hep-th0610284/RG.tex
1: \chapter{Out of the conformal point: Renormalization Group Flows}
2: \label{cha:renorm-group-flows}
3: 
4: \chapterprecis{This chapter is devoted to the study of the relaxation
5:   of squashed \textsc{wzw} models further deformed by the insertion of
6:   non-marginal operators. The calculation is carried from both the
7:   target space and world-sheet points of view, once more highlighting
8:   the interplay between the two complementary descriptions. In the
9:   last part such techniques are used to outline the connection between
10:   the time evolution and the \textsc{rg}-flow which is seen as a large
11:   friction limit description; we are hence naturally led to a
12:   \textsc{frw}-type cosmological model.}
13: 
14: \lettrine{S}{tring theory}, at least in its world-sheet formulation,
15: is most easily studied on-shell. Thanks to the power of conformal
16: field theory, this permits a profound analysis of specific
17: backgrounds.  At the same time, though, it makes it difficult to
18: describe more general effects that require a less local knowledge of
19: the theory and its moduli space. In particular it is not obvious how
20: to describe transitions between two different solutions or even the
21: relaxation of an unstable background towards an on-shell solution.
22: 
23: In this chapter we deviate from conformality by adding non-marginal
24: deformations on the top of exact solutions, such as
25: \textsc{wzw}-models or squashed-group models. The resulting
26: \textsc{rg}-flow then drives those systems back to or away from the
27: conformal point, depending on the character of the deformation. As it
28: is usually the case, these calculations can be faced from two
29: complementary points of view: either in terms of the target space
30: description or in terms of world-sheet two-dimensional theory. We will
31: consider both approaches and show how they do really complete each
32: other, in the sense that they can be considered as two different
33: series expansions of the same quantity. As such, each side contains
34: more information than the other at any given order in perturbation.
35: This will allow us in particular to make a prediction on the outcome
36: of a technically involved one-loop calculation in the \textsc{wzw} and
37: squashed group \textsc{cft}s on the basis of a two-loop result on
38: target space renormalization.
39: 
40: In the last part of this chapter we use the technology developed so
41: far to show how an \textsc{rg}-flow analysis can allow for further
42: insights on the issue of time-dependent solutions. More precisely we
43: will see how for a given class of systems whose geometry is the warped
44: product of a constant curvature space and a time direction the
45: \textsc{rg}-flow equations are a sort of large-friction approximation
46: with the central charge deficit playing the r\^ole of an effective
47: friction coefficient.
48: 
49: 
50: \section{The target space point of view}
51: \label{sec:target-space-renormalization}
52: 
53: 
54: \subsection{Renormalization in a dimensional regularization scheme}
55: \label{sec:renorm-dimens-regul}
56: 
57: \danger Consider the $\sigma$-model with Lagrangian density
58: \begin{equation}
59:   \mathcal{L} = \frac{1}{2 \lambda} \left( g_{ij} + B_{ij} \right) 
60:   \Xi^{ij} ,
61: \end{equation}
62: where $g_{ij} $ is a metric, $B_{ij}$ a two-form and $\Xi^{ij} = \partial_\mu X^i
63: \partial^\mu X^j + \epsilon_{\mu \nu} \partial^\mu X^i \partial^\nu X^j$. We will say that the model is
64: renormalizable if the corresponding counterterms at any given order in
65: the loop expansion can be reabsorbed into a renormalization of the
66: coupling constant and other parameters that appear in the expressions
67: for $g_{ij}$ and $B_{ij}$.\footnote{When this is not the case the
68:   model might nevertheless be renormalizable in a more general sense,
69:   in the infinite-dimensional space of metrics and torsions}
70: 
71: The standard technique for dealing with this kind of Lagrangian
72: consists in incorporating the Kalb--Ramond field (or, equivalently,
73: the \textsc{wz} term) into the geometry by reading it as a
74: torsion. This means that instead of the usual Levi-Civita connection
75: one uses the connection $\Gamma^-$ defined as
76: \begin{equation}
77:     {\Gamma^-}^{i}_{\phantom{i}jk}  = {i
78:   \atopwithdelims\{\} j k } - \frac{1}{2} H^i_{\phantom{i}jk } .  
79: \end{equation}
80: where ${i \atopwithdelims\{\} j k }$ is the Christoffel symbol and with
81: respect to this connection one defines the Riemann and Ricci tensors
82: $R^-$ and $Ric^-$.
83: 
84: \marginlabel{Two-loop target space \textsc{rg}-flow}Now, using the
85: background field method in a dimensional regularization scheme (see
86: \cite{Osborn:1989bu,Alvarez-Gaume:1981hn,Friedan:1980jm,Friedan:1980jf,Braaten:1985is,Hull:1987pc}
87: and for various applications
88: \cite{Balog:1996im,Balog:1998br,Sfetsos:1998kr,Sfetsos:1999zm}) we can
89: calculate the one- and two\nb-loop counterterms that turn out to be:
90: \begin{align}
91:   \mu^\varepsilon \mathcal{L}^{(1)} &= \frac{1}{\pi \varepsilon}
92:   T^{(1)} = \frac{\alpha^\prime}{2 \varepsilon \lambda } {Ric^-}_{ij}
93:   \Xi^{ij} ,\\
94:   \mu^\varepsilon \mathcal{L}^{(2)} &= \frac{\lambda}{8 \pi^2
95:     \varepsilon} T^{(2)} = \frac{\alpha^{\prime 2}}{16 \varepsilon
96:     \lambda } Y^{lmk}_{\phantom{lmk}j}{R^-}_{iklm} \Xi^{ij} ,
97: \end{align}
98: where $Y$ is given by
99:  \begin{equation}
100:   Y_{i j k l } = -2 {R^-}_{i j k l } + 3 {R^-}_{[k i j ] l } +
101:   \frac{1}{2} \left( {H^2}_{k i } g_{j l } - {H^2}_{k j } g_{i l } 
102:   \right) ,
103: \end{equation}
104: and
105: \begin{equation}
106:   {H^2}_{ij} = {H_{[3]}}_{ilm} {H_{[3]}}_j^{\phantom{j}lm} .
107: \end{equation}
108: 
109: In general the metric and the Kalb--Ramond field depend on a set of
110: bare parameters $a^{(0)}_k$. In this case we can convert the
111: counterterms given above into coupling and parameter renormalizations
112: if we write perturbatively the bare quantities as:
113: \begin{equation}
114: \label{eq:bare-stuff}
115:   \begin{cases}
116:     \lambda^{(0)} = \mu^\varepsilon \lambda \left( 1 + \frac{J_1
117:         (a)}{\pi \varepsilon} \lambda + \ldots \right) =
118:     \mu^\varepsilon \lambda \left( 1 + \frac{y_\lambda}{\epsilon} +
119:       \ldots \right) ,\\
120:     a^{(0)}_k = a_k + \frac{a_k^{(1)}(a)}{\pi \varepsilon} + \ldots =
121:     a_k \left( 1 + \frac{y_{a_k}}{\epsilon} + \ldots \right) ,\\
122:     X^{(0) \mu } = X^\mu + \frac{X^{(1) \mu }( X, a)}{\pi \varepsilon} + \ldots , 
123:   \end{cases}
124: \end{equation}
125: where we allowed for a slight generalization with respect to the
126: definition of renormalization given above in terms of a coordinate
127: reparametrization of the target space\footnote{This redefinition is in
128:   general non-linear; in the special case when the $X^{(i)}$'s depend
129:   linearly on $X$ the last equation of the system~\ref{eq:bare-stuff}
130:   reduces to a multiplicative wavefunction renormalization. The only
131:   condition is that $X^{(i)}$ shouldn't depend on the derivatives of
132:   $X$. In more geometric terms we are just using the diffeomorphism
133:   invariance of the renormalized theory.}. Then the one- and
134: two\nb-loop $\beta$\nb-functions are given by:
135: \begin{equation}
136:   \begin{cases}
137:     \beta_\lambda = \displaystyle{\Deriv{\lambda}{t} = \lambda^2
138:     \deriv{y_\lambda}{\lambda}} = \frac{\lambda^2 }{\pi} \left(
139:       J^{(1)} (a) + \frac{\lambda}{4 \pi} J^{(2)} (a) \right) ,\\
140:     \beta_{a_k} = \displaystyle{\Deriv{a_k}{t}= \lambda a_k
141:     {\deriv{y_{a_k}}{\lambda}}} = \frac{\lambda }{\pi} \left( a^{(1)}_k
142:       (a) + \frac{\lambda}{4 \pi} a^{(2)}_k (a) \right) .
143:   \end{cases}
144: \end{equation}
145: 
146: The unknown functions $J^{(i)}, a^{(i)}, X^{(i)\mu}$ are determined by
147: the equation
148: \begin{equation}
149: \label{eq:renorm-counter}
150:   T^{(i)} = - J^{(i)} \mathcal{L} + \deriv{\mathcal{L}}{a_k} a_k^{(i)} 
151:   + \deriv{\mathcal{L}}{X^{\mu}} X^{(i)\mu}.
152: \end{equation}
153: This corresponds to demanding the generalized quantum effective action
154: $\Gamma[X]$ to be finite order by order.
155: 
156: 
157: \subsection{Two-loop $\beta$ equations for a \textsc{wzw} model.}
158: \label{sec:two-loop-equations}
159: 
160: As we have already announced in Ch.~\ref{cha:wess-zumino-witten}, the
161: normalization for the \textsc{wz} term in a \textsc{wzw} action can be
162: fixed by an \textsc{rg}-flow calculation. This is precisely what we
163: will do in the following at two-loop order using the dimensional
164: regularization scheme outlined above. In this way we will find a new
165: apparent non-trivial solution that doesn't show up at first order (and
166: which will prove to be an artifact as we'll see in the following, by
167: using a \textsc{cft} description in Sec.~\ref{sec:cft-approach-1}).
168: Moreover we will see how the roles of the \textsc{ir} and \textsc{uv}
169: limits are interchanged between the compact and the non-compact case,
170: \emph{ie} how the same kind of deformation is relevant or irrelevant
171: depending on the compactness of the starting model.
