} \lambda^n\, \delta_\n(\ep)\,.
699: \ee
700: %%%%%
701: By equating coefficients of each power of $\lambda$ in the Taylor expansions
702: of the two sides of equation (\ref{delcV0}), one therefore obtains a 
703: hierarchy of transformations $\delta_\n(\ep)\cV$.  
704: One can take independent $\ep$ parameters
705: for each $n$.  The transformations at the $n=0$ order are just the original
706: infinitesimal global $\cG$ symmetries.
707: 
708:    In the flat-space case discussed in 
709: \cite{schwarz1,lupepo} $w$ is spacetime-independent, and one can take $w=1$, 
710: but in the gravity-coupled situation we are considering in this paper,
711: one must choose a very specific $x^\mu$-dependence for $w$ in order
712: to ensure that (\ref{delcV0}) describes a symmetry of the equation
713: of motion $d(\rho{*A})=0$.  The calculation is performed by first noting
714: from (\ref{MAdef}) that (\ref{delcV0}) implies
715: %%%%%
716: \be
717: \delta M = w M\eta + w \eta^\dagger M\,,\qquad 
718: \delta A= D(w\eta) + D(w M^{-1} \eta^\dagger M)\,,\label{delMA}
719: \ee
720: %%%%%
721: where we define $Df\equiv df + [A,f]$ on any $\cG$-valued function $f$.
722: Note that $\rho$ will be inert under the Kac-Moody transformations.
723: 
724:     Use of the Lax equation (\ref{Lax2}) shows that
725: %%%%%
726: \be
727: D\eta= \fft1{\tau}\, {*d\eta}\,,\qquad
728:  D(M^{-1}\eta^\dagger M)= \tau\, {*d}(M^{-1}\eta^\dagger M)\,,
729: \ee
730: %%%%%
731: as in the flat-space case \cite{lupepo}. From this, it follows that
732: %%%%%
733: \bea
734: d(\rho {*\delta A}) &=& \Big[d\Big(\fft{\rho w}{\tau}\Big)-
735:      \rho {*d}w \Big]\wedge d\eta + 
736:  \big[d(\rho w \tau)- \rho {*d}w\big]\wedge d(M^{-1}\eta^\dagger M)\nn\\
737: %%
738: && +
739: d(\rho{*d}w)\, \big(\eta + M^{-1}\eta^\dagger M\big)\,.
740: \eea
741: %%%%
742: and so $d(\rho {*\delta A})$ will vanish (and thus the equation of
743: motion (\ref{eom}) will be preserved by the transformations) if 
744: \cite{schwarz2}
745: %%%%%
746: \be
747: \rho {*dw} = d\Big(\fft{\rho w}{\tau}\Big)\,,\qquad 
748:    \hbox{and}\qquad \rho{*dw} = d(\rho w \tau)\,.
749: \ee
750: %%%%
751: Using (\ref{dtheta}) one sees that this can be achieved, by taking $w$
752: to be a constant multiple of $s/\rho$.  
753: We shall take $w=s/(\lambda\rho)$, and thus the 
754: transformations (\ref{delcV0}) we shall consider are given by
755: %%%%%
756: \be
757: \delta \cV = \fft{s}{\lambda\rho}\, \cV\eta + \delta h \cV\,,\qquad
758: \eta\equiv X(\lambda)\ep X(\lambda)^{-1}\,.\label{delcV}
759: \ee
760: %%%%%
761: (The reason for choosing to divide out by the constant parameter $\lambda$
762: is that then, as can be seen from (\ref{scdef}) and (\ref{taulambda}), 
763: the prefactor $w=s/(\lambda\rho)$ approaches 1 as $\lambda$ goes to zero.)
764: 
765:    We now need to consider the transformations of $X$.  These are determined
766: by the requirement that the Lax equation must be preserved, with $A$
767: transforming as in (\ref{delMA}).  We shall need to know how $X(\lambda)$ with
768: spectral parameter $\lambda_2$ transforms under variations 
769: $\delta(\ep,\lambda)$ with an independent choice of spectral parameter
770: $\lambda_1$.  We denote $X(\lambda_2)$ by $X_2$, and $\delta(\ep_1,\lambda_1)$
771: by $\delta_1$.  From (\ref{Lax3}) we therefore have
772: %%%%%
773: \be
774: \delta_1(dX_2 X_2^{-1}) = \fft{\tau_2}{1-\tau_2^2}\, {*\delta_1 A} + 
775:          \fft{\tau_2^2}{1-\tau_2^2}\, \delta_1 A\,.\label{varlax}
776: \ee
777: %%%%%
778: ($\tau_2$ means the solution for $\tau$ given in (\ref{taulambda}) with
779: the spectral parameter $\lambda$ taken to be $\lambda_2$.) 
780: 
781:    Let us write the transformation of $X$ as $\delta_1 X_2 = U X_2$.  It 
782: follows that $\delta_1(dX_2 X_2^{-1})=dU + [U, dX_2 X_2^{-1}]$, and so 
783: from (\ref{varlax}) we have
784: %%%%%
785: \be
786: dU + [U,dX_2 X_2^{-1}] = \fft{\tau_2}{1-\tau_2^2}\, {*\delta_1 A} +
787:          \fft{\tau_2^2}{1-\tau_2^2}\, \delta_1 A\,.\label{varlax2}
788: \ee
789: %%%%%
790: We can write the solution for $U$ as the sum 
791: %%%%%
792: \be
793:   U= U^{\rm{hom}} + U^{\rm{inhom}}\,,
794: \ee
795: %%%%%
796: where $U^{\rm{hom}}$ is the solution of the homogeneous equation with
797: the right-hand side of (\ref{varlax2}) set to zero, while $ U^{\rm{inhom}}$
798: solves the inhomogeneous equation with the variations of $A$ on the
799: right-hand side providing the source.
800: 
801:    It is easily seen that the homogeneous equation is solved by
802: %%%%%
803: \be
804: U^{\rm{hom}} = X_2\ep_1 X_2^{-1}\,.\label{homsol}
805: \ee
806: %%%%%
807: For the inhomogeneous equation, we make the ansatz
808: %%%%%
809: \be
810: U^{\rm{inhom}} = u\, \eta_1  + v\, M^{-1}\eta_1^\dagger M \,,
811: \ee
812: %%%%%
813: where $u$ and $v$ are $\cG$-singlet functions to be determined.  Note that
814: $\eta_1\equiv X_1 \ep_1 X_1^{-1}$, with $X_1\equiv X(\lambda_1)$, 
815: and that therefore
816: $d\eta_1 = -[\eta_1, dX_1 X_1^{-1}]$.  Substituting the variations of
817: $A$, and the ansatz for $U=U^{\rm{inhom}}$, into (\ref{varlax2}) one
818: finds, after some straightforward although slightly elaborate algebra
819: involving the use of the Lax equation (\ref{Lax3}), (\ref{dtau}) and 
820: (\ref{dtheta}), that there is a solution with \cite{schwarz2}
821: %%%%%
822: \be
823: u= \fft{s_1}{\lambda_1 \rho}\, \fft{\tau_2}{(\tau_1-\tau_2)}\,,\qquad
824:  v= \fft{s_1}{\lambda_1\rho}\, \fft{\tau_1\tau_2}{(1-\tau_1\tau_2)}\,.
825: \label{uvsol}
826: \ee
827: %%%%%
828: Here $s_1$ denotes $s=\sinh\theta=2\tau/(1-\tau^2)$ with $\tau$ given by
829: (\ref{taulambda}) for $\lambda=\lambda_1$.
830: A remarkable property, which is absolutely crucial in what follows, is
831: that $d(u-v)=0$ and hence
832: %%%%%
833: \be
834: u-v=\hbox{constant}\,.\label{uvconst}
835: \ee
836: %%%%%
837: This can be verified using (\ref{dtau}) and (\ref{dtheta}).
838: 
839:    To summarise the results so far, we have shown that there exist symmetry
840: transformations of $X$ of the form
841: %%%%%
842: \bea
843: \delta_1^{\rm{hom}} X_2 &=& X_2 \ep_1\,,\label{deltahom}\\
844: %%
845: \delta_1^{\rm{inhom}} X_2 &=& \fft{s_1}{\lambda_1 \rho}\, 
846:    \fft{\tau_2}{(\tau_1-\tau_2)}\, \eta_1 X_2 + 
847:     \fft{s_1}{\lambda_1\rho}\, \fft{\tau_1\tau_2}{(1-\tau_1\tau_2)}\,
848:    M^{-1}\eta_1^\dagger M X_2\,.\label{deltainhom}
849: \eea
850: %%%%%
851: The inhomogeneous transformations are accompanied by the transformations
852: of $\cV$, $M$ and $A$ given by (\ref{delcV}) and (\ref{delMA}), 
853: whilst the homogeneous transformations
854: are completely independent, acting only on $X$ and leaving $\cV$, $M$
855: and $A$ invariant. 
856: 
857:    The complete set of global symmetry transformations of $X$ will be read off
858: by expanding the various $\delta_1$ variations, and $X_2$, as power series
859: in $\lambda_1$ and $\lambda_2$ respectively, both around $\lambda_i=0$.