172: 
173: Consider the following action
174: \begin{equation}
175:   S_{\lambda,H} = \frac{1}{2 \lambda } \int_\Sigma \di^2 z \:
176:   \left( g_{ab} + \h B_{ab}\right) J^a_\mu J^b_\nu \partial
177:   X^\mu \bar \partial X^\nu 
178: \end{equation}
179: where $J^a$ are the Maurer--Cartan one-forms for a group $G$ whose
180: algebra has structure constants $f_{abc}$, $g_{ab} = -1/(2 g^\ast)
181: \F{t}{as} \F{s}{bt}$ ($g^\ast $ is the dual Coxeter number) and
182: $B_{ab}$ is an antisymmetric matrix satisfying $ \di \left(B_{ab} J^a
183:   \land J^b\right) = 1/3! f_{abc} J^a \land J^b \land J^c$ as in
184: Sec.~\ref{sec:target-space-wzw}. Since the deformation (parameterized
185: by having $\h \neq 1$) doesn't affect the geometric part (but for the
186: overall normalization) we can still express the curvature in terms of
187: the Lie algebra structure constants.  In particular it is easy to
188: recover that the Riemann tensor is written as:
189: \begin{equation}
190:   R^a_{\phantom{a}bcd} = \frac{1}{4}\F{a}{be} \F{e}{cd}
191: \end{equation}
192: and the Ricci tensor is obtained by contracting:
193: \begin{equation}
194:   Ric_{ab} = \frac{g^\ast }{2} g_{ab}.
195: \end{equation}
196: as in Eqs.~\eqref{eq:group-curvatures}.
197: 
198: In order to write the beta equations as described above we need to
199: incorporate the \textsc{wz} term (or, more precisely, the Kalb--Ramond
200: field) into the geometry. The most natural approach is to consider
201: $H^a_{\phantom{a}bc}$ as a torsion and include it in the
202: connection~\cite{Braaten:1985is}.  We hence define:
203: \begin{equation}
204:   \label{eq:minus-connection}
205:   {\Gamma^-}^{a}_{\phantom{a}bc}  = {a
206:   \atopwithdelims\{\} b c } - \frac{1}{2} H^a_{\phantom{a}bc } .
207: \end{equation}
208: The covariant derivative of a one-form is then defined as:
209: \begin{equation}
210:   {\nabla^-}_a V_b = \partial_a V_b - {\Gamma^-}^c_{\phantom{c}ab} V_c =
211:   \nabla_a V_b + \frac{1}{2} H^c_{\phantom{c}ab } V_c
212: \end{equation}
213: where $\nabla_a$ is the covariant derivative with respect to the
214: Levi--Civita connection. Similarly we define the curvature:
215: \begin{equation}
216:   \comm{{\nabla^-}_a, {\nabla^-}_b} V_c = {R^-}_{c \phantom{d} a b}^{\phantom{c}d} V_d +
217:   H^d_{\phantom{d}ab} {\nabla^-}_d V_c 
218: \end{equation}
219: and it is straightforward to show that:
220: \begin{equation}
221:   {R^-}_{abcd} = R_{abcd} + \frac{1}{2} \nabla_c H_{abd} - \frac{1}{2} \nabla_d
222:   H_{abc} + \frac{1}{4} H_{fac} H^f_{\phantom{f}db} - \frac{1}{4} H_{fad}
223:   H^{f}_{\phantom{f}cb}. 
224: \end{equation}
225: 
226: Let us now specialize this general relation to our particular
227: deformation. Since $H_{abc} = \h f_{abc} $ it is immediate that
228: $\nabla_a H_{bcd} = 0$ and that the Jacobi identity holds.  We
229: then derive:
230: \begin{equation}
231:   {R^-}_{abcd} = \left( 1 - \h^2 \right)R_{abcd} ,
232: \end{equation}
233: whence in particular:
234: \begin{equation}
235:   {Ric^-}_{ab}  = \left( 1 - \h^2\right) Ric_{ab} = \left( 1 - \h^2 \right)
236:   \frac{g^\ast }{2} g_{ab}.
237: \end{equation}
238: The one-loop counterterm becomes:
239: \begin{equation}
240:   T^{(1)}_{ab} = \frac{g^\ast }{4} \left( 1 - \h^2 \right) g_{ab} .
241: \end{equation}
242: 
243: The evaluation of the two-loop counterterm is lengthy but straightforward
244: once $R_{abcd}$ is written in terms of the structure constants. The result
245: is:
246: \begin{equation}
247:    T^{(2)}_{ab} = \frac{{g^\ast}^2 }{8} \left( 1-\h^2\right)
248:    \left( 1-3 \h^2\right) g_{ab} .
249: \end{equation}
250: 
251: Substituting these expressions in Eq.~(\ref{eq:renorm-counter}) (and
252: using the fact that $g_{ab}$ and $\epsilon_{ab}$ are orthogonal) one
253: sees that they become identities for the following choice of
254: parameters:
255: \begin{align}
256:   \begin{cases}
257:     J^{(1)} = - \frac{g^\ast }{4} \left( 1 - \h^2\right) \\
258:     a_\h^{(1)} = - \frac{g^\ast }{4} \h \left( 1 - \h^2\right)
259:   \end{cases} &&
260:   \begin{cases}
261:     J^{(2)} = - \frac{{g^\ast}^2 }{8}  \left( 1 - \h^2 \right)
262:     \left( 1 - 3 \h^2 \right) \\
263:     a_\h^{(2)} = - \frac{{g^\ast}^2 }{8} \h \left( 1 - \h^2 \right)
264:     \left( 1 - 3 \h^2 \right)
265:   \end{cases}
266: \end{align}
267: corresponding to the following beta equations:
268: \begin{align}
269:   \beta_\h &= - \frac{1}{4 \pi } \lambda^\ast \h \left( 1 - \h^2 \right)
270:   \left( 1 +
271:     \frac{1 }{8 \pi } \lambda^\ast \left( 1-3 \h^2 \right)\right) \\
272:   \beta_{\lambda^\ast} &= - \frac{1 }{4 \pi } {\lambda^\ast}^2 \left( 1 - \h^2
273:   \right) \left( 1 + \frac{1 }{8 \pi } \lambda^\ast \left( 1-3 \h^2
274:     \right)\right)
275: \end{align}
276: where $\lambda^\ast = g^\ast \lambda$ is the effective coupling
277: constant (this is precisely the fixed parameter in a 't~Hooft limit
278: since for a $SU(N) $ group $g^\ast = N$). The difference between a
279: compact and a non-compact group lies in the sign of the dual Coxeter
280: number that is respectively positive/negative. In both cases we remark
281: that $\h / \lambda^\ast$ remains constant, which is a nice check of
282: our construction, since in the notation of
283: Ch.~\ref{cha:wess-zumino-witten} this is just the level of the model
284: that, in the compact case, is quantized and hence is not expected to
285: receive any perturbative correction. On the other hand non
286: perturbative effects do eventually lead to the $k \to k+g^\ast $ shift
287: which is the reason for the two-loop behaviour of the flow. 
288: 
289: Let us analyze the flow in detail:
290: \begin{itemize}
291: \item The flow diagram for the compact case is drawn in
292:   Fig.\ref{fig:flow-lines-2l-compact} where we see the presence of
293:   three phases:
294:   \begin{itemize}
295:   \item region {1} is the basin of attraction for the \textsc{wzw} model
296:     ($z = 1$);
297:   \item the points in region {2} describe systems that flow towards
298:     asymptotic freedom;
299:   \item region {3} seems to be the basin of attraction for a different theory,
300:     always with a group manifold geometry but with a
301:     differently-normalized \textsc{wz} term.
302:   \end{itemize}
303:   only a discrete set of trajectories is allowed and, in particular,
304:   region {3} -- separated from region {1} by the line
305:   $\lambda^\ast = 4 \pi \h$ -- is only accessible for levels $k <
306:   g^\ast /2 $.
307: \item The flow diagram for the non\nb-compact case is drawn in Fig...;
308:   again we see three different phases:
309:   \begin{itemize}
310:   \item region {1} describe theories flowing to asymptotic freedom;
311:   \item region {2} looks like the basin of attraction for the non\nb-trivial
312:     solution with the group manifold metric and a new normalization
313:     topological term;
314:   \item region {3} describe theories flowing to a strong coupling
315:     regime.
316:   \end{itemize}
317:   In particular it is interesting to remark that the roles of the
318:   \textsc{uv} and \textsc{ir} are somehow inverted. The
319:   \textsc{wzw} model appears as a \textsc{uv} fixed point and thus an
320:   unstable solution from the point of view of dynamical systems.
321: \end{itemize}
322: 
323: 
324: \begin{figure}
325:   \includegraphics[width=.8\linewidth]{Compact2loops}
326:   \caption{Two-loop \textsc{rg}-flow diagram for compact groups.}
327:   \label{fig:flow-lines-2l-compact}
328: \end{figure}
329: 
330: 
331: \subsection{Renormalization group-flow in squashed compact groups}
332: 
333: The models that we have presented in Ch.~\ref{cha:deformations} are
334: conformal; for this reason we expect to find them as fixed points in
335: an \textsc{rg} flow. To verify this claim let us introduce a
336: two-parameter family of $\sigma$~models generalizing the exact
337: backgrounds; a possible choice consists in adding an extra magnetic
338: field on the top of the one responsible for the squashing, but now
339: coming from a higher-dimensional right sector. Explicitly
340: \begin{equation}
341:   \label{eq:general-deformed}
342:   \begin{cases}
343:     \di s^2 = \displaystyle{\sum_{\mu \in G/T}} J^\mu J^\mu + \left( 1 - \h^2 \right)
344:     \displaystyle{\sum_{a \in T}} J^a J^a ,\\
345:     H_{[3]} = \frac{\hb}{2\h} f_{\mu \nu\rho} J^\mu \land J^\nu \land
346:     J^\rho & \mu \in G/T, \\
347:     F^a = \frac{\h + \hb}{2} \sqrt{\frac{k}{k_g}} \F{a}{\mu \nu} J^\mu \land
348:     J^\nu & \text{$\mu \in G/T$, $a \in T$}, \\
349:     \bar F^a = \frac{\h - \hb}{2} \sqrt{\frac{k}{k_g}} \F{a}{\mu \nu} J^\mu \land J^\nu & \text{$\mu \in G/T$, $a \in T$}
350:   \end{cases}
351: \end{equation}
352: and in particular for $SU(2)$:
353: \begin{equation}
354: \label{eq:2par-su2}
355:   \begin{cases}
356:     \di s^2 = \di \theta^2 + \di \psi^2 + \di \phi^2
357: + \cos \theta \di \psi \di \phi - \h^2 \left( \di \psi + \cos \theta \di \phi \right)^2\ ,\\
358:     B = \frac{\hb}{\h} \cos \theta \di \psi \land \di \phi\ ,\\
359:     A = \left( \h + \hb \right) \left( \di \psi + \cos \theta \di \phi \right)\ ,\\
360:     \bar A = \left( \h - \hb \right) \left( \di \psi + \cos \theta \di \phi \right)\ ,
361:   \end{cases}
362: \end{equation}
363: where $\hb $ is a new parameter, describing the deviation from the
364: conformal point. It is clear that the above background reduces to the
365: one we are used to in the $\hb \to \h $ limit.  In particular we see
366: that the metric is unchanged, the Kalb--Ramond field has a different
367: normalization and a new field $\bar A$ appears.  This configuration
368: can be described in a different way: the geometry of a squashed sphere
369: supports two covariantly constant magnetic fields with charge $Q = \h
370: + \hb$ and $\bar Q =\h - \hb$; the \textsc{rg} flow will describe the
371: evolution of these two charges from a generic $\left( Q, \bar Q
372: \right)$ to $\left( 2 \h, 0 \right)$, while the sum $Q + \bar Q = 2
373: \h$ remains constant.  In this sense the phenomenon can be interpreted
374: as a charge transmutation of $\bar Q $ into $Q$.  The conservation of
375: the total charge is in fact a consequence of having chosen a
376: perturbation that keeps the metric and only changes the antisymmetric
377: part of the background.