860: Since $\tau_i$ vanishes at $\lambda_i=0$ (see (\ref{taulambda})) 
861: this appears to present a difficulty for
862: the inhomogeneous transformations (\ref{deltainhom}), because of the pole
863: associated with the 
864: denominator $(\tau_1-\tau_2)$ in the first term.  In the analogous
865: treatment of the flat-space case (see \cite{schwarz1,lupepo}) this was
866: not a problem, because there $\tau_1$ and $\tau_2$ were themselves constant 
867: spectral parameters and so the pole could subtracted by including an
868: appropriate constant multiple of the homogeneous transformation.  Here,
869: we cannot simply subtract the analogous multiple of $\delta^{\rm{hom}}$ 
870: with a prefactor of the form $(s_1/\lambda_1\rho)\, 
871:   \tau_2(\tau_1-\tau_2)^{-1}$, 
872: because this is not a constant, and so $(s_1/\lambda_1\rho)\, 
873: \tau_2(\tau_1-\tau_2)^{-1} 
874: \delta_1^{\rm{hom}}X_2$ is not a symmetry transformation.  At this
875: point the remarkable property (\ref{uvconst}) comes to the rescue.  In fact,
876: by substituting (\ref{tautheta}) and (\ref{thetasol2}) into (\ref{uvsol})
877: we find that
878: %%%%%
879: \be
880: \fft{s_1}{\lambda_1 \rho}\, \fft{\tau_2}{(\tau_1-\tau_2)} =
881: \fft{s_1}{\lambda_1\rho}\, \fft{\tau_1\tau_2}{(1-\tau_1\tau_2)}
882:   + \fft{\lambda_2}{\lambda_1-\lambda_2}\,.\label{uvrel2}
883: \ee
884: %%%%%
885: Thus, we may rewrite (\ref{deltainhom}) as
886: %%%%%
887: \be
888: \delta_1^{\rm{inhom}} X_2 = \fft{s_1}{\lambda_1\rho}\, 
889:   \fft{\tau_1\tau_2}{(1-\tau_1\tau_2)}\, \big(
890:  \eta_1 + M^{-1}\eta_1^\dagger M\big) X_2 + 
891:   \fft{\lambda_2}{\lambda_1-\lambda_2}\, \eta_1 X_2\,.\label{deltainhom2}
892: \ee
893: %%%%%
894: This shows that, despite the original appearance in (\ref{deltainhom}), 
895: the pole at 
896: $\lambda_1=\lambda_2$ in the inhomogeneous transformation rule actually
897: has a pure constant coefficient, and so by
898: subtracting the homogeneous symmetry transformation 
899: $\lambda_2/(\lambda_1-\lambda_2)\, \delta_1^{\rm{hom}}$ we can easily
900: remove the pole.\footnote{In Schwarz's discussion in \cite{schwarz2}, the 
901: important point that one must remove the $\tau_1=\tau_2$ singularity in 
902: $\delta_1^{\rm{inhom}}X_2$, and furthermore that this can actually be done,
903: for all spacetime points $x^\mu$ simultaneously, appears to have been 
904: unnoticed.  In fact Schwarz instead made a subtraction such that 
905: $\delta_1^{\rm{inhom}}X_2$ vanished at a preferred point $x_0^\mu$ in 
906: the two-dimensional spacetime.  However, this subtraction does not in fact
907: cancel the singularity at $\tau_1=\tau_2$ for other points in the 
908: spacetime.  The absence of the cancellation did not show up in Schwarz's
909: subsequent calculation of the commutator $[\delta_1,\delta_2]$ because
910: he evaluated it only on $\cV$ (which is inert under the homogeneous
911: variation in question)
912: and not on $X_3$ (which is not inert).}
913: 
914:    With these points understood, we can now present the full set of global
915: symmetry transformations of the two-dimensional system in the final
916: form that we shall use in what follows.  We shall denote them by
917: $\delta_1$ and $\td\delta_1$, and their actions on the original
918: sigma model fields in $\cV$ (and hence on $M$), 
919: and on the fields in $X(\lambda)$, are as follows:
920: %%%%%
921: \bea
922: \delta_1 \cV &=& \fft{s_1}{\lambda_1\rho}\, \cV\eta_1 + \delta h \cV\,,
923: \label{deltacV}\\
924: %%
925: \delta_1 M &=&  \fft{s_1}{\lambda_1\rho}\, 
926:   \big(M\eta_1 + \eta_1^\dagger M\big)\,,\label{deltaM}\\
927: %% 
928: \delta_1 X_2 &=& \fft{\lambda_2}{\lambda_1-\lambda_2}\, 
929:   \big(\eta_1 X_2 - X_2 \ep_1\big) + \fft{s_1}{\lambda_1\rho}\,
930:    \fft{\tau_1\tau_2}{1-\tau_1\tau_2}\, 
931:             \big(\eta_1 + M^{-1}\eta_1^\dagger M\big) X_2\,,\label{deltaX}
932: \eea
933: %%%%%
934: where $\eta_1\equiv X_1\ep_1 X_1^{-1}$, and 
935: %%%%%
936: \bea
937: \td\delta_1 \cV &=&0\,,\label{tddeltacV}\\
938: %%
939: \td\delta_1 M &=&0\,,\label{tddeltaM}\\
940: %%
941: \td\delta_1 X_2 &=& \fft{\lambda_1\lambda_2}{1-\lambda_1\lambda_2}\, 
942:        X_2\ep_1\,.\label{tddeltaX}
943: \eea
944: %%%%%    
945: The $\delta$ transformations are the inhomogeneous transformations we
946: discussed above, with the subtraction of the necessary homogeneous term 
947: in $\delta_1 X_2$ to ensure analyticity when $\lambda_1$ approaches 
948: $\lambda_2$.  The $\td\delta$ transformations are 
949: independent and purely homogeneous, thus leaving $\cV$ and $M$ invariant.  The
950: inclusion of the specific $\lambda_i$-dependent prefactor in
951: (\ref{tddeltaX}) is purely for convenience; it ensures that the final
952: algebra obtained by calculating the commutators of the transformations
953: arises in a simple and conventional basis.  
954: 
955:    It is evident from the
956: transformations above that the expansions of the variations $\delta$ and
957: $\td\delta$ will be of the forms
958: %%%%%
959: \be
960: \delta(\ep,\lambda) = \sum_{n\ge0} \lambda^n\, \delta_\n(\ep)\,,\qquad
961: \td\delta(\ep,\lambda) = \sum_{n\ge1} \lambda^n\, \td\delta_\n(\ep)\,.
962: \label{deltaexps}
963: \ee
964: %%%%% 
965: There is an independent $\cG$-valued infinitesimal parameter for each 
966: $n$ in $\delta_\n(\ep)$, and for each $n$ in $\td\delta_\n(\ep)$.
967: 
968: \subsection{The Kac-Moody algebra}
969: 
970:    Having obtained the explicit expressions (\ref{deltacV})--(\ref{tddeltaX})
971: for the $\delta$ and $\td\delta$ transformations of the fields, it is
972: now a mechanical exercise to calculate the commutators of these 
973: transformations.  Specifically, we calculate the commutators 
974: $[\delta_1,\delta_2]$, $[\delta_1,\td\delta_2]$ and 
975: $[\td\delta_1,\td\delta_2]$ acting on $M$ and on $X_3$.  After some algebra,
976: we find
977: %%%%%
978: \bea
979: {[}\delta_1,\delta_2{]} &=&
980:  \fft{\lambda_1}{\lambda_1-\lambda_2}\,
981:       \delta(\ep_{12}, \lambda_1)- \fft{\lambda_2}{\lambda_1-\lambda_2}\,
982:    \delta(\ep_{12},\lambda_2)\,,
983: \label{nn}\\
984: %%
985: {[}\delta_1,\td\delta_2{]} &=& 
986:   \fft{\lambda_1 \lambda_2}{1-\lambda_1 \lambda_2}\,
987:                \delta(\ep_{12},\lambda_1) + \fft1{1-\lambda_1 \lambda_2}\,
988:           \td\delta(\ep_{12},\lambda_2)\,,\label{ntd}\\
989: %%
990: {[}\td\delta_1,\td\delta_2{]} &=& \fft{\lambda_2}{\lambda_1-\lambda_2}\,
991:   \td\delta(\ep_{12},\lambda_1)
992:   -\fft{\lambda_1}{\lambda_1-\lambda_2}\, 
993:          \td\delta(\ep_{12}, \lambda_2)\,,\label{tdtd}
994: \eea
995: %%%%%
996: where $\ep_{12}\equiv [\ep_1,\ep_2]$.  Note that there are no poles
997: at $\lambda_1=\lambda_2$, because in (\ref{nn}) and (\ref{tdtd}) the numerator
998: on the right-hand side has a zero that cancels the denominator there.
999: 
1000:    Using (\ref{deltaexps}), expanding the expressions 
1001: (\ref{nn})--(\ref{tdtd}) in powers of $\lambda_1$ and $\lambda_2$, and
1002: then equating the coefficients of each power, we can read off the algebra
1003: of the modes, finding
1004: %%%%%
1005: \bea
1006: {[}\delta_\m(\ep_1),\delta_\n(\ep_2){]} &=& \delta_{\sst{(m+n)}}(\ep_{12})\,,
1007: \qquad m\ge0\,,\ n\ge0\,,\label{nn2}\\
1008: %%
1009: {[}\delta_\m(\ep_1),\td\delta_\n(\ep_2){]} &=& \delta_{\sst{(m-n)}}(\ep_{12})
1010:         + \td\delta_{\sst{(n-m)}}(\ep_{12})\,,\qquad m\ge0\,,\ n\ge1\,,
1011:    \label{ntd2}\\
1012: %%
1013:  {[}\td\delta_\m(\ep_1),\td\delta_\n(\ep_2){]} &=&
1014:        \td\delta_{\sst{(m+n)}}(\ep_{12})\,,
1015:   \qquad m\ge1\,,\ n\ge1\,,\label{tdtd2}
1016: \eea
1017: %%%%%
1018: where in (\ref{ntd2}) it is to be understood that $\delta_\n=0$ for
1019: $n\le -1$ and $\td\delta_\n=0$ for $n\le0$.
1020: 
1021:    As in the flat-space case discussed in \cite{lupepo}, these three sets
1022: of commutation relations can be combined into one
1023: by introducing a new set $\Delta_\n$ of variations, defined for all $n$
1024: with $-\infty\le n\le \infty$, according to
1025: %%%%%
1026: \bea
1027:    \Delta_\n &=& \delta_\n\,,\qquad n\ge0\,,\nn\\
1028: %%
1029:  \Delta_{\sst{(-n)}} &=& \td\delta_\n\,,\qquad n\ge 1\,.\label{Deltadef}
1030: \eea
1031: %%%%%
1032: It is then easily seen that (\ref{nn2}), (\ref{ntd2}) and (\ref{tdtd2})
1033: become
1034: %%%%%
1035: \be
1036: [\Delta_\m(\ep_1),\Delta_\n(\ep_2)] = \Delta_{\sst{(m+n)}}(\ep_{12})\,,
1037: \qquad m,n\in\Z \,,\label{kacmoody}
1038: \ee
1039: %%%%%
1040: with $\ep_{12}=[\ep_1,\ep_2]$.
1041: This defines the affine Kac-Moody algebra $\hat \cG$.  If we associate
1042: generators $J_n^i$ with the transformations $\Delta_\n(\ep^i)$, where
1043: $\ep= \ep^i\, T^i$ and $T^i$ are the generators of the Lie algebra $\cG$
1044: satisfying $[T^i,T^j]=f^{ij}{}_k\, T^k$, then (\ref{kacmoody}) implies
1045: %%%%%
1046: \be
1047: [J_m^i,J_n^j]= f^{ij}{}_k\, J^k_{m+n}\,.\label{kacmoody2}
1048: \ee
1049: %%%%% 
1050: 
1051: \section{Virasoro-Like Symmetries}\label{virasorosec}
1052: 
1053: \subsection{Virasoro-like transformations}\label{virsubsec}
1054: 
1055:      In addition to the Kac-Moody symmetries that we discussed in section
1056: \ref{kac-moodysec}, which are the affine extension of the original
1057: $\cG$ global symmetry of the sigma model, there is also a further
1058: infinite-dimensional symmetry that is a singlet under $\cG$ 
1059: \cite{houli,maison,nicolai}.   This is related to the Virasoro algebra. 
1060: It was discussed in detail for sigma models in flat two-dimensional
1061: spacetime in \cite{schwarz1}, but the generalisation to the gravity-coupled
1062: case was not found in \cite{schwarz2}.  Here, we use the methods developed
1063: in \cite{schwarz1,schwarz2}, and extend them to obtain explicit Virasoro-like
1064: transformations in the gravity-coupled sigma models.
1065: 
1066:    The action of the Virasoro-like transformations on the coset representative
1067: $\cV$ will be taken to be\footnote{We should really include an infinitesimal 
1068: parameter as a prefactor in the definition of $\xi$ in
1069: equation (\ref{deltaVcV}).  However, since it is a singlet
1070: it plays no significant r\^ole, and so it may be omitted without any
1071: risk of ambiguity.} 
1072: %%%%%
1073: \be
1074: \delta^V \cV = h\cV \xi\,,\qquad \xi \equiv \dot X X^{-1}\,,
1075: \label{deltaVcV}
1076: \ee
1077: %%%%%
1078: where $h$ is a $\cG$-singlet function to be determined, and $\dot X$ means
1079: $dX(\lambda)/d\lambda$.  From this it follows that
1080: %%%%%
1081: \be
1082: \delta^V A = D(h(\xi + M^{-1} \xi^\dagger M))\,.\label{deltaVA}
1083: \ee
1084: %%%%%
1085: We can show from (\ref{thetasol2}) that
1086: %%%%%
1087: \be
1088: \dot\theta = \fft{s^2}{\lambda^2 \rho}\,,\qquad
1089: \dot\tau= \fft{s\tau}{\lambda^2\rho}\,.
1090: \ee
1091: %%%%%
1092: By taking the $\lambda$ derivative of the Lax equation (\ref{Lax3}),
1093: we can then show after a little algebra that
1094: %%%%%
1095: \be
1096: D\xi= \fft1{\tau}\, {*d}\xi -\fft{s^2}{2\lambda^2\rho\tau}\, (\tau {*A} +A)\,,
1097: \qquad 
1098: D(M^{-1}\xi^\dagger M)= \tau {*d}(M^{-1} \xi^\dagger M) +
1099:   \fft{s^2}{2\lambda^2\rho}\, ({*A}+\tau A)\,,
1100: \ee
1101: %%%%%
1102: and hence that
1103: %%%%%
1104: \be
1105: D(\xi + M^{-1}\xi^\dagger M)=\fft1{\tau}\, {*d}\xi + 
1106:   \tau {*d}(M^{-1} \xi^\dagger M) -\fft{s}{\lambda^2\rho}\, A\,.\label{virid}
1107: \ee
1108: %%%%%
1109: 
1110:    We now examine the variation of the equation of motion for $A$,
1111: namely 
1112: %%%%%
1113: \be
1114: \delta^V\big(d(\rho {*A})\big)= 
1115: d\big( (\delta^V\rho){*A} + \rho {*\delta^V A}\big) =0\,.\label{eomvar}
1116: \ee
1117: %%%%%
1118: Note that unlike the Kac-Moody transformations, which leave $\rho$ invariant,
1119: here we shall find that the Virasoro-like transformations must act also
1120: on $\rho$.   Substituting (\ref{deltaVA}) into (\ref{eomvar}), and
1121: using (\ref{virid}), we obtain
1122: %%%%
1123: \bea
1124: \delta^V\big(d(\rho {*A})\big) &=&
1125: \Big[d\Big(\fft{\rho h}{\tau}\Big) - \rho {*dh}\Big]\wedge d\xi +
1126:  [d(\rho h\tau) - \rho{*dh}]\wedge d(M^{-1}\xi^\dagger M)\nn\\
1127: %%
1128: && +d(\rho {*dh})[\xi + M^{-1}\xi^\dagger M] -d\Big(\fft{h s}{\lambda^2}\,
1129:   {*A}\Big) + d(\delta^V\rho\, {*A})\,.\label{var2}
1130: \eea
1131: %%%%%
1132: This leads us to require $h$ to satisfy
1133: %%%%%
1134: \be
1135: d\Big(\fft{\rho h}{\tau}\Big) =d(\rho h\tau)=\rho{*dh}\,,
1136: \ee
1137: %%%%%
1138: which has as solution
1139: %%%%%
1140: \be
1141:    h=-\fft{s}{\rho}\,,\label{hsol2}
1142: \ee
1143: %%%%
1144: (the constant factor is arbitrary, and we take it to be $-1$ for
1145: later convenience).  Equation 
1146: (\ref{var2}) now reduces to 
1147: %%%%%
1148: \be
1149: \delta^V\big(d(\rho {*A})\big)= d\Big(\fft{s^2}{\lambda^2\rho}\,
1150:   {*A}\Big) + d(\delta^V\rho\, {*A})  \,,
1151: \ee
1152: %%%%%
1153: which, in view of the fact that $d(\rho {*A})=0$, vanishes if we take
1154: %%%%%
1155: \be
1156: \delta^V\rho = -\fft{s^2}{\lambda^2\rho} +\alpha \rho\,,\label{deltarho}
1157: \ee
1158: %%%%%
1159: where $\alpha$ is an arbitrary constant parameter.  However, since
1160: by its definition $\delta^V$ acting on $\cV$ has no $\lambda^0$ term
1161: in its Taylor expansion, we may choose $\alpha$ so as to remove the
1162: $\lambda^0$ term in (\ref{deltarho}).  (We shall return to discussing 
1163: the extra symmetry associated with $\alpha$ later.)  This means setting
1164: %%%%%
1165: \be
1166: \alpha=1\,.
1167: \ee
1168: %%%%%
1169: 
1170:    Since $\rho$ satisfies the free wave equation (\ref{rhoeq}), we must
1171: also check that its variation (\ref{deltarho}) is consistent with this.
1172: This is easily done, by substituting (\ref{thetasol2}) into (\ref{deltarho}),
1173: giving
1174: %%%%%
1175: \be
1176: \delta^V \rho = -\fft{2\lambda\rho_+^2}{1-2\lambda \rho_+} +
1177:    \fft{2\lambda \rho_-^2}{1+2\lambda \rho_-} \,.\label{deltarho2}
1178: \ee
1179: %%%%%
1180: Since this is manifestly the sum of a function of $\rho_+(x^+)$ and a
1181: function of $\rho_-(x^-)$, it clearly satisfies 
1182: $\del_+\del_-(\delta^V\rho)=0$.  In fact we can read off from 
1183: (\ref{deltarho2}) that
1184: %%%%%
1185: \be
1186: \delta^V\rho_+ = -\fft{2\lambda\rho_+^2}{1-2\lambda\rho_+} + \beta\,,\qquad
1187: \delta^V\rho_- = \fft{2\lambda\rho_-^2}{1+2\lambda\rho_-} -\beta\,,
1188: \label{deltarhoprhom}
1189: \ee
1190: %%%%%
1191: where $\beta$ is another arbitrary constant parameter.  Again, like
1192: the previous discussion which implied we could take $\alpha=1$, here
1193:  too we could require that there be no $\lambda^0$ term in the expansion, 
1194: and this constrains $\beta$ also, to 
1195: %%%%%
1196: \be
1197: \beta=0\,.