378: 
379: We can also see the background in Eq.\eqref{eq:general-deformed} from
380: a higher dimensional perspective where only the metric and the
381: Kalb-Ramond field are switched on. Pictorially:
382: \begin{align}
383:   \label{eq:4d-2par-su2}
384:   g = \left(
385:     \begin{tabular}{ccc|c}
386:       & & & \\
387:       & $g_{\textsc{wzw}}$ & & $\h J_a$  \\
388:       & & &  \\ \hline
389:       & $\h J_a$ & & 1
390:     \end{tabular} \right) &&
391:   B = \left(
392:     \begin{tabular}{ccc|c}
393:       & & & \\
394:       & $\frac{\hb}{\h} B_{\textsc{wzw}}$ & & $\hb J_a$  \\
395:       & & &  \\ \hline
396:       & $-\hb J_a$ & & 0
397:     \end{tabular} \right)
398: \end{align}
399: where $g_{\textsc{wzw}}$ and $B_{\textsc{wzw}}$ are the usual metric
400: and Kalb--Ramond fields for the \textsc{wzw} model on the group $G$.
401: More explicitly in the $SU(2)$ case:
402: \begin{align}
403:   g = \begin{pmatrix}
404:     1 & 0 & 0 & 0 \\
405:     0 & 1 & \cos \theta & \h \\
406:     0 & \cos \theta & 1 & \h \cos \theta \\
407:     0 & \h & \h \cos \theta & 1
408:   \end{pmatrix} &&
409:   B = \begin{pmatrix}
410:     0 & 0 & 0 & 0 \\
411:     0 & 0 & \frac{\hb}{\h} \cos \theta & \hb \\
412:     0 & -\frac{\hb}{\h} \cos \theta & 0 & \hb \cos \theta \\
413:     0 & -\hb & -\hb \cos \theta & 0
414:   \end{pmatrix}
415: \end{align}
416: where the fourth entry represents the bosonized internal current.  In
417: particular this clarifies the stated right-sector origin for the new
418: gauge field $\bar A$. This higher dimensional formalism is the one we
419: will use in the following \textsc{rg} analysis.
420: 
421: \bigskip 
422: 
423: The beta-equations at two-loop order in the expansion in powers of the
424: overall coupling constant $\lambda$ and the field redefinitions for
425: the internal coordinates $X^i$ turn out to be:
426: \begin{equation}
427: \label{eq:beta-compact-2loops}
428:   \begin{cases}
429:     \beta_{\lambda^\ast} =
430:  \Deriv{\lambda^\ast}{t} = - \frac{\lambda^{\ast 2}}{4 \pi} \left( 1-
431:       \frac{\hb^2}{\h^2} \right) \left( 1+ \frac{\lambda^\ast}{8 \pi}
432: \left( 1 - 3 \frac{\hb^2}{\h^2} \right) \right), \\
433:     \beta_{\h} = \Deriv{\h}{t} = \frac{\lambda^\ast \h}{8 \pi} \left( 1- \h^2 \right)
434:     \left( 1- \frac{\hb^2}{\h^2} \right)
435: \left( 1+ \frac{\lambda^\ast}{8 \pi} \left( 1 - 3 \frac{\hb^2}{\h^2} \right) \right) , \\
436:     \beta_{\hb} = \Deriv{\hb}{t} = - \frac{\lambda^\ast \hb}{8 \pi} \left( 1 + \h^2
437:     \right) \left( 1- \frac{\hb^2}{\h^2} \right)
438: \left( 1+ \frac{\lambda^\ast}{8 \pi} \left( 1 - 3 \frac{\hb^2}{\h^2} \right) \right) , \\
439:     X^i = X^i - \frac{\lambda^\ast}{16}
440: \left( 1 - \h^2 \right) \left( 1- 4 \frac{\hb^2}{\h^2} + 3 \frac{\hb^4}{\h^4} \right) ,
441:   \end{cases}
442: \end{equation}
443: where $\lambda^\ast = \lambda g^\ast$, $g^\ast$ being the dual Coxeter
444: number, is the effective coupling constant ($\lambda^\ast = N \lambda
445: $ for $G = SU(N)$). The contributions at one- and two-loop order are
446: clearly separated. In the following we will concentrate on the
447: one-loop part and we will comment on the two-loop result later.
448: Let us then consider the system:
449: \begin{equation}
450: \label{eq:compact-beta}
451:   \begin{cases}
452:     \beta_{\lambda^\ast} = \Deriv{\lambda^\ast}{t} = - \frac{\lambda^{\ast 2}}{4 \pi} \left( 1-
453:       \frac{\hb^2}{\h^2} \right) , \\
454:     \beta_{\h} = \Deriv{\h}{t} = \frac{\lambda^\ast \h}{8 \pi} \left( 1- \h^2 \right)
455:     \left( 1- \frac{\hb^2}{\h^2} \right), \\
456:     \beta_{\hb} = \Deriv{\hb}{t} = - \frac{\lambda^\ast \hb}{8 \pi} \left( 1 + \h^2
457:     \right) \left( 1- \frac{\hb^2}{\h^2} \right)\ .
458:   \end{cases}
459: \end{equation}
460: This can be integrated by introducing the parameter $z = \hb / \h$ which makes one
461: of the equations redundant. The other two become:
462: \begin{equation}
463: \label{eq:compact-red-system}
464:   \begin{cases}
465:     \dot \lambda^\ast = - \frac{\lambda^{\ast 2}}{4 \pi} ( 1 - z^2 ) ,\\
466:     \dot z = - \frac{\lambda^\ast z}{4 \pi} ( 1 - z^2 )\ .
467:   \end{cases}
468: \end{equation}
469: By inspection one easily sees that $\dot \lambda / \lambda = \dot z /
470: z $, implying $ \lambda (t) = C z (t)$, where $C$ is a constant.
471: This was to be expected since $C$ is proportional to the normalization
472: of the topological \textsc{wz} term. Since we are dealing with a
473: compact group it turns out that $C$ is, as in \cite{Witten:1983ar},
474: quantized with:
475: \begin{equation}
476:   C_k = \frac{2 \pi}{k},\ \ k \in \setN \ .
477: \end{equation}
478: Now it's immediate to separate the system and find that $z (t)$ is
479: defined as the solution to the implicit equation:
480: \begin{equation}
481:   - \frac{t}{2 k} = \frac{1}{z_0} - \frac{1}{z(t)} + \log \left[ \frac{
482:       \left( z(t) + 1 \right) \left( z_0 - 1 \right) }{\left( z(t) - 1
483:       \right) \left( z_0 + 1 \right)} \right]
484: \end{equation}
485: with the initial condition $z(0) = z_0$. A similar expression was found in
486: \cite{Braaten:1985is,Witten:1983ar}. The reason for this is, as pointed out
487: previously~\cite{Kiritsis:1995iu}, that the conformal model ($\hb = \h$) in its
488: higher-dimensional representation (the one in Eq.~(\ref{eq:4d-2par-su2}))
489: coincides with a $G \times H$ \textsc{wzw} model after a suitable local
490: field redefinition.
491: 
492: As it is usually the case in the study of non-linear dynamics, a
493: better understanding of the solution is obtained by drawing the
494: \textsc{rg} flow. In a $\left( z, \lambda^\ast \right)$ plane, the
495: trajectories are straight lines through the origin and only a discrete
496: set of them are allowed.  Moreover the line $z = 1 $ is an \textsc{ir}
497: fixed-point locus. This situation is sketched in
498: Fig.~\ref{fig:flow-lines-su2}(a). Just as expected the $z = \hb / \h =
499: 1 $ point, corresponding to the initial exact model described in
500: Ch.~\ref{cha:deformations}, is an \textsc{ir} fixed point for the
501: \textsc{rg} flow.
502: 
503: \begin{figure}
504:   \subfigure[$(z, \lambda) $ plane]{\includegraphics[width=7cm]{Compact1loop}}
505:   \subfigure[$(\h, \hb)$ plane]{\includegraphics[width=7cm]{HHbarFlow}}
506:   \caption{Flow lines for the deformed (non-conformal) squashed
507:     \textsc{wzw} model in (a) the $(z, \lambda )$ and (b) the $(\h, \hb)$
508:     planes. The arrows point in the negative $t$ direction, \emph{i.e.}
509:     towards the infrared; in (a) we see how the squashed \textsc{wzw} model
510:     $z=1$ appears as an \textsc{ir} fixed point, in (b) how perturbing the
511:     conformal $\hb = \h $ model by increasing $\hb$ leads to a a new fixed
512:     point corresponding to a value of $\h$ closer to $1$.}
513:   \label{fig:flow-lines-su2}
514: \end{figure}
515: 
516: Further insights can  be gained if
517: we substitute  the condition $\lambda^\ast  = C_k  \hb /  \h$ into the
518: system (\ref{eq:compact-beta}) thus getting:
519: \begin{equation}
520:   \label{eq:compact-HHb}
521:   \begin{cases}
522:     \Deriv{\h}{t} =  \frac{\hb}{4 k} \left( 1- \h^2 \right)
523:     \left( 1 - \frac{\hb^2}{\h^2} \right) , \\
524:     \Deriv{\hb}{t} = - \frac{\hb^2}{4 k \h} \left( 1 + \h^2 \right)
525:     \left( 1 - \frac{\hb^2}{\h^2} \right) .
526:   \end{cases}
527: \end{equation}
528: The flow diagram for this system in the $\left( \h, \hb \right)$
529: plane, Fig.~\ref{fig:flow-lines-su2}(b), shows how the system relaxes
530: to equilibrium after a perturbation. In particular we can see how
531: increasing $\hb$ leads to a a new fixed point corresponding to a value
532: of $\h$ closer to $1$.