1198: \ee
1199: %%%%%
1200: We shall discuss the extra symmetry associated with a non-zero $\beta$
1201: below.
1202: 
1203:    It is useful also to record the expressions for $\delta_1^V\theta_2$ and
1204: $\delta_1^V\tau_2$, for which we find
1205: %%%%%
1206: \bea
1207: \delta_1^V \theta_2 &=& -\fft{2 \tau_1\tau_2}{\rho^2 
1208:       (\t_1-\tau_2)(1-\tau_1 \tau_2)}\, \Big(\fft{s_1^2}{\lambda_1^2}
1209:                -\fft{s_2^2}{\lambda_2^2}\Big)\,,\label{delVtheta}\\
1210: %%
1211: \delta_1^V \tau_2 &=& - \fft{\tau_1\tau_2(1-\tau_2^2)}{\rho^2
1212:       (\t_1-\tau_2)(1-\tau_1 \tau_2)}\, \Big(\fft{s_1^2}{\lambda_1^2}
1213:                -\fft{s_2^2}{\lambda_2^2}\Big)\,,\label{delVtau}
1214: \eea
1215: %%%%%
1216: 
1217:    The next step is to calculate $\delta_1^V X_2$.  This is done by
1218: requiring the invariance of the Lax equation (\ref{Lax3}) 
1219: (for $\lambda=\lambda_2$) under the transformation $\delta_1^V$,
1220: with $\delta_1^V A$ and $\delta_1^V\tau_2$ read off from (\ref{deltaVA})
1221: and (\ref{delVtau}).  After some algebra, we find
1222: %%%%%
1223: \be
1224: \delta_1^V X_2 = -\fft{s_1}{\rho} \fft{\tau_1\tau_2}{(1-\tau_1\tau_2)}\,
1225: (\xi_1 + M^{-1} \xi_1^\dagger M) X_2 - 
1226:   \fft{\lambda_1\lambda_2}{\lambda_1-\lambda_2}\, (\xi_1-\xi_2) X_2\,.
1227: \label{deltaVX}
1228: \ee
1229: %%%%%
1230: 
1231: \subsection{Virasoro-like symmetries}
1232: 
1233:    We are now in a position to calculate the commutator
1234: $[\delta_1^V,\delta_2^V]$.  We have evaluated this on all the
1235: fields, namely $\cV$, $\rho$ and $X$.   We find that the abstract algebra
1236: turns out to be
1237: %%%%%
1238: \be
1239: [\delta_1^V,\delta_2^V] = 
1240: \fft{\lambda_1\lambda_2}{\lambda_1-\lambda_2}\,
1241: (\dot\delta_1^V + \dot\delta_2^V)
1242: -\fft{2\lambda_1\lambda_2}{(\lambda_1-\lambda_2)^2}\,
1243: (\delta_1^V -\delta_2^V)
1244: \,.\label{valg1}
1245: \ee
1246: %%%%%
1247: 
1248:   Making the mode expansion
1249: %%%%%
1250: \be
1251: \delta^V(\lambda) = \sum_{n\ge1} \lambda^n \delta^V_\n\,,
1252: \label{virmodes}
1253: \ee
1254: %%%%%
1255: we find upon substituting into (\ref{valg1}) that
1256: the modes satisfy the commutation relations
1257: %%%%%
1258: \be
1259: [\delta^V_\m,\delta^V_\n] = (m-n) \delta^V_{\sst{(m+n)}}\,,\qquad m\ge1\,\ \ 
1260:    n\ge 1\,.\label{virexp1}
1261: \ee
1262: %%%%%
1263: As in the flat-space case we have obtained ``half'' the Virasoro algebra.
1264: It is interesting that here we obtain precisely the upper half of a standard
1265: Virasoro algebra of the $L_n$, whereas in the flat-space case we found 
1266: generators $K_n=L_n-L_{-n}$.  The difference between the two results from
1267: the use of a different spectral parameterisation.  (See appendix \ref{appsec} 
1268: for a
1269: discussion of how the flat-space limit of our gravity-coupled construction
1270: here is related to the previous flat-space construction in \cite{lupepo}.)
1271: 
1272:    We now return to the additional symmetries associated with the 
1273: parameter $\alpha$ in (\ref{deltarho}) and $\beta$ in (\ref{deltarhoprhom}). 
1274: It will be recalled that we ``fixed'' these symmetries by requiring the
1275: absence of $\lambda^0$ terms in $\delta^V\rho$ and in $\delta^V\rho_\pm$,
1276: on the grounds that the original Virasoro-like transformations $\delta^V\cV$
1277: had contributions only for strictly positive powers of $\lambda$.  However,
1278: we can interpret the transformations parameterised by $\alpha$
1279: and $\beta$ as additional symmetries of the system that act non-trivially
1280: on $\rho$, but which happens to leave $\cV$ inert.  In fact it is
1281: natural to denote the $\alpha$ symmetry by $\hat\delta^V_{\sst{(0)}}$ and
1282: the $\beta$ symmetry by $\hat\delta^V_{\sst{(-1)}}$, since we find that they
1283: are the natural continuation of the Virasoro transformations $\delta^V_\n$
1284: in (\ref{virexp1}) to include the $n=0$ and $n=-1$ terms.  
1285: 
1286:    In detail, we find that the action of the $\hat\delta^V_{\sst{(0)}}$ and
1287: $\hat\delta^V_{\sst{(-1)}}$ transformations are as follows:
1288: %%%%%
1289: \bea
1290: \hat\delta^V_{\sst{(0)}}\rho_+ &=&  -\rho_+\,,\qquad 
1291:       \hat\delta^V_{\sst{(0)}} \rho_- = -\rho_-\,,\qquad 
1292: \hat\delta^V_{\sst{(0)}} X = -\lambda X(\lambda)\,, 
1293: \qquad \hat\delta^V_{\sst{(0)}} \cV =0\,, \nn\\
1294: %%
1295: \hat\delta^V_{\sst{(-1)}}\rho_+ &=&  -\ft12 \,,\qquad
1296:       \delta^V_{\sst{(-1)}} \rho_- = \ft12 \,,\qquad
1297: \hat\delta^V_{\sst{(-1)}} X = -\lambda^2 X(\lambda)\,,
1298: \qquad \hat\delta^V_{\sst{(-1)}} \cV =0\,,\label{virextra}
1299: \eea
1300: %%%%%
1301: It is also useful to record that
1302: %%%%%
1303: \be
1304: \hat\delta^V_{\sst{(0)}}\rho= -\rho\,,\quad \hat\delta^V_{\sst{(0)}}\tau=
1305:    -\lambda \dot\tau\,;\qquad
1306: \hat\delta^V_{\sst{(-1)}}\rho= 0\,,\quad \hat\delta^V_{\sst{(-1)}}\tau=
1307:    -\lambda^2 \dot\tau\,.
1308: \ee
1309: %%%%%
1310: 
1311:    It is a straightforward matter to calculate the commutators of 
1312: $\hat\delta^V_{\sst{(0)}}$ and $\hat\delta^V_{\sst{(-1)}}$ with each other, and
1313: with the previously-defined Virasoro-type transformations appearing in
1314: (\ref{virmodes}).  We first define a total extended set of transformations 
1315: $\hat\delta^V_\n$ with $n\ge -1$, according to
1316: %%%%%
1317: \be
1318: \hat\delta^V_\n = \delta^V_\n\,,\qquad n\ge 1\,,
1319: \ee
1320: %%%%%
1321: where $\delta^V_\n$ is defined in (\ref{virmodes}), and with 
1322: $\hat\delta^V_\n$ for $n=0$ and $n=-1$ as in (\ref{virextra}).  We find
1323: that these satisfy precisely the same algebra as in (\ref{virexp1}),
1324: now with the index range extended appropriately, {\it viz.}
1325: %%%%%
1326: \be
1327: [\hat\delta^V_\m,\hat\delta^V_\n] = (m-n) 
1328:    \hat\delta^V_{\sst{(m+n)}}\,,\qquad m\ge-1\,\ \
1329:    n\ge -1\,.\label{virexp2}
1330: \ee
1331: %%%%%
1332: Thus we may associate standard Virasoro generators $L_n$ with 
1333: $\hat\delta^V_\n$,
1334: and we obtain the Virasoro subalgebra generated by $L_n$ with $n\ge -1$:
1335: %%%%%
1336: \be
1337: [L_m,L_n]=(m-n) L_{m+n}\,,\qquad m\ge 1\,,\ n\ge1\,.
1338: \ee
1339: %%%%% 
1340: 
1341:    It is natural to introduce an extended Virasoro-like transformation
1342: $\hat\delta^V$ constructed from the extended mode set 
1343: $\hat\delta^V_\n$, by analogy
1344: with (\ref{virmodes}).  Thus we may define
1345: %%%%%
1346: \be
1347: \hat\delta^V(\lambda) = \sum_{n\ge -1} \lambda^n \hat\delta^V_\n =
1348:    \delta^V(\lambda) + \hat\delta^V_{\sst{(0)}} + \fft{1}{\lambda}\,
1349:       \hat\delta^V_{\sst{(-1)}}\,.
1350: \ee
1351: %%%%%
1352: Acting, for example, on $\rho_+$, we find from (\ref{deltarhoprhom}) 
1353: (with $\beta=0$) and (\ref{virextra}) that
1354: %%%%%
1355: \be
1356: \hat\delta^V \rho_+= -\fft{2\lambda\rho_+}{1-2\lambda\rho_+} -
1357:   \rho_+ - \fft1{2\lambda} = -\fft{1}{2\lambda(1-2\lambda\rho_+)}\,.
1358: \ee
1359: %%%%%
1360: The $\hat\delta^V$ variations of $\rho_-$, $\rho$ and $X$ can all easily
1361: be worked out in a similar way from our previous results.  ($\cV$, as
1362: we noted already, is inert under the extra terms $ \hat\delta^V_{\sst{(0)}}$
1363: and $\hat\delta^V_{\sst{(-1)}}$.)  In summary, the extended transformations
1364: $\hat\delta^V$ act on all the fields as follows:
1365: %%%%%
1366: \bea
1367: \hat\delta^V \rho_+ &=& -\fft1{2\lambda(1-2\lambda\rho_+)}\,,\qquad
1368:  \hat\delta^V \rho_- = \fft1{2\lambda(1+2\lambda\rho_-)}\,,\qquad
1369: \hat\delta^V \rho = -\fft{s^2}{\lambda^2\rho}\,,\nn\\
1370: %%
1371: \hat\delta_1^V X_2&=& -\fft{s_1}{\rho}\, \fft{\tau_1\tau_2}{(1-\tau_1\tau_2)}\,
1372:  \big(\xi_1 + M^{-1}\xi_1^\dagger M\big) X_2 - 
1373: \fft{\lambda_2 /\lambda_1}{(\lambda_1-\lambda_2)}\, 
1374:   \big(\lambda_1^2\, \xi_1 -\lambda_2^2 \, \xi_2\big)\,,\nn\\
1375: %%
1376: \hat\delta^V \cV &=& -\fft{s}{\rho}\, \cV \xi\,.\label{finalV}
1377: \eea
1378: %%%%%
1379: 
1380: 
1381: 
1382: \subsection{Commutators of Virasoro and Kac-Moody transformations}
1383: 
1384:    Having obtained the Kac-Moody transformations in section \ref{kac-moodysec},
1385: and the ``half'' Virasoro transformations in section \ref{virsubsec}, we may
1386: now compute also the mixed commutators between these two sets of symmetry
1387: transformations.  The calculations are entirely mechanical, although in
1388: some cases somewhat involved, and we shall just present our final results. 
1389: 
1390:    For the Virasoro-like transformations $\delta^V$ defined in 
1391: (\ref{deltaVcV}), (\ref{deltarho}) and (\ref{deltaVX}), we find that their
1392: commutation relations with the Kac-Moody transformations $\delta$ and
1393: $\td\delta$ defined in (\ref{deltacV})--(\ref{tddeltaX}) are given
1394: by
1395: %%%%%
1396: \bea
1397: {[}\delta_1^V, \delta_2{]} \!\!\!\ &=& \!\!\!\
1398:   \fft{\lambda_1\lambda_2}{(\lambda_1-\lambda_2)}\,
1399: \dot\delta(\lambda_2,\ep_2) + 
1400:  \fft{\lambda_1\lambda_2}{(\lambda_1-\lambda_2)^2}\,
1401: \big( \delta(\lambda_2,\ep_2) -\delta(\lambda_1,\ep_2)\big)\,,
1402: \label{Vdelta}\\
1403: %%
1404: {[}\delta_1^V,\td\delta_2{]}\!\!\! &=& \!\!\!\ 
1405:  \fft{\lambda_1\lambda_2^2}{(1-\lambda_1\lambda_2)}\,
1406:   \dot{\td\delta}(\lambda_2,\ep_2) + 
1407:   \fft{\lambda_1\lambda_2}{(1-\lambda_1\lambda_2)^2}\,
1408:    \td\delta(\lambda_2,\ep_2) +
1409:   \fft{\lambda_1\lambda_2}{(1-\lambda_1\lambda_2)^2}\,
1410:   \delta(\lambda_1,\ep_2)\,.\label{Vtddelta}
1411: \eea
1412: %%%%%
1413: (We have calculated these commutators acting on all the fields, namely
1414: $\cV$, $X$ and $\rho$.)
1415: Substituting the expansions (\ref{deltaexps}) and (\ref{virexp1}) into
1416: these expressions, and equating the coefficients of each power of 
1417: $\lambda_1$ and $\lambda_2$, we can read off the commutation relations
1418: %%%%%
1419: \bea
1420: {[}\delta^V_\m, \delta_\n {]} &=& -n \delta_{\sst{(n+m)}}\,,
1421: \qquad m\ge 1\,,\quad n\ge0\,,\label{com1}\\
1422: %%
1423: {[}\delta^V_\m, \td\delta_\n {]} &=& n \td\delta_{\sst{(n-m)}} + 
1424: n \delta_{\sst{(m-n)}}\,,\qquad m\ge 1\,,\quad n\ge1\,,\label{com2}
1425: \eea
1426: %%%%%
1427: where in (\ref{com2}) it is understood that $\delta_{\sst{(m-n)}}$ is zero
1428: if $n> m$ and that $\td\delta_{\sst{(n-m)}}$ is zero if $n\le m$.  The 
1429: commutation relations (\ref{com1}) and (\ref{com2}) can be unified into
1430: one formula if we defined the full set of Kac-Moody modes $\Delta_\n$
1431: as in (\ref{Deltadef}).  We then find
1432: %%%%%
1433: \be
1434: {[}\delta^V_\m, \Delta_\n {]} =  -n \Delta_{\sst{(n+m)}}\,,
1435: \label{mixed1}
1436: \ee
1437: %%%%%
1438: with $m\ge 1$ and $-\infty\le n\le \infty$.
1439: 
1440:    With our previous associations in which $\Delta_\n$ corresponds to
1441: the Kac-Moody generators $J_n^i$, and 
1442: $\delta^V_\m$ corresponds to the Virasoro generator $L_m$ with $m\ge 1$, 
1443: we therefore have
1444: %%%%%
1445: \be
1446: [L_m, J_n^i] = - n J_{n+m}^i\,.\label{VKM}
1447: \ee
1448: %%%%%
1449: Note that it is because $\delta^V_\m$ is associated with the positive
1450: half of the Virasoro algebra, and thus it can only {\it increase} the
1451: Kac-Moody level number in (\ref{VKM}), that (\ref{Vdelta}) has only 
1452: $\delta$ Kac-Moody transformations on the right-hand side, 
1453: whereas (\ref{Vtddelta}) has both $\delta$ and $\td\delta$ Kac-Moody
1454: transformations.
1455: 
1456:    We can also extend the Virasoro-like transformations from $\delta^V$
1457: to $\hat\delta^V$ in these mixed commutator calculations.  The 
1458: necessary calculations are just a simple extension of those already presented,
1459: and the upshot is that one merely has to extend the range of the $m$
1460: index in (\ref{mixed1}) and (\ref{VKM}) down to $m\ge -1$ when $\delta^V$
1461: is replaced with $\hat\delta^V$.
1462: 
1463: \section{Conclusions}
1464: 
1465:    In this paper, we have extended our previous work on the symmetries of
1466: two-dimensional symmetric-space sigma models, by now considering the case
1467: where the sigma model is coupled to gravity.  These gravity-coupled sigma 
1468: models arise in the toroidal reduction of gravity and supergravity
1469: theories from higher dimensions.  There has been some considerable 
1470: discussion of the Kac-Moody symmetries of these models in earlier literature,
1471: and we took the
1472: paper \cite{schwarz2} by Schwarz as the starting point for
1473: our construction.
1474: 
1475:    We were able to improve on the construction of Kac-Moody transformations 
1476: in \cite{schwarz2} in several important respects.  Firstly, we gave a
1477: proper treatment of the use of the ``homogeneous'' Kac-Moody transformations
1478: (\ref{deltahom}) to subtract out the pole in the ``inhomogeneous'' 
1479: transformations (\ref{deltainhom}) that would otherwise occur at 
1480: $\tau_1=\tau_2$.  The fact that this can be done simultaneously at all 
1481: points $x^\mu$ in the two-dimensional spacetime depends upon the quite
1482: remarkable identity (\ref{uvrel2}).  
1483: 
1484:    Secondly, were able to
1485: obtain transformations for the full affine Kac-Moody extension $\hat\cG$ of
1486: the manifest symmetry algebra $\cG$ of the coset $\cG/\cH$.  This result
1487: extends the one obtained in \cite{schwarz2}, where only a certain subalgebra
1488: $\hat\cG_H$ of symmetries was found.  
1489: 
1490:    We also constructed further infinite-dimensional symmetry 
1491: transformations of the SSM fields that are singlets under the original
1492: $\cG$ Lie algebra.  They had not been found in \cite{schwarz2}.  These
1493: symmetries correspond to a subalgebra of the centreless Virasoro algebra,
1494: corresponding to the generators $L_n$ with $n\ge -1$. 
1495: It would be of considerable interest to see if there exist further 
1496: symmetries that could extend this subalgebra to the full Virasoro algebra.
1497: 
1498:    Finally, we remark that all of our symmetry analyses have been restricted
1499: purely to the infinitesimal level.  It would be very interesting to 
1500: extend the methods used in this paper to the non-infinitesimal level.  This
1501: would be of importance both from the perspective of understanding the
1502: U-duality symmetries of string theory, and in order to make use of the
1503: infinite-dimensional symmetries for the purpose of generating new
1504: solutions from old ones. 
1505: 
1506: 
1507: \section*{Acknowledgements}
1508: 
1509:    We are very grateful to John Schwarz for discussions, and for
1510: drawing our attention to references \cite{schwarz1,schwarz2}.  We thank
1511: also Hermann Nicolai for discussions.  This research has
1512: been generously supported by George Mitchell and the Mitchell Family
1513: Foundation.  The research of
1514: H.L. and C.N.P. is also supported in part by DOE grant DE-FG03-95ER40917.