533: 
534: We would like to pause for a moment and put the above results in
535: perspective. Consider for simplicity the $SU(2)$ case: the
536: target-space of the sigma-model under consideration is a squashed
537: three-sphere with two different magnetic fields. Along the flow, a
538: transmutation of the two magnetic charges occurs: the system is driven
539: to a point where one of the magnetic charges vanishes. This fixed
540: point is an ordinary squashed-\textsc{wzw} (of the type studied in
541: Ch.~\ref{cha:deformations}), that supports a single magnetic charge.
542: 
543: As we pointed out in Ch.~\ref{cha:deformations}, in the
544: squashed-\textsc{wzw}, the magnetic field is bounded by a critical
545: value, $\h = 1$. As long as $\h \leq 1$, the geometry is a genuine
546: squashed three-sphere. For $\h > 1$, the signature becomes Lorentzian
547: and the geometry exhibits closed time-like curves.  Although of
548: limited physical interest, such a background can be used as a
549: laboratory for investigating the fate of chronological pathologies
550: along the lines described above.  In the case under consideration and
551: under the perturbation we are considering the model shows a symmetry
552: between the $\h >1 $ and $\h<1$ regions. The presence of closed
553: time-like curves doesn't seem to effect the stability (note that
554: regions with different signatures are disconnected, \emph{i.e.}  the
555: signature of the metric is preserved under the \textsc{rg} flow).  It
556: is clear however that these results are preliminary. To get a more
557: reliable picture for the fate of closed time-like curves, one should
558: repeat the above analysis in a wider parameter space, where other
559: \textsc{rg} motions might appear and deliver a more refined stability
560: landscape.
561: 
562: \bigskip
563: 
564: A final remark concerns the fact that we find the same \textsc{rg}
565: flow behaviour as for a compact (non-squashed) group. We have already
566: made extensive use of the fact that formally the squashed $SU(2)$
567: behaves like a $SU(2) \times U(1)$ \textsc{wzw} model, in particular
568: in Sec.~\ref{sec:no-renorm-theor} where this was at the root of the
569: no-renormalization theorem. In some sense, then, the present
570: calculation is just a perturbative confirmation of that statement.
571: 
572: \subsection{Renormalization group-flow in squashed anti de Sitter}
573: \label{sec:non-compact}
574: 
575: As we've already discussed in Sec.~\ref{sec:deformed-sl2}, sigma
576: models based on non-compact group offer richer (\emph{i.e.}  more
577: complex) phase diagrams than the compact ones. In our particular
578: models this is because the possible choices for a Cartan torus are not
579: pairwise conjugated by inner automorphisms and this is why different
580: choices correspond to inequivalent backgrounds, exhibiting different
581: physical properties. If we concentrate our attention on the $SL (2,
582: \setR )$ \textsc{wzw} model (that is the only non-compact case with
583: just one time direction), we see that the three possible choices for
584: the Cartan generator (elliptic, parabolic, hyperbolic) respectively
585: lead to the exact backgrounds we introduced in
586: Sec.~\ref{sec:deformed-sl2} and we report here for convenience:
587: \begin{align}
588:   &\begin{cases}
589:     \di s^2 = \di \rho^2 - \di t^2 + \di \phi^2 - 2 \sinh \rho \di t
590:     \di \phi - \h^2 \left( \di t + \sinh \rho \di \phi \right)^2 ,\\
591:     B = \sinh \rho \di t \land \di \phi ,\\
592:     A = 2 \h \left( \di t + \sinh \rho \di \phi \right) .
593:   \end{cases} \\
594:   &\begin{cases}
595:     \di s^2 = \frac{\di u^2}{u^2} + \frac{\di x^+ \di x^-}{u^2} - \h^2
596:     \frac{\di x^+ \di x^+}{u^4} , \\
597:     B = \frac{\di x^+ \land \di x^-}{u^2} ,\\
598:     A = 2 \h \frac{\di x^+}{u^2}.
599:   \end{cases}\\
600:   &\begin{cases}
601:     \di s^2 = \di r^2 + \di x^2 - \di \tau^2 + 2 \sinh r \di x \di
602:     \tau - \h^2 \left( \di x + \sinh r \di \tau \right)^2 \\
603:     B = \sinh r \di x \land \di \tau ,\\
604:     A = 2 \h \left( \di x + \sinh r \di \tau \right).
605:   \end{cases}
606: \end{align}
607: 
608: Since these solutions are exact \textsc{cft} backgrounds, we expect
609: them to appear as fixed points for an \textsc{rg} flow, like the
610: compact configuration described in the previous section. As we will
611: see in the following this is actually the case, but with a difference
612: regarding the role of the \textsc{uv} and \textsc{ir} which is proper
613: to non\nb-compact groups (as explained in
614: Sec.~\ref{sec:two-loop-equations}). Using the same technique as above,
615: the first step consists in generalizing the three backgrounds by
616: introducing the following three families of low energy configurations:
617: \begin{align}
618:   &\begin{cases} \di s^2 = \di \rho^2 - \di t^2 + \di \phi^2 - 2 \sinh
619:     \rho \di t \di \phi - \h^2 \left( \di t + \sinh \rho \di \phi
620:     \right)^2  \\
621:     B = \frac{\hb}{\h} \sinh \rho \di t \land \di \phi \\
622:     A = \left( \h + \hb \right) \left( \di t + \sinh \rho \di \phi \right) \\
623:     \bar A = \left( \h - \hb \right) \left( \di t + \sinh \rho \di
624:       \phi \right)
625:   \end{cases} \\
626:   &\begin{cases} \di s^2 = \frac{\di u^2}{u^2} + \frac{\di x^+ \di
627:       x^-}{u^2} - \h^2 \frac{\di x^+ \di x^+}{u^4}  \\
628:     B = \frac{\hb}{\h} \frac{\di x^+ \land \di x^-}{u^2} \\
629:     A = \left( \h + \hb \right) \frac{\di x^+}{u^2} \\
630:     \bar A = \left( \h - \hb \right) \frac{\di x^+}{u^2}
631:   \end{cases} \\
632:   &\begin{cases} \di s^2 = \di r^2 + \di x^2 - \di \tau^2 + 2 \sinh r
633:     \di x \di \tau - \h^2 \left( \di x + \sinh r \di \tau \right)^2 \\
634:     B = \frac{\hb}{\h} \sinh r \di x \land \di \tau \\
635:     A = \left( \h + \hb \right) \left( \di x + \sinh r \di \tau \right) \\
636:     \bar A = \left( \h - \hb \right)\left( \di x + \sinh r \di \tau
637:     \right)
638:   \end{cases}
639: \end{align}
640: The guiding principle remains the same, \emph{i.e.} keep the same
641: geometry, rescale the \textsc{kr} field and introduce a new
642: electromagnetic field, coming (in a four-dimensional perspective) from
643: the right\nb-moving sector. Again we will observe the same
644: charge-transmutation effect as before, this time in terms of charge
645: density (or charge at infinity).
646: 
647: The backgrounds above can be equivalently described in four dimensions
648: by a metric and a \textsc{kr} field as follows:
649: \begin{small}
650:   \begin{align}
651:     \hspace{-1cm}g &=
652:     \begin{pmatrix}
653:       1 & 0 & 0 & 0 \\
654:       0 & -1 & -\sinh \rho & \h \\
655:       0 & -\sinh \rho & 1 & \h \sinh \rho \\
656:       0 & \h & \h \sinh \rho & 1
657:     \end{pmatrix} &
658:     B &= \begin{pmatrix}
659:       0 & 0 & 0 & 0 \\
660:       0 & 0 & \frac{\hb}{\h} \sinh \rho & \hb \\
661:       0 & -\frac{\hb}{\h} \sinh \rho & 0 & \hb \sinh \rho \\
662:       0 & -\hb & -\hb \sinh \rho & 0
663:     \end{pmatrix} \\
664:     \hspace{-1cm}g & = \begin{pmatrix}
665:       \frac{1}{u^2} & 0 & 0 & 0 \\
666:       0 & 0 & \frac{1}{2u^2} & \frac{\h}{u^2} \\
667:       0 & -\frac{1}{2u^2} & 0 & 0 \\
668:       0 & \frac{\h}{u^2} & 0 & 1
669:     \end{pmatrix} &
670:     B &= \begin{pmatrix}
671:       0 & 0 & 0 & 0 \\
672:       0 & 0 & \frac{\hb}{\h} \frac{1}{2u^2} & \frac{\hb}{u^2} \\
673:       0 & \frac{\hb}{\h} \frac{1}{2u^2} & 0 & 0 \\
674:       0 & -\frac{\hb}{u^2} & 0 & 0
675:     \end{pmatrix} \\
676:     \hspace{-1cm}g &= \begin{pmatrix}
677:       1 & 0 & 0 & 0 \\
678:       0 & 1 & \sinh r & \h \\
679:       0 & \sinh r & -1 & \h \sinh r \\
680:       0 & \h & \h \sinh r & 1
681:     \end{pmatrix} &
682:     B &= \begin{pmatrix}
683:       0 & 0 & 0 & 0 \\
684:       0 & 0 & \frac{\hb}{\h} \sinh r & \hb \\
685:       0 & -\frac{\hb}{\h} \sinh r & 0 & \hb \sinh r \\
686:       0 & -\hb & -\hb \sinh r & 0
687:     \end{pmatrix}
688:   \end{align}  
689: \end{small}
690: We must now evaluate the $R^-$ tensor (\emph{i.e.}  the Ricci tensor
691: with respect to the connection $\Gamma^- = \Gamma + 1/2 H$) and read
692: the counterterms in a dimensional regularization scheme as described
693: in Eq.~\eqref{eq:bare-stuff}:
694: \begin{small}
695:   \begin{align}
696:     \hspace{-1cm}\begin{cases}
697:       J^{(1)} = \frac{1}{4} \left( 1 - \frac{\hb^2}{\h^2} \right) ,\\
698:       a_{\h}^{(1)} = - \frac{1+\h^2}{8} \h \left( 1 - \frac{\hb^2}{\h^2} \right) ,\\
699:       a_{\hb}^{(1)} = \frac{1 - \h^2}{8} \hb \left( 1 - \frac{\hb^2}{\h^2} \right), \\
700:       X^{(1)}_X = \frac{ 1 + \h^2 }{8} \left( 1 - \frac{\hb^2}{\h^2} \right) X .
701:     \end{cases} %&&
702:     \begin{cases}
703:       J^{(1)} = \frac{1}{4} \left( 1 - \frac{\hb^2}{\h^2} \right) ,\\
704:       a_{\h}^{(1)} = - \frac{1}{8} \h \left( 1 - \frac{\hb^2}{\h^2} \right) ,\\
705:       a_{\hb}^{(1)} = \frac{1}{8} \hb \left( 1 - \frac{\hb^2}{\h^2} \right) ,\\
706:       X^{(1)}_X = \frac{ 1 }{8} \left( 1 - \frac{\hb^2}{\h^2} \right) X .
707:     \end{cases} &&
708:     \begin{cases}
709:       J^{(1)} = \frac{1}{4} \left( 1 - \frac{\hb^2}{\h^2} \right) ,\\
710:       a_{\h}^{(1)} = - \frac{1 - \h^2}{8} \h \left( 1 - \frac{\hb^2}{\h^2} \right) ,\\
711:       a_{\hb}^{(1)} = \frac{1 + \h^2}{8} \hb \left( 1 - \frac{\hb^2}{\h^2} \right) ,\\
712:       X^{(1)}_X = \frac{ 1 - \h^2 }{8} \left( 1 - \frac{\hb^2}{\h^2} \right) X .
713:     \end{cases}
714:   \end{align}
715: \end{small}
716: The analogies among the three configurations are clear, but become
717: striking when we introduce the parameter $z = \hb / \h $ and all 
718: three $\beta$\nb-functions systems all reduce to the following:
719: \begin{equation}
720:   \label{eq:ad3-one-loop}
721:   \begin{cases}
722:     \dot \lambda = \frac{\lambda^2}{4 \pi} ( 1 - z^2 ) ,\\
723:     \dot z = \frac{\lambda z}{4 \pi} ( 1 - z^2 ) .
724:   \end{cases}
725: \end{equation}
726: This is (up to a sign) the same system we found in the compact case
727: and it is hence immediate to write the solution
728: \begin{gather}
729:   \lambda (t) = C z(t) \\
730:   \frac{C t}{4 \pi } = \frac{1}{z_0} - \frac{1}{z (t)} + \log \left[
731:     \frac{ \left( z(t) + 1 \right) \left( z_0 - 1 \right) }{\left( z(t) - 1
732:       \right) \left( z_0 + 1 \right)} \right] .
733: \end{gather}
734: Although, as expected, $z = 1$ is a fixed point (corresponding to the
735: conformal points) some differences are important. First of all the
736: background is non\nb-compact, so $C$ is not quantized and, although
737: the flow trajectories are still straight lines through the origin, the
738: angular parameter is now arbitrary. The other difference is that $z =
739: 1$ is a fixed point, but it doesn't correspond to a \textsc{ir} stable
740: configuration but to a \textsc{uv} stable one. This is precisely the
741: same behaviour that one encounters for non\nb-compact \textsc{wzw}
742: models when varying the normalization of the \textsc{wz} term (as in
743: Sec.~\ref{sec:two-loop-equations}). Again the flow diagram is the same
744: as for the original $SL(2,\setR)$ group and is summarized in
745: Fig.~\ref{fig:flow-lines-sl2}.
746: 
747: 
748: \begin{figure}
749:   \centering
750:   \includegraphics[width=.8\linewidth]{NonCompact1loop}
751:   \caption{Flow diagram for the system in Eq.~(\ref{eq:ad3-one-loop}).
752:     The arrows point from the \textsc{uv} to the \textsc{ir} and $z=1$
753:     appears as a \textsc{uv} stable solution locus.}
754:   \label{fig:flow-lines-sl2}  
755: \end{figure}
756: 
757: 
758: \section{The CFT approach}
759: \label{sec:cft-approach-1}
760: 
761: 
762: In order to make contact with genuine \textsc{cft} techniques, we must
763: identify the relevant operators which are responsible for the $(\h,
764: \hb)$ deformation of the $G \times H$ original \textsc{wzw} model ($H
765: = U(1)^{\rank G}$). At lowest approximation, all we need is their
766: conformal dimensions in the unperturbed theory.
767: 
768: Following~\cite{Zamolodchikov:1986gt}, let $S_0$ be the unperturbed
769: (conformal) action and $ \mathcal{O}_i$ the relevant operators of
770: conformal dimension $\Delta_i = 1 - \epsilon_i$. Consider the
771: perturbed model,
772: \begin{equation}
773:   S = S_0 + g^i \mathcal{O}_i.
774: \end{equation}
775: The tree-level beta-functions read:
776: \begin{equation}
777:   \label{eq:cft-beta}
778:   \beta^i (g) = - \epsilon_i g^i ,
779: \end{equation}
780: where $g^i$ is supposed to be small, for the perturbative
781: expansion of $\beta^i$ to hold\footnote{One should be very careful in the
782:   choice of signs in these formulae. In~\cite{Zamolodchikov:1986gt} the time
783:   variable, in fact, describes the evolution of the system towards the infrared
784:   and as such it is opposite with respect to the $t = \log \mu $ convention
785:   that we used in the previous section (as in \cite{Witten:1983ar}).}.
786: 
787: The $G \times H$ primary operator we need can be written as follows:
788: \begin{equation}
789:   \mathcal{O} = \sum_{\ssc{a,b}} \braket{ t^{\ssc{a}} g t^{\ssc{b}} g^{-1}} \braket{ t^{\ssc{a}} \partial g g^{-1}  } \braket{ t^{\ssc{b}} g^{-1} \bar \partial g} =  \sum_{\ssc{a,b}} \Phi^{\ssc{ab}}  J^{\ssc{a}} \bar J^{\ssc{b}} ,
790: \end{equation}
791: where $\Phi^{\ssc{ab}}$ is a primary field transforming in the adjoint
792: representation of the left and right groups $G$. As such, the total
793: conformal dimensions (as we've seen in Sec.~\ref{sec:cft-approach})
794: are
795: \begin{equation}
796:   \Delta = \bar \Delta = 1 + \frac{g^\ast}{g^\ast + k} ,
797: \end{equation}
798: where $g^\ast$ is the dual Coxeter number and as such the operator is
799: irrelevant (in the infrared).
800: 
801: Specializing this general construction to our case we find that the
802: action for the fields in Eq.~\eqref{eq:KK-action} is:
803: \mathindent=0em
804: \begin{equation}
805:   \label{eq:cft-action}
806:   S = \frac{k}{4 \pi} \left\{ S_0 + \left( \frac{\h}{\hb} - 1\right)  \sum_{\ssc{a,b}} \Phi^{\ssc{ab}} J^{\ssc{a}} \bar J^{\ssc{b}} + \frac{\h}{\hb} \left( \h + \hb \right) \sum_i J^{a_i} \bar J^i +  \frac{\h}{\hb} \left( \h - \hb \right) \sum_{i,\ssc{a}} J^i \Phi^{a_i \ssc{a}} \bar J^{\ssc{a}}  \right\}  .
807: \end{equation}
808: \mathindent=\oldindent
809: where $\ssc{a}$ runs over all currents, $i$ over the internal currents
810: (in $H$) and $J^{a_i}$ is the \textsc{wzw} current of the Cartan
811: subalgebra of $G$ coupled to the internal $\bar J^i$.  The extra terms
812: can be interpreted as relevant combinations of operators in the $G \times
813: H$ model. The beta-functions are thus computed following
814: Eq.~(\ref{eq:cft-beta}):
815: \begin{equation} \label{eq:hhb-beta}
816:   \begin{cases}
817:     \left. \Deriv{}{t} \log \left[ \left( \frac{\h}{\hb} -1 \right) \right]
818:     \right|_{\hb = \h} = \frac{ g^\ast}{g^\ast + k } = \frac{g^\ast}{k } - 
819:     \frac{g^{\ast 2}}{k^2} + \mathcal{O} \left(\frac{1}{k^3}\right), \\
820:     \left. \Deriv{}{t} \log \left[ \frac{\h}{\hb} \left( \h + \hb
821:         \right) \right] \right|_{\hb = \h} =  \mathcal{O} \left(\frac{1}{k^3}\right) , \\
822:     \left. \Deriv{}{t} \log \left[ \frac{\h}{\hb} \left( \h - \hb \right)
823:       \right] \right|_{\hb = \h}=  \frac{ g^\ast}{g^\ast + k } =
824:     \frac{g^\ast}{k } -  \frac{g^{\ast 2}}{k^2} + \mathcal{O}
825:     \left(\frac{1}{k^3}\right).
826:   \end{cases}
827: \end{equation}
828: 
829: Equations~(\ref{eq:hhb-beta}) agree with the results of the
830: field-theoretical approach (up to the overall normalization), at least
831: in the regime where~(\ref{eq:hhb-beta}) are valid, namely for small
832: $\h$ and $\hb$ perturbations. But there's more: as pointed out before
833: the conformal model ($\hb = \h$) is exact because it coincides with a
834: $G \times H$ \textsc{wzw} model after a suitable field redefinition
835: for any value of $\h$.  As a consequence the equations remain valid
836: for any finite $\h$. This is reassuring both for the validity of the
837: geometrical approach\footnote{There is no doubt on the method
838:   itself. It could simply fail to describe the desired phenomenon due
839:   to an inappropriate ansatz for the off-criticality excursion in
840:   parameter space.} and for the conclusions on the stability picture
841: of the models under consideration.
842: 
843: The extra information that we obtain from this calculation is about
844: the interpretation for the two-loop $\beta$-function we described
845: in the previous section. In fact it is now clear that with the
846: target-space approach we just describe the Taylor expansion of the
847: tree-level \textsc{cft} result:
848: \begin{equation}
849:   \frac{g^\ast}{g^\ast + k } = \frac{g^\ast}{k} - \frac{g^{\ast 2}}{k^2} + \mathcal{O} \left(\frac{1}{k^3} \right).
850: \end{equation}
851: This is not surprising since the would-be non-trivial fixed point of
852: the two-loop expansion lay out of the validity range for our
853: approximation. If we really want to go beyond the large $k$ limit, we
854: need to push the analysis from this, \textsc{cft}, side. 
855: 
856: \begin{comment}
857:   The one-loop corrections to~(\ref{eq:cft-beta}) are then of the form
858:   $ C_{ijk} \, g^i \, g^j$, where $ C_{ijk}$ are related to the
859:   three-point function of the unperturbed theory. This coefficient is
860:   a measure of the dimension of the operator $\mathcal{O}_i$ in the
861:   theory perturbed by the set $\{\mathcal{O}_i \}$. Such a computation
862:   requires techniques that go beyond the scope of the present work.