1515: 
1516: 
1517: \newpage
1518: 
1519: \appendix
1520: 
1521: \section{The Flat-Space Limit}\label{appsec}
1522: 
1523: \subsection{Decoupling of gravity}
1524: 
1525:    In a previous paper \cite{lupepo}, we studied the somewhat simpler
1526: problem of symmetric-space coset models in a purely flat two-dimensional
1527: spacetime.  In that case, there is no conformal function $\rho$, and the
1528:  equations (\ref{eom}) and (\ref{cartmau}) become simply
1529: %%%%%
1530: \be
1531: d{*A}=0\,,\qquad dA+ A\wedge A=0\,.
1532: \ee
1533: %%%%
1534: It is evident, therefore, that we can recover the flat-space limit by taking
1535: %%%%%
1536: \be
1537: \rho=\hbox{constant}\,,
1538: \ee
1539: %%%%%
1540: and furthermore this is consistent with its equation of motion (\ref{rhoeom}).
1541: In fact since in general $\rho$ has the solution $\rho=\rho_+(x^+) +
1542:  \rho_-(x^-)$, as in (\ref{rhosol}), we can conveniently parameterise a
1543: family of flat-space limits by 
1544: %%%%%
1545: \be
1546: \rho_+ = 1+\gamma\,,\qquad \rho_- = 1-\gamma\,,\label{alphapar}
1547: \ee
1548: %%%%%
1549: where $\gamma$ is a constant.
1550: 
1551:    In the flat-space limit the spectral function $\tau$ appearing the the
1552: Lax equation (\ref{Lax2}) is a constant, which was denoted by $t$ in 
1553: \cite{lupepo}.   From (\ref{tautheta}) and (\ref{thetasol2}), we see
1554: that (\ref{alphapar}) implies
1555: %%%%%
1556: \be
1557: \lambda= \fft{t}{1+2\gamma\, t + t^2}\,.\label{lamt}
1558: \ee
1559: %%%%%
1560: Thus the constant spectral parameter $t$ used in \cite{lupepo} and the
1561: constant spectral parameter $\lambda$ that we have been using in this paper
1562: are not identical in the flat-space limit.  As a consequence, the
1563: expansions of the Kac-Moody and Virasoro transformations in mode sums
1564: as in (\ref{deltaexps}) and (\ref{virmodes}) will take different forms
1565: depending on whether we use the $t$ parameter or the $\lambda$ parameter
1566: in the flat-space limit, and this amounts to a change of basis for the
1567: algebras.
1568: 
1569:    In what follows, we shall illustrate how the bases for the Kac-Moody and 
1570: Virasoro algebras are related in the ``symmetric'' choice for the flat-space
1571: limit, where we set $\gamma=0$ (and hence $\rho_+=\rho_-=1$).  The calculation
1572: for a general choice of $\gamma$ proceeds in a very similar manner.  With
1573: the choice $\gamma=0$ we have from (\ref{lamt}) that
1574: %%%%%
1575: \be
1576: \lambda=\fft{t}{1+t^2}\,,\qquad t= \fft{1-\sqrt{1-4\lambda^2}}{2\lambda}\,.
1577: \label{lamt2}
1578: \ee
1579: %%%%%
1580: (We choose the negative root in the solution for $t$ so that small $\lambda$
1581: corresponds to small $t$.)
1582: 
1583: \subsection{Kac-Moody symmetries in the flat-space limit}
1584: 
1585:    First, let us consider the Kac-Moody transformations $\delta_1 X_2$
1586: defined in (\ref{deltaX}).  Using (\ref{uvsol}) and (\ref{uvrel2}), we
1587: can first rewrite $\delta_1 X_2$ as
1588: %%%%%
1589: \be
1590: \delta_1 X_2 =\fft{s_1}{\lambda_1\rho}\,\Big[ \fft{\tau_2}{(\tau_1-\tau_2)}\,
1591:     (\eta_1 X_2-X_2\ep_1 X_2^{-1}) + \fft{\tau_1\tau_2}{(1-\tau_1\tau_2)}\,
1592: (M^{-1}\eta_1^\dagger M + X_2\ep_1X_2^{-1})\Big] X_2\,.
1593: \ee
1594: %%%%%
1595: Taking the flat-space limit with $\rho=2$, $\tau_i=t_i$ and $\lambda_i$
1596: related to $t_i$ by (\ref{lamt2}), we therefore have
1597: %%%%%
1598: \bea
1599: \delta_1^{\rm lim} X_2 &=& \fft{1+t_1^2}{1-t_1^2}\, \Big[
1600:   \fft{t_2}{(t_1-t_2)}\, (\eta_1 X_2-X_2\ep_1 X_2^{-1}) +
1601:          \fft{t_1 t_2}{(1-t_1 t_2)}\, M^{-1}\eta_1^\dagger M\Big] X_2\nn\\
1602: %%
1603:  && +
1604:  \fft{1+t_1^2}{1-t_1^2}\, \fft{t_1 t_2}{(1-t_1 t_2)}\, X_2\ep_1\,.
1605: \eea
1606: %%%%%
1607: (The superscript ``lim'' denotes the flat-space limit of the general
1608: gravitationally-coupled transformations we have derived in this paper.)
1609: Comparing with the definitions of the Kac-Moody transformations that
1610: we used in the purely flat-space discussion in \cite{lupepo}, and which
1611: we denote by $\delta_1^{\rm flat}$ and $\td\delta_1^{\rm flat}$ for
1612: this present discussion, we read off that 
1613: %%%%%
1614: \be
1615: \delta_1^{\rm lim} X_2= \fft{1+t_1^2}{1-t_1^2}\,(\delta_1^{\rm flat} 
1616:             +\td\delta_1^{\rm flat}) X_2\,.\label{oldnew1}
1617: \ee
1618: %%%%%
1619: 
1620:    By definition, the modes of $\delta_1^{\rm lim}$ are read off from 
1621: an expansion in powers of $\lambda_1$, as in (\ref{deltaexps}).  Also,
1622: by definition, the modes of $\delta_1^{\rm flat}$ and 
1623: $\td\delta_1^{\rm flat}$ that we used in the flat-space situation in 
1624: \cite{lupepo}) were expanded in powers of $t_1$.  Thus we have only to
1625: substitute the expansions into (\ref{oldnew1}) to obtain
1626: %%%%%
1627: \be
1628: \sum_{n\ge0}\lambda_1^n \delta_\n^{\rm lim} = \fft{1+t_1^2}{1-t_1^2}\,
1629: \Big(\delta_{\sst{(0)}}^{\rm flat} + 
1630:  \sum_{n\ge 1} t_1^n (\delta_\n^{\rm flat} + \td\delta_\n^{\rm flat})
1631:       \Big)\,.
1632: \ee
1633: %%%%%
1634: We now use (\ref{lamt2}) to express $t_1$ in terms of $\lambda_1$ on the
1635: right-hand side, and equate coefficients of each power of $\lambda_1$.  It
1636: is convenient to recall from \cite{lupepo} that, just as we did for the
1637: the gravity-coupled case in this paper in (\ref{Deltadef}), the full set 
1638: of Kac-Moody 
1639: transformations $\Delta_\n^{\rm flat}$ were defined from the non-negative
1640: modes $\delta_\n^{\rm flat}$  and the negative modes $\td\delta_\n^{\rm flat}$
1641: by $\Delta_\n^{\rm flat}= \delta_\n^{\rm flat}$ ($n\ge0$) and 
1642: $\Delta_\n^{\rm flat}= \td\delta_{\sst{(-n)}}^{\rm flat}$ ($n\le -1$).  
1643: We then find that
1644: %%%%%
1645: \bea
1646: \delta^{\rm lim}_{\sst{(0)}} &=& \Delta_{\sst{(0)}}^{\rm flat}\,,\quad
1647: \delta^{\rm lim}_{\sst{(1)}} = \Delta_{\sst{(-1)}}^{\rm flat} +
1648:                      \Delta_{\sst{(1)}}^{\rm flat}\,,\quad
1649: \delta^{\rm lim}_{\sst{(2)}} = \Delta_{\sst{(-2)}}^{\rm flat} +
1650:                           2 \Delta_{\sst{(0)}}^{\rm flat}+
1651:                      \Delta_{\sst{(2)}}^{\rm flat}\,,\nn\\
1652: %%
1653: \delta^{\rm lim}_{\sst{(3)}} &=& \Delta_{\sst{(-3)}}^{\rm flat} +
1654:                           3 \Delta_{\sst{(-1)}}^{\rm flat}+
1655:                          3 \Delta_{\sst{(1)}}^{\rm flat}+
1656:                      \Delta_{\sst{(3)}}^{\rm flat}\,,
1657: \eea
1658: %%%%%
1659: and so on, with the general $n$ case given by
1660: %%%%%
1661: \be
1662: \delta^{\rm lim}_\n = \sum_{p=0}^n C^n_p\, \Delta_{\sst{(n-2p)}}^{\rm flat}\,,
1663: \qquad C^n_p\equiv \fft{n!}{p!\, (n-p)!}\,.
1664: \ee
1665: %%%%%
1666: This shows that the non-negative half of the Kac-Moody algebra in the
1667: flat-space limit of the gravity-coupled models is related to a combination
1668: of the positive and negative halves of the Kac-Moody algebra that arose in 
1669: the previous flat-space discussion in \cite{lupepo}. 