863: 
864: 
865:   \cnote{pezzo aggiunto} 
866:   \begin{equation} \label{eq:hhb-beta}
867:     % CFT:\hskip 1 cm
868:     \Deriv{}{t}  \left( \frac{\h}{\hb} -1 \right)
869:     \Bigg|_{\hb = \h} = \frac{ g^\ast}{g^\ast + k }\left({h\over\hb}-1\right) + \cdots  =
870:     \left(\frac{g^\ast}{k } - \frac{g^{\ast 2}}{k^2} \right)
871:     \left({h\over\hb}-1\right) + \cdots\ ,
872:   \end{equation}
873:   where the dots after the first equality denote higher order terms in
874:   the $(h/\hb-1)$-expansion and after the second equality, in addition
875:   to that, higher order terms in the $1/k$-expansion.  This result is
876:   the same as the one in \cite{Knizhnik:1984nr} since, as we have
877:   mentioned, there is the a local field redefinition that maps this
878:   model at the conformal point to the $G\times H$ \textsc{wzw} model.
879: \end{comment}
880: 
881: \marginlabel{Target space \emph{vs} \textsc{cft} renormalization}From
882: the target space view point, the renormalization approach remains
883: valid in the large $k$ limit for any value of $\h / \hb$. This enables
884: us to use Eq.~\eqref{eq:beta-compact-2loops} and push (for $k \to
885: \infty$) Eq.~\eqref{eq:hhb-beta} at least to the next leading order in
886: $\left( 1 - \h / \hb \right)$ so to get \mathindent=0em
887: \begin{equation} 
888:   \label{eq:hhb-beta1}
889:   \left. \Deriv{}{t} \left( \frac{\h}{\hb} -1 \right)
890:   \right|_{\hb = \h} = \left(\frac{g^\ast}{k } -
891:     \frac{g^{\ast 2}}{k^2}\right)\left(\frac{\h }{\hb } - 1 \right)+
892:   \frac{1}{2} \left(- \frac{g^\ast }{k} + 7 \frac{{g^\ast}^2}{k^2} \right)\left(\frac{\h}{\hb} - 1\right)^2
893:   + \ldots 
894: \end{equation}
895: \mathindent=\oldindent
896: that obviously agrees to first order in the coupling $\left(\h / \hb-1
897: \right)$ with the expression above.
898: 
899: The extra information that we obtain from this calculation is about
900: the interpretation for the two-loop beta-function we described in the
901: previous section. The one-loop corrections to~(\ref{eq:cft-beta}) are
902: of the form $ C_{ijk} \, g^i \, g^j$, where $ C_{ijk}$ are related to
903: the three-point function of the unperturbed theory
904: \cite{Zamolodchikov:1986gt}.  This coefficient is a measure of the
905: dimension of the operator $\mathcal{O}_i$ in the theory perturbed by
906: the set of all operators. Eq.(\ref{eq:hhb-beta1}), based on the
907: target-space approach, precisely predicts the coefficient of the term
908: $\left(\h / \hb - 1 \right)^2$ to second order in the
909: $1/k$-expansion. It seems that such a computation is feasible from the
910: \textsc{cft} viewpoint at least as a series expansion for large
911: $k$. This would allow for a genuine two-loop comparison of the two
912: methods, and is left for future investigation.
913: 
914: 
915: 
916: \section{RG flow and friction}
917: \label{sec:rg-flow-friction}
918: 
919: It has already been noted in literature \cite{Gutperle:2002ki} that a
920: deep link exists between the equations of motion and the
921: \textsc{rg}-flow. In an oversimplified toy model one can consider the
922: equations of motion for a system with friction:
923: \begin{equation}
924: \label{eq:point-dynamics}
925:   \frac{\di^2 r}{\di t^2} = - V^\prime (r) - k \frac{\di r}{\di t} .
926: \end{equation}
927: Large friction corresponds to the $k\to \infty $ limit where the
928: dynamics described by this second order equation is well approximated
929: by a first order one:
930: \begin{equation}
931: \label{eq:large-friction}
932:   \frac{\di r}{\di t^\prime } = - V^\prime (r), \hspace{2cm} t^\prime = \frac{t}{k}.
933: \end{equation}
934: At least in some cases the same link exists between the second order
935: equations of motion and the first order \textsc{rg} flow equation: the
936: latter provide a good approximation for the dynamics of the system in
937: some region of the moduli space. In this section we will provide a
938: class of systems (with constant-curvature metrics and no dilaton)
939: where this can be explicitly verified and the ``friction'' identified
940: with the expectation value for the dilaton which appears out of
941: equilibrium.  More precisely we will consider the \textsc{rg} flow for
942: the coupling constant of the metric with respect to the Kalb-Ramond
943: field and then show that the equations that one obtains in this way
944: are an approximation of those for a system in which the constant is a
945: field depending on an extra time direction.
946: 
947: 
948: \subsection{The \textsc{rg}-flow approach}
949: \label{sec:textscrg-flow-appr}
950: 
951: As announced above we would like to study the \textsc{rg}-flow for the
952: coupling of the metric in a system without dilaton, that is for the
953: sigma model
954: \begin{equation}
955: \label{eq:low-dimensional-sigma}
956:   S = \frac{1}{2 \lambda } \int \di^2 z \:  \left( c g_{\mu \nu} + B_{\mu \nu}  \right) \partial X^\mu \bar \partial X^\nu 
957: \end{equation}
958: knowing that for $c = 1$ the model is conformal. Using the geometric
959: \textsc{rg}-flow approach developed in
960: Sec.~\ref{sec:target-space-renormalization} we find that Riemann
961: tensor with respect to the connection of
962: Eq.~\eqref{eq:minus-connection}:
963: \begin{equation}
964:   {R^-}\ud{\mu}{\nu \rho \sigma} = \left( 1 - \frac{1}{c^2} \right) R\ud{\mu}{ \nu \rho \sigma} .
965: \end{equation}
966: It follows that the one-loop counterterm is given by
967: \begin{equation}
968:   T_{\mu \nu} = \frac{R}{d} \left( 1 - \frac{1 }{c^2} \right) g_{\mu \nu}  
969: \end{equation}
970: where for simplicity we supposed the manifold to be Einstein, which is
971: consistent with the fact that the conformal model with fields $g$ and
972: $B$ doesn't include a dilaton. Hence we immediately find the
973: parameters
974: \begin{align}
975:   J(c) = 0 && a (c) = \frac{R}{d} \left(1 - \frac{1}{c^2} \right)
976: \end{align}
977: and the corresponding beta equations
978: \begin{equation}
979:   \begin{cases}
980:     \beta_{\lambda} = 0 ,\\
981:     \beta_{c} = \frac{\lambda }{\pi } a_c = \frac{\lambda R }{d \pi } \left(1 -
982:       \frac{1}{c^2} \right) .
983:   \end{cases}
984: \end{equation}
985: 
986: In order to compare this result with what we will find in the
987: following we can write
988: \begin{equation}
989:   c (\mu) = e^{2\sigma (\mu)}  
990: \end{equation}
991: where $\mu $ is the energy scale. Then the energy evolution of $\sigma
992: (\mu)$ (going towards the infrared) gives:
993: \begin{equation}
994:   \label{eq:energy-evol-sigma}
995:   \frac{\di \sigma }{\di \mu } = -\frac{\lambda R }{2 d \pi } e^{-2\sigma (\mu)}\left(1 - e^{-4\sigma (\mu)}  \right) = - V^\prime (\sigma(\mu)) 
996: \end{equation}
997: which admits the implicit solution
998: \begin{equation}
999:   \log \mu = - \frac{1}{4} \left( 2 e^{2\sigma(\mu)} + \log (\tanh \sigma(\mu)) \right) .
1000: \end{equation}
1001: 
1002: This is for us the equivalent of Eq.~\eqref{eq:large-friction}. Now we
1003: move to the $\left( d+1 \right)$-dimensional spacetime to find the
1004: corresponding Eq.~\eqref{eq:point-dynamics}.
1005: 
1006: \subsection{Spacetime interpretation}
1007: \label{sec:equations-motion}
1008: 
1009: 
1010: 
1011: \paragraph{Equations of motion.}
1012: 
1013: As we said above we want to describe the same system by introducing an
1014: extra time dimension and reading the coupling as a time-dependent
1015: field. In other words we would like to write the equations of motion
1016: for the following sigma model:
1017: \begin{equation}
1018:   S = \int \di^2 z \: \left[ - \partial t \bar \partial t + \left( c(t) g_{\mu \nu} + B_{\mu \nu}  \right) \partial X^\mu \bar \partial X^\nu \right] 
1019: \end{equation}
1020: where, $g$ and $B$ are background fields solving the low-energy string
1021: equations of motion. In order to write the equations of motion let us
1022: rewrite the $d+1$ dimensional metric in terms of a Weyl rescaling as:
1023: \begin{equation}
1024: \label{eq:g-bar}
1025:   \bar g_{\textsc{mn}} = e^{2 \sigma(t)}
1026:   \begin{pmatrix}
1027:     - e^{-2\sigma(t)} & 0 \\
1028:     0 & g_{\mu \nu}
1029:   \end{pmatrix} = e^{2 \sigma (t)} g_{\ssc{mn}}
1030: \end{equation}
1031: where $c(t) = e^{2 \sigma (t)}$. This means in particular that the
1032: Ricci tensor (this time with respect to the standard Levi-Civita
1033: connection) can be written as
1034: \begin{equation}
1035:   \overline{Ric}_{\ssc{mn}} = Ric_{\ssc{mn}} - g_{\ssc{mn}} K\du{\ssc{l}}{\ssc{l}} - \left(d - 1 \right) K_{\ssc{mn}}  
1036: \end{equation}
1037: where $K_{\ssc{mn}}$ is defined as
1038: \begin{align}
1039:   K\du{\ssc{m}}{\ssc{n}} &= - \partial_{\ssc{m}} \sigma
1040:   g^{\ssc{nl}} \partial_{\ssc{l}} \sigma + g^{\ssc{nl}}
1041:   \left( \partial_{\ssc{m}} \partial_{\ssc{l}} \sigma -
1042:     \Gamma\ud{\ssc{p}}{\ssc{ml}} \partial_{\ssc{p}} \sigma \right) +
1043:   \frac{1}{2} g^{\ssc{lp}} \partial_{\ssc{l}}
1044:   \sigma \partial_{\ssc{p}} \sigma
1045:   \delta_{\ssc{m}}^{\phantom{\ssc{m}}\ssc{n}} \\
1046:   K_{\ssc{mn}} &= g_{\ssc{nl}} B\du{\ssc{m}}{\ssc{l}}
1047: \end{align}
1048: After some algebra one finds that
1049: \begin{align}
1050:   \Gamma\ud{t}{tt} = - \dot \sigma(t) && \Gamma\ud{t}{t\mu} = 0 && \Gamma\ud{t}{\mu \nu} = 0 
1051: \end{align}
1052: \begin{subequations}
1053:   \begin{align}
1054:     K\du{t}{t} = - e^{2 \sigma(t)} \left( \ddot \sigma(t) + \frac{\dot
1055:         \sigma^2(t)}{2} \right) &&
1056:     K\du{\mu}{\nu} = -\frac{1}{2} e^{2\sigma(t)} \dot \sigma^2(t) \delta\du{\mu}{\nu} \\
1057:     K_{tt} = \left(\ddot \sigma(t) + \frac{\dot \sigma^2(t)}{2} \right) && K_{\mu
1058:       \nu} = - \frac{e^{2 \sigma(t) }}{2} \dot \sigma^2(t) g_{\mu \nu}
1059:   \end{align}
1060: \end{subequations}
1061: %&& B\du{t}{\mu} = 0
1062: where $\dot \sigma(t) $ is the notation for
1063: \begin{equation}
1064:   \dot \sigma(t) = \frac{\di \sigma (t)}{\di t}  
1065: \end{equation}
1066: In particular this implies that
1067: \begin{equation}
1068:   K\du{\ssc{l}}{\ssc{l}} = - e^{2\sigma(t)} \left( \frac{d+1}{2} \dot \sigma^2(t) + \ddot \sigma(t)  \right)  .