1670: 
1671:    We turn now to the negative half of the Kac-Moody algebra of
1672: the present paper, described by $\td\delta_1$ in (\ref{tddeltaX}).  We have
1673: %%%%%
1674: \be
1675: \td\delta_1 X_2 = \fft{\lambda_1\lambda_2}{(1-\lambda_1\lambda_2)}\,
1676:      X_2 \ep_1\,,
1677: \ee
1678: %%%%%
1679: from which it follows that
1680: %%%%%
1681: \be
1682: \td\delta_\n^{\rm lim} X_2 = \lambda_2^n\, X_2\ep_1\,.\label{lamv}
1683: \ee
1684: %%%%%
1685: Equally, in the flat-space expansion from \cite{lupepo} we have
1686: %%%%%
1687: \be
1688: \td\delta_\n^{\rm flat} X_2 = t_2^n\, X_2\ep_1\,.\label{tv}
1689: \ee
1690: %%%%%
1691: By substituting (\ref{lamt2}) into (\ref{lamv}), expanding the 
1692: right-hand side in powers of $t_2$, and then using (\ref{tv}), we can 
1693: read off the expressions for the modes $\td\delta_\n^{\rm lim}$ in terms
1694: of the modes $\td\delta_\n^{\rm flat}$.  The first few modes are 
1695: given by
1696: %%%%%
1697: \bea
1698: \td\delta_{\sst{(1)}}^{\rm lim} &=& \td\delta_{\sst{(1)}}^{\rm flat}
1699:  -\td\delta_{\sst{(3)}}^{\rm flat} + \td\delta_{\sst{(5)}}^{\rm flat}
1700:   -\td\delta_{\sst{(7)}}^{\rm flat}+\cdots\,,\nn\\
1701: %%
1702: \td\delta_{\sst{(2)}}^{\rm lim} &=& \td\delta_{\sst{(2)}}^{\rm flat}
1703:  -2\td\delta_{\sst{(4)}}^{\rm flat} + 3\td\delta_{\sst{(6)}}^{\rm flat}
1704:   -4\td\delta_{\sst{(8)}}^{\rm flat}+\cdots\,,\nn\\
1705: %%
1706: \td\delta_{\sst{(3)}}^{\rm lim} &=& \td\delta_{\sst{(3)}}^{\rm flat}
1707:  -3\td\delta_{\sst{(5)}}^{\rm flat} + 6\td\delta_{\sst{(7)}}^{\rm flat}
1708:   -10\td\delta_{\sst{(9)}}^{\rm flat}+\cdots\,,
1709: \eea
1710: %%%%%
1711: and the general case is given by
1712: %%%%%
1713: \be
1714: \td\delta_\n^{\rm lim}= \sum_{p=0}^\infty \fft{(-1)^p\,(n+p-1)!}{p!\, (n-1)!}\,
1715:        \td\delta_{\sst{(n+2p)}}^{\rm flat}\,,\qquad n\ge 1\,.
1716: \ee
1717: %%%%%
1718: 
1719:     Combining our results for $\delta_\n^{\rm lim}$ and 
1720: $\td\delta_\n^{\rm lim}$, and using the definition (\ref{Deltadef}) for
1721: the full set of Kac-Moody transformations, we therefore have
1722: %%%%%
1723: \bea
1724: \Delta_\n^{\rm lim} &=& \sum_{p=0}^n C^n_p\, 
1725: \Delta^{\rm flat}_{\sst{(n-2p)}}\,,\qquad
1726:    n\ge 0\label{dela}\\
1727: %%
1728: \Delta_\n^{\rm lim} &=& \sum_{p=0}^n (-1)^p\, C^{p-n-1}_p\, 
1729:    \Delta^{\rm flat}_{\sst{(n-2p)}}\,,\label{delb}
1730: \qquad  n\le -1\,.
1731: \eea
1732: %%%%%
1733: Now if $n$ is temporarily generalised to be an arbitrary real variable 
1734: whilst $p$ is an
1735: integer, then we have $C^n_p/C^{p-n-1}_p =(-1)^p$, and so we see 
1736: that (\ref{dela})
1737: and (\ref{delb}) can really be combined into a single formula of the 
1738: form (\ref{dela}), with the $p$ summation appropriately extended;
1739: %%%%%
1740: \be
1741: \Delta_\n^{\rm lim} = \sum_{p=0}^\infty C^n_p\,
1742: \Delta^{\rm flat}_{\sst{(n-2p)}}\,,\qquad n\in\Z\,.
1743: \ee
1744: %%%%%
1745: 
1746:    A
1747: direct
1748: demonstration that $\Delta_\n^{\rm lim}$ satisfies the standard Kac-Moody
1749: algebra (\ref{kacmoody}) if $\Delta_\n^{\rm flat}$ satisfies the 
1750: standard Kac-Moody algebra
1751: can easily be given, based on the identity that 
1752: $(1+t)^p (1+t)^q = (1+t)^{p+q}$.
1753: 
1754: \subsection{Virasoro-type symmetries in the flat-space limit}
1755: 
1756:     In our discussion of the Virasoro-type symmetries in section 
1757: \ref{virasorosec}, we found that not only the fields $\cV$ and $X$, but
1758: also the field $\rho$, is subject to these transformations, 
1759: as in (\ref{finalV}).  When we take the flat-space limit we are fixing 
1760: $\rho$ to a constant, and consequently some of the Virasoro-type symmetry
1761: of the general gravity-coupled case will be broken.  In fact, as we shall
1762: see below, it is the two ``extra'' generators $\hat\delta^V_{\sst{(0)}}$ and 
1763: $\hat\delta^V_{\sst{(-1)}}$, whose action is given in (\ref{virextra}), that
1764: are broken when we fix $\rho$ (and also $\rho_+$ and $\rho_-$ separately).  
1765: In fact these two transformations will be treated as compensating 
1766: transformations that restore $\rho_+$ and $\rho_-$ to their chosen fixed
1767: values when the remaining Virasoro-type transformations  
1768: $\hat\delta^V_\n$ with $n\ge1$ act.  This loss of the $n=0$ and $n=-1$ 
1769: symmetries in the flat-space limit is consistent with the fact that they
1770: were never seen in the purely flat-space results in \cite{lupepo}.
1771: 
1772:    It can be seen from (\ref{deltarhoprhom}) (with $\beta=0$) and  
1773: (\ref{virextra}) that in order to preserve the flat-space limit with
1774: $\rho_+=\rho_-=1$, the $\delta_1^V$ Virasoro transformations for the
1775: modes $n\ge1$ should be accompanied by $\hat\delta^V_{\sst{(0)}}$ and
1776: $\hat\delta^V_{\sst{(-1)}}$ compensating transformations in the combination
1777: %%%%%
1778: \be
1779: \td\delta_1^V = \delta_1^V - s_1^2\, \hat\delta^V_{\sst{(0)}} -
1780:                      2 s_1 c_1\, \hat\delta^V_{\sst{(-1)}}\,.
1781: \label{compvir}
1782: \ee
1783: %%%%%
1784: We can therefore now pass to the flat-space limit by setting 
1785: $\rho_+=\rho_-=1$, and calculate the commutators of these
1786: compensated Virasoro-type transformations, using the expressions for the
1787: commutators of the transformations appearing on the right-hand side of
1788: (\ref{compvir}) that we have obtained in section \ref{virasorosec}.  Thus
1789: we find that
1790: %%%%%
1791: \bea
1792: {[} \td\delta_1^V,\td\delta_2^V {]} &=& 
1793: \fft{\lambda_1\lambda_2}{\lambda_1-\lambda_2}\, \left[
1794: \fft{1-4\lambda_1^2}{1-4\lambda_2^2}\, \del_{\lambda_1} \td\delta_1^V +
1795:  \fft{1-4\lambda_2^2}{1-4\lambda_1^2}\, \del_{\lambda_2} 
1796:   \td\delta_2^V\right]\nn\\
1797: %%
1798: &&- \fft{2\lambda_1\lambda_2}{(\lambda_1-\lambda_2)^2}\,
1799: \left[ \fft{1+4\lambda_1^2-8\lambda_1\lambda_2}{1-4\lambda_2^2}\, 
1800:     \td\delta_1^V -
1801:    \fft{1+4\lambda_2^2-8\lambda_1\lambda_2}{1-4\lambda_1^2}\,
1802:     \td\delta_2^V\right]\,.\label{tdVcom}
1803: \eea
1804: %%%%%
1805: 
1806:    Since $\cV$ is inert under $\delta_{\sst{(0)}}^V$ and 
1807:   $\delta_{\sst{(-1)}}^V$, it follows that $\td\delta_1^V\cV=\delta_1^V \cV$,
1808: and so from (\ref{deltaVcV}) and (\ref{hsol2}) we have, after taking the
1809: flat-space limit in which $\lambda$ and $t=\tau$ are related by (\ref{lamt2}),
1810: %%%%%
1811: \bea
1812: \td\delta^V \cV &=& -\fft{t}{1-t^2}\, \cV \big( \del_{\lambda} X\big) X^{-1}
1813:                             \,,\nn\\
1814: %%
1815: &=& -\fft{t}{1-t^2}\, \fft{\del t}{\del\lambda}\, 
1816:              \big( \del_t X\big) X^{-1}\,,\nn\\
1817: %%
1818:   &=& \left(\fft{1+t^2}{1-t^2}\right)^2\, \delta^{V,\, {\rm flat}}\, \cV\,,
1819: \label{virlimflat}
1820: \eea
1821: %%%%%
1822: where $\delta^{V,\, {\rm flat}}$ is the Virasoro-type transformation that
1823: we found in the purely flat-space discussion in \cite{lupepo}.  (It is
1824: given by $\delta^{V,\, {\rm flat}} \cV = -t \big(\del_t X(t)\big) X(t)^{-1}$.)