1069: \end{equation}
1070: It then easily follows that
1071: \begin{subequations}
1072: \label{eq:Ric-bar}
1073:   \begin{align}
1074:     \overline{Ric}_{tt} &= -d \left( \ddot \sigma (t)+ \dot \sigma^2(t) \right) \\
1075:     \overline{Ric}_{t\mu } &= 0 \\
1076:     \overline{Ric}_{\mu \nu } &= Ric_{\mu \nu } + \bar g_{\mu \nu} \left( d \dot
1077:       \sigma^2(t) + \ddot \sigma(t) \right) .
1078:   \end{align}
1079: \end{subequations}
1080: The other terms in the equations of motion read
1081: \begin{align}
1082:   \bar H_{\mu \nu}^2 &= H_{\mu \alpha \beta } H_{\nu \gamma \delta} \bar g^{\alpha \gamma} \bar g^{\beta \delta} = e^{-4 \sigma(t)} H_{\mu \nu}^2 \\
1083:   \bar \nabla_{\ssc{m}} \bar \nabla_{\ssc{n}} \Phi &= \partial_{\ssc{m}} \partial_{\ssc{n}} \Phi -
1084:   \bar \Gamma\ud{\lambda}{\ssc{mn}} \partial_\lambda \Phi
1085: \end{align}
1086: now, $\bar \Gamma\ud{t}{\mu \nu } = - \dot \sigma(t) \bar g_{\mu \nu}$ so
1087: \begin{subequations}
1088:   \begin{align}
1089:     \bar \nabla_t \bar \nabla_t \Phi  &= \ddot \Phi(t) \\
1090:     \bar \nabla_\mu  \bar \nabla_t \Phi  &= 0 \\
1091:     \bar \nabla_\mu \bar \nabla_\nu \Phi &= \dot \sigma(t) \dot
1092:     \Phi(t) \bar g_{\mu \nu} .
1093:   \end{align}
1094: \end{subequations}
1095: These are all the ingredients we need to write the equations of motion:
1096: \begin{equation}
1097:   \overline{Ric}_{\ssc{mn}} - \frac{1}{4} \bar H_{\ssc{mn}}^2 + 2  \bar \nabla_{\ssc{m}} \bar \nabla_{\ssc{n}} \Phi = 0  .
1098: \end{equation}
1099: Splitting the time component we obtain
1100: \begin{subequations}
1101:   \begin{align}
1102:     Ric_{tt} &+ 2 \partial_t \partial_t \Phi(t) = -d \left( \ddot \sigma(t) + \dot \sigma^2(t) \right) + 2 \ddot \Phi(t) = 0 \\
1103:     Ric_{\mu \nu} &\left( 1 - e^{-4\sigma(t)} \right) + \bar g_{\mu
1104:       \nu} \left( d \dot \sigma^2(t) + \ddot \sigma(t) - 2 \dot
1105:       \sigma(t) \dot \Phi(t) \right) = 0
1106:   \end{align}
1107: \end{subequations}
1108: where we have used the equations of motion for the system in $\sigma = 0$:
1109: \begin{equation}
1110:   Ric_{\mu \nu} = \frac{1}{4} H_{\mu\nu}^2 
1111: \end{equation}
1112: The system admits a solution if and only if $g_{\mu \nu}$ is Einstein
1113: (since the original system didn't have any dilaton). Taking the trace
1114: with $\bar g^{\ssc{mn}}$ we obtain the system:
1115: \begin{equation}
1116:   \begin{cases}
1117:     d \left( \ddot \sigma(t)  + \dot \sigma^2(t) \right) - 2 \ddot \Phi(t) = 0 \\
1118:     R e^{-2\sigma(t)} \left( 1 - e^{-4\sigma(t)} \right) + d \left( d \dot
1119:       \sigma^2(t) + \ddot \sigma(t) - 2 \dot \sigma(t) \dot \Phi(t) \right) = 0
1120:   \end{cases}
1121: \end{equation}
1122: Introducing
1123: \begin{equation}
1124:   \label{eq:defin-Psi}
1125:   Q (t) = -\dot \Phi (t) + \frac{d}{2} \dot \sigma (t)
1126: \end{equation}
1127: the equations become:
1128: \begin{equation}
1129: \label{eq:friction-syst}
1130:   \begin{cases}
1131:     \dot Q(t) = - d \dot \sigma^2(t) \\
1132:     \ddot \sigma(t) = - \frac{R}{d} e^{-2\sigma(t)} \left( 1 - e^{-4 \sigma(t)} \right) - 2
1133:     \dot \sigma(t) Q(t)
1134:   \end{cases}
1135: \end{equation}
1136: This second equation has precisely the structure of the motion in a potential
1137: \begin{equation}
1138:   V (\sigma) - V(0) = \frac{R}{6d} e^{-6 \sigma} \left( 1 - 3 e^{4\sigma} \right) \sim - \frac{R}{3d} +  \frac{2 R }{d} \sigma^2 .
1139: \end{equation}
1140: and with a time-dependent friction coefficient $Q(t)$. In the limit of
1141: $Q \to \infty $ we clearly recover Eq.~\eqref{eq:energy-evol-sigma}
1142: with the same potential $V(\sigma)$ when identifying the \emph{energy
1143:   scale} $\mu$ for the off-shell system with the \emph{time direction}
1144: here following
1145: \begin{equation}
1146:   \log \mu = \frac{\pi \bar Q}{\lambda} t .  
1147: \end{equation}
1148: 
1149: 
1150: \paragraph{Linearization.}
1151: 
1152: The system~\eqref{eq:friction-syst} can be solved numerically and
1153: typical results for large $Q (0)$ and small $Q(0)$ are presented in
1154: Fig.~\ref{fig:num-solution-Psi}.
1155: 
1156: \begin{figure}
1157:   \begin{center}
1158:     \subfigure[Small $Q(0)$]{
1159:       \includegraphics[width=.4\linewidth]{SmallPhi-crop}}
1160:     \subfigure[Large $Q(0)$]{
1161:       \includegraphics[width=.4\linewidth]{LargePhi-crop}}
1162:   \end{center}
1163:   \caption{Typical behaviour for $\sigma(t)$ in the
1164:     system~\eqref{eq:friction-syst} for (a)~small and (b)~large
1165:     (positive) initial values of $Q(t)$.}
1166:   \label{fig:num-solution-Psi}
1167: \end{figure}
1168: 
1169: A further step can be made by linearization. Introduce
1170: \begin{equation}
1171:   \Sigma (t)= \dot \sigma (t)   
1172: \end{equation}
1173: the system becomes a first order one:
1174: \begin{equation}
1175:   \begin{cases}
1176:     \dot  Q(t) = - d \Sigma^2(t) \\
1177:     \dot \sigma(t) = \Sigma(t) \\
1178:     \dot \Sigma(t) = - V^\prime (\sigma(t)) - 2 \Sigma(t) Q(t)
1179:   \end{cases}
1180: \end{equation}
1181: which has a fixed point for $\left( Q, \sigma, \Sigma \right) = \left(
1182:   \bar Q, 0, 0 \right)$ where $\bar Q$ is a constant. Around this
1183: point the equations read:
1184: \begin{equation}
1185: \label{eq:linear-system}
1186:   \begin{cases}
1187:     \dot Q(t) = 0 \\
1188:     \dot \sigma(t) = \Sigma(t) \\
1189:     \dot \Sigma(t) = - V^{\prime \prime} (0) \sigma(t) - 2 \bar Q \Sigma(t) = - \frac{4 R}{d}
1190:     \sigma(t) - 2 \bar Q \Sigma(t)
1191:   \end{cases}
1192: \end{equation}
1193: so, $Q $ decouples (and remains constant) and the equation of motion
1194: around the fixed point is
1195: \begin{equation}
1196:   \frac{\di^2 \sigma (t)}{\di t^2} = - \frac{4 R}{d}
1197:   \sigma(t) - 2 \bar Q \frac{\di \sigma (t)}{\di t} ,
1198: \end{equation}
1199: which can be integrated giving
1200: \begin{equation}
1201:   \sigma (t) = C_1 \exp \left[ -\left(  \bar Q + \sqrt{\bar Q^2 -
1202:         \frac{4R}{d}}  \right) t \right] + C_2 \exp 
1203:   \left[ -\left( \bar Q - \sqrt{\bar  Q^2 - \frac{4R}{d}}  \right)
1204:     t \right]
1205: \end{equation}
1206: with $C_1 $ and $C_2$ integration constants.
1207: 
1208: 
1209: For positive $\bar Q$ the solution converges to $\sigma = 0$ with or
1210: without oscillations if $\bar Q^2 \lessgtr 4R/d$.  In terms of $\sigma(t) $
1211: and $\Phi(t)$ this limit solution is
1212: \begin{align}
1213:   \sigma(t) \xrightarrow[t\to \infty ]{} 0 && \Phi (t) \sim - \bar Q t ,
1214: \end{align}
1215: which is not surprisingly the initial conformal model in
1216: Eq.~\eqref{eq:low-dimensional-sigma} plus a linear dilaton.
1217: 
1218: 
1219: \paragraph{The meaning of $\bar Q$.}
1220: 
1221: 
1222: $\bar Q $ is linked to the dilaton: larger values correspond to
1223: negative and larger absolute values for $\Phi$, \emph{i.e.} moving
1224: further inside the perturbative regime.  On the other hand, negative
1225: values of $\bar Q $ give diverging solutions, but in this case the
1226: dilaton grows (see Eq.~\ref{eq:defin-Psi}) and the very underlying
1227: perturbative approach collapses. It is worth to remark that if we make
1228: an hypothesis of uniqueness for the system~\eqref{eq:friction-syst},
1229: $Q$ can't change sign because $Q(t) = 0$, $\sigma(t) = 1$ is a solution
1230: (the starting conformal model with constant dilaton).
1231: 
1232: 
1233: A better understanding of the actual meaning of this parameter can be
1234: obtained if we consider the limiting situation of linear dilaton. In
1235: this case, in fact, it is immediate to derive the central-charge of
1236: the overall system:
1237: \begin{equation}
1238:   c = \left( d + 1 \right) - 3 \bar Q^2 - c_d + c_I = 0 ,
1239: \end{equation}
1240: where $c_d$ is the central-charge of the conformal system in
1241: Eq.~\eqref{eq:low-dimensional-sigma} (\emph{e.g.} $6 / \left(k+2
1242: \right)$ for the $SU(2)$ \textsc{wzw} model) and $c_I$ is the internal
1243: central-charge. If follows that for a critical model
1244: \begin{equation}
1245:   \bar Q^2 = \frac{1}{3} \left( d + 1 - c_d + c_I \right)  
1246: \end{equation}
1247: and $\bar Q$ is essentially a measure of the deficit.
1248: 
1249: 
1250: A final remark regards the consistency of the approximation for the
1251: dynamics one obtains from the \textsc{rg}-flow
1252: equation~\eqref{eq:energy-evol-sigma}, corresponding to a $Q \to \infty$
1253: limit. The linearized system~\eqref{eq:linear-system} provides a
1254: justification for such limit: in fact the time scale for $Q (t)$ is
1255: comparably larger than $\sigma(t)$'s -- to the point that the former
1256: decouples around the fixed point. For this reason it can be taken as a
1257: constant (fixed by the initial conditions) if we just concentrate on
1258: the evolution of the warping factor $\sigma (t)$.
1259: 
1260: \section{Cosmological interpretation}
1261: \label{sec:friedm-roberts-walk}
1262: 
1263: The type of backgrounds we are studying are time-dependent and as such
1264: can have a cosmological interest. For this reason, since there is a
1265: non-trivial dilaton, one should better move to the Einstein frame (as
1266: opposed to the string frame we've been using thus far). This means
1267: that the metric is written as:
1268: \begin{equation}
1269:   \tilde g_{\ssc{mn}} = e^{- \Phi (t) /2 } \bar g_{\ssc{mn}}
1270: \end{equation}
1271: and after a coordinate change
1272: \begin{equation}
1273:   \tau(t) = \int  e^{-\Phi(t)/4} \di t
1274: \end{equation}
1275: can be put back to the same warped product form as in
1276: Eq.~\eqref{eq:g-bar}:
1277: \begin{multline}
1278:   \label{eq:Einst-metric}
1279:   \widetilde{\di s^2} = \tilde g_{\ssc{mn}} \di x^{\ssc{m}} \di
1280:   x^{\ssc{n}} = -\di \tau^2 + \left. e^{2 \sigma(t) - \Phi (t)/2}
1281:   \right|_{t = t(\tau)} \left( g_{\mu\nu} \di x^\mu \di x^\nu \right)
1282:   = \\ = -\di \tau^2 + w(\tau) \left( g_{\mu\nu} \di x^\mu \di x^\nu
1283:   \right).
1284: \end{multline}
1285: 
1286: Cosmologically interesting solutions are obtained when $d=3$. In this
1287: case the $H$ field is proportional to the volume form on $g$. This
1288: implies that $H_{\mu \nu}^2 \propto g_{\mu \nu}$ and then the
1289: equations reduce to
1290: \begin{equation}
1291:   Ric_{\mu \nu} = \Lambda^2 g_{\mu \nu}
1292: \end{equation}
1293: \emph{ie} $g_{\mu \nu}$ is to be the metric of an Einstein
1294: three-manifold (the most simple case being a three-sphere). What we
1295: get then is a typical example of Friedmann-Robertson-Walker
1296: (\textsc{frw}) spacetime such as those already studied in
1297: \cite{Tseytlin:1991ss,Tseytlin:1992ye,Goldwirth:1993ha,Copeland:1994vi}.
1298: As such it describes the time evolution of an isotropic spacetime (or
1299: more in general of a spacetime with the symmetries of the conformal
1300: theory in Eq.~\eqref{eq:low-dimensional-sigma}). Some intuition about
1301: the time evolution can be developed if we take the linearized system
1302: in Eq.~\eqref{eq:linear-system} and consider the large $t$ limit. In
1303: fact, as remarked above the solution asymptotically approaches a
1304: linear dilaton background (which was already studied from this point
1305: of view in \cite{Antoniadis:1988vi}):
1306: \begin{align}
1307:   \sigma (t) \xrightarrow[t\to \infty ]{} 0 && Q (t) = \bar Q && \Phi(t) \sim - \bar Q t
1308: \end{align}
1309: hence one verifies that the metric in the Einstein frame is
1310: asymptotically
1311: \begin{equation}
1312:   \widetilde{\di s^2} \sim - \di \tau^2 + \bar Q^2 \tau^2 \left( g_{\mu\nu} \di x^\mu \di x^\nu \right) 
1313: \end{equation}
1314: which corresponds to an expanding universe with curvature
1315: \begin{equation}
1316:   \tilde R \sim \frac{R + \bar Q^2 d \left( d - 1 \right)}{\bar Q^2 \tau^2} .%=  \frac{6 \left( 1 + \bar Q^2 \right)}{\tau^2 \bar Q^2}
1317: \end{equation}
1318: 
1319: A similar result, with a polynomial expansion is found if we consider
1320: an exponential decrease for $\sigma(t)$, or better for $c(t)$ (in the
1321: linear limit $c(t) - 1$ obeys the same equations as
1322: $\sigma(t)$). After a redefinition of the variables we can let
1323: \begin{equation}
1324:   c(t) = e^{-t} + 1  .
1325: \end{equation}
1326: It is easy to check that in general\footnote{On a side note, since $c(t)>0$ by construction the relation $\tau = \tau (t)$ is always invertible.}
1327: \begin{equation}
1328:   \tau (t) = \int c(t)^{-d/16} e^{1/4\int Q(t^\prime) \di t^\prime} \di t
1329: \end{equation}
1330: and in this linearized approximation the latter becomes
1331: \begin{equation}
1332:   \tau (t)= \int \left( e^{-t} + 1 \right)^{-d/16} e^{\bar Q t / 4} \di t .
1333: \end{equation}
1334: This integral can be solved analytically: \mathindent=0em
1335: \begin{equation}
1336:   \tau (u) = \frac{16}{d + 4 \bar Q} \left( 1 + \frac{1}{u} \right)^{-d/16} u^{\bar Q/4} \left( 1 + u \right)^{d/16} {}_2F_1 \left( \frac{d}{16} , \frac{d+4 \bar Q}{16}; \frac{d + 4 \bar Q}{16} + 1, - u\right) ,  
1337: \end{equation}
1338: \mathindent=\oldindent
1339: where $u = e^t$ and ${}_2F_1$ is an hypergeometric
1340: function\footnote{The hypergeometric function ${}_2F_1$ is defined as
1341:   follows:
1342:   \begin{equation}
1343:     {}_2F_1 (a,b;c,u) = \sum_{k=0}^\infty \frac{ \left( a \right)_k \left( b \right)_k}{\left(c \right)_k} \frac{z^k}{k!}    
1344:   \end{equation}
1345:   where $\left(a \right)_k$ is the Pochhammer symbol
1346:   \begin{equation}
1347:     \left( a \right)_k = \frac{\Gamma (a+k)}{\Gamma (a)}
1348:   \end{equation}%
1349: }. It is better however to consider the asymptotic behaviours. For $u
1350: \to \infty$ one finds that $\tau (u)$ and the warping factor $w(u)$ go
1351: like:
1352: \begin{align}
1353:   \tau (u) \sim \frac{4}{\bar Q} u^{\bar Q /4} , && w(u) \sim u^{\bar Q /2} ,
1354: \end{align}
1355: and consistently with the results above for the linear dilaton case (which is
1356: precisely the large-$u$ limit):
1357: \begin{equation}
1358:   w (\tau) \sim \tau^2 ;
1359: \end{equation}
1360: similarly for small $u$:
1361: \begin{align}
1362:   \tau (u) \sim \frac{16}{d+4 \bar Q} u^{\left( d+4 \bar Q\right)/16} , && w(u) \sim u^{d/4+2+\bar Q/2}
1363: \end{align}
1364: and then
1365: \begin{equation}
1366:   w(\tau) \sim \tau^{4 + 8 \left( 4 - \bar Q \right)/\left( d + 4 \bar Q \right)} .
1367: \end{equation}
1368: Note that this behaviour precisely measures the effect of a finite
1369: value for $\bar Q$ and in fact for $\bar Q \to \infty$ we recover
1370: again $w (\tau) \sim \tau^2$.  Summarizing, just as advertised, we get
1371: again a polynomially expanding universe (a so-called big-bang
1372: solution).
1373: 
1374: The analysis for the small-$\bar Q$ regime is clearly more difficult
1375: to be carried out analytically. Apart from numerical solutions (see
1376: Fig.~\ref{fig:num-solution-warp}), in general we can study $w(\tau)$
1377: as a parametric curve in the $\left(w,\tau\right)$ plane defined by
1378: $\left(w(t),\tau(t) \right)$. Then $\tau(t)$ appears to be a
1379: monotonically increasing function since $c(t)>0$ which implies that
1380: $w(\tau)$ has an extremum for each extremum in $w(t)$. This means that
1381: we expect the superposition of a polynomial expansion and a damped
1382: oscillation. The limiting situation is obtained when $\bar Q$ is small
1383: (but not vanishing), and for large $t$, $\tau (t) \sim t$ so that
1384: $w(\tau)$ slowly converges, oscillating, to a constant value.
1385: 
1386: \begin{figure}
1387:   \begin{center}
1388:     \subfigure[$\bar Q \sim 0$]{
1389:       \includegraphics[width=.45\linewidth]{WarpZeroPhi}}
1390:     \subfigure[Small $\bar Q$]{
1391:       \includegraphics[width=.45\linewidth]{WarpSmallPhi}}
1392:     \subfigure[Medium $\bar Q$]{
1393:       \includegraphics[width=.45\linewidth]{WarpMedPhi}}
1394:   \end{center}
1395:   \caption{Typical behaviour for the warping factor in the small-$\bar
1396:     Q$ regime. We consider (a) $\bar Q$ very small but not
1397:     vanishing, (b) small $\bar Q$ and (c) larger $\