1825: We have also verified that the same relation (\ref{virlimflat}) holds
1826: when acting on $X$ instead of $\cV$.  
1827: Using (\ref{virlimflat}), we can translate the commutator (\ref{tdVcom}) into
1828: the commutator ${[} \delta^{V,\, {\rm flat}}_1, \delta^{V,\, {\rm flat}}_2 
1829:  {]}$, finding
1830: %%%%%
1831: \bea
1832: {[} \delta^{V,\, {\rm flat}}_1, \delta^{V,\, {\rm flat}}_2 {]} &=&
1833: -2 t_1 t_2 \left[ \fft{1}{(t_1-t_2)^2} +\fft1{(1-t_1 t_2)^2}\right]
1834:   \, \delta_1^{V,\,{\rm flat}}
1835:   +\fft{t_1t_2(1-t_1^2)}{(t_1-t_2)(1-t_1 t_2)} \, \del_{t_1}
1836: \delta_1^{V\, {\rm flat}}\nn\\
1837: %%
1838: &&
1839: -[1\leftrightarrow 2]\,,\label{comVM}
1840: \eea
1841: %%%%%%
1842: This agrees precisely with the commutator of Virasoro-type transformations 
1843: that we found previously for the flat space models in \cite{lupepo}.
1844: 
1845:    The relation (\ref{virlimflat}) can be re-expressed as a relation
1846: between the modes in the expansions of the two variations, namely
1847: %%%%%
1848: \be
1849: \td\delta_1^V =\sum_{n\ge1} \lambda_1^n\, \td\delta_\n^V\,,\qquad
1850: \delta_1^{V\,{\rm flat}} = \sum_{n\ge 1} t_1^n\,
1851:      \delta_\n^{V\,{\rm flat}}\,.
1852: \ee
1853: %%%%%
1854: For the first few levels we find
1855: %%%%%
1856: \bea
1857:  \td\delta_{\sst{(1)}}^V &=& \delta_{\sst{(1)}}^{V\,{\rm flat}}\,,\qquad
1858:  \td\delta_{\sst{(2)}}^V = \delta_{\sst{(2)}}^{V\,{\rm flat}}\,,\qquad
1859: \td\delta_{\sst{(3)}}^V = \delta_{\sst{(3)}}^{V\,{\rm flat}} +
1860:                       5 \delta_{\sst{(1)}}^{V\,{\rm flat}}\,,\nn\\
1861: %%
1862: \td\delta_{\sst{(4)}}^V &=& \delta_{\sst{(4)}}^{V\,{\rm flat}} +
1863:                       6 \delta_{\sst{(2)}}^{V\,{\rm flat}}\,,\qquad
1864: \td\delta_{\sst{(5)}}^V = \delta_{\sst{(5)}}^{V\,{\rm flat}} +
1865:                       7 \delta_{\sst{(3)}}^{V\,{\rm flat}}+
1866:                      22 \delta_{\sst{(1)}}^{V\,{\rm flat}}\,.
1867: \eea
1868: %%%%%
1869: 
1870: 
1871: \newpage
1872: 
1873: \begin{thebibliography}{99}
1874: 
1875: \bm{crejul} E. Cremmer and B. Julia, {\it The $N=8$ supergravity theory. 1:
1876: The Lagrangian}, Phys. Lett. {\bf B80}, 48 (1978);
1877: {\it The $SO(8)$ supergravity},  Nucl. Phys. {\bf B159}, 141 (1979).
1878: 
1879: \bm{cj2} B. Julia, {\it Group disintegrations}; E. Cremmer, {\it Supergravity
1880: in 5 dimensions}, in {\sl Superspace and Supergravity}, Eds. S.W. Hawking and
1881: M. Rocek (Cambridge University Press, 1981) 331; 267.
1882: 
1883: \bm{cjlp} E. Cremmer, B. Julia, H. L\"u and C.N. Pope,
1884: {\it Dualisation of dualities. I},
1885: Nucl. Phys. {\bf B523}, 73 (1998), hep-th/9710119.
1886: %%CITATION = NUPHA,B523,73;%%
1887: 
1888: \bm{geroch} R. Geroch, {\it  A method for generating new solutions of
1889: Einstein's equations 2}, J. Math. Phys. {\bf 13}, 394 (1972).
1890: 
1891: \bm{lupepo} H. L\"u, M.J. Perry and C.N. Pope,
1892: {\it Infinite-dimensional symmetries of two-dimensional coset models},
1893: arXiv:0711.0400 [hep-th].
1894: %%CITATION = ARXIV:0711.0400;%%
1895: 
1896: \bm{pol} K. Pohlmeyer,
1897: {\it Integrable Hamiltonian systems and interactions through quadratic
1898: constraints},
1899: Commun. Math. Phys. {\bf 46}, 207 (1976).
1900: %%CITATION = CMPHA,46,207;%%
1901: 
1902: \bm{lus} M. L\"uscher,
1903: {\it Quantum nonlocal charges and absence of particle production in the
1904: two-dimensional nonlinear sigma model},
1905: Nucl. Phys. {\bf B135}, 1 (1978).
1906: %%CITATION = NUPHA,B135,1;%%
1907: 
1908: \bm{polus} M. L\"uscher and K. Pohlmeyer,
1909: {\it Scattering of massless lumps and nonlocal charges in the
1910: two-dimensional classical nonlinear sigma model},
1911: Nucl. Phys. {\bf B137}, 46 (1978).
1912: %%CITATION = NUPHA,B137,46;%%
1913: 
1914: \bm{kinchi} W. Kinnersley and D.M. Chitre,
1915: {\it Symmetries of the stationary Einstein-Maxwell field equations. II, III},
1916: J. Math. Phys. {\bf 18}, 1538 (1977); J. Math. Phys. {\bf 19}, 1926 (1978).
1917: 
1918: \bm{belzak} V.A. Belinsky and V.E. Zakharov,
1919: {\it Integration of the Einstein equations by the inverse scattering problem
1920: technique and the calculation of the exact soliton solutions},
1921: Sov. Phys. JETP {\bf 48}, 985 (1978); Zh. Eksp. Teor. Fiz.  {\bf 75}, 1953
1922: (1978).
1923:   %%CITATION = ZETFA,75,1953;%%
1924: 
1925: \bm{mais1} D. Maison,
1926: {\it Are the stationary, axially symmetric Einstein equations completely
1927: integrable?},
1928: Phys. Rev. Lett. {\bf 41}, 521 (1978).
1929: %%CITATION = PRLTA,41,521;%%
1930: 
1931: \bm{breitmais} P. Breitenlohner and D. Maison,
1932: {\it On the Geroch group},
1933: Ann. Inst. H. Poincar\'e, {\bf 46}, 215 (1987).
1934: %%CITATION = AHPAA,46,215;%%
1935: 
1936: \bm{dolan} L. Dolan,
1937: {\it Kac-Moody algebras and exact solvability in hadronic physics},
1938: Phys. Rept.  {\bf 109}, 1 (1984).
1939: %%CITATION = PRPLC,109,1;%%
1940: 
1941: \bm{devfair} C. Devchand and D.B. Fairlie,
1942: {\it A generating function for hidden symmetries of chiral models},
1943: Nucl. Phys. {\bf B194}, 232 (1982).
1944: %%CITATION = NUPHA,B194,232;%%
1945: 
1946: \bm{yswu} Y.S. Wu,
1947: {\it Extension of the hidden symmetry algebra in classical principal chiral
1948: models}, Nucl. Phys. {\bf B211}, 160 (1983).
1949: %%CITATION = NUPHA,B211,160;%%
1950: 
1951: \bm{jul} B. Julia, {\it On infinite-dimensional symmetry groups in physics},
1952: published in Niels Bohr Symposium, 1985: 0215.
1953: 
1954: \bm{nicolai}  H.~Nicolai,
1955: {\it Two-dimensional gravities and supergravities as integrable system},
1956: in ``Schladming 1991, Proceedings, Recent aspects of quantum fields,'' 231,
1957: and Hamburg DESY - DESY 91-038
1958: %%CITATION = C91-02-27;%%
1959: 
1960: \bm{nicjul} B. Julia and H. Nicolai, {\it Conformal internal symmetry of
1961: 2d sigma-models coupled to gravity and  a dilaton},
1962: Nucl. Phys. {\bf B482}, 431 (1996), hep-th/9608082.
1963: %%CITATION = NUPHA,B482,431;%%
1964: 
1965: \bm{schwarz1} J.H. Schwarz,
1966: {\it Classical symmetries of some two-dimensional models},
1967: Nucl. Phys. {\bf B447}, 137 (1995), hep-th/9503078.
1968: %%CITATION = NUPHA,B447,137;%%
1969: 
1970: \bm{schwarz2} J.H. Schwarz,
1971: {\it Classical symmetries of some two-dimensional models coupled to
1972: gravity},
1973: Nucl. Phys. {\bf B454}, 427 (1995), hep-th/9506076.
1974: %%CITATION = NUPHA,B454,427;%%
1975: 
1976: \bm{houli} B.Y. Hou and W. Li, 
1977: {\it Virasoro algebra in the solution space of the Ernst equation},
1978: Lett. Math. Phys. {\bf 13}, 1 (1987).
1979: 
1980: \bm{maison} D. Maison, {\it Geroch group and inverse scattering 
1981:   method,} MPI-PAE/PTh 80/88 (1988).
1982: 
1983: 
1984: \end{thebibliography}
1985: 
1986: \end{document}
1987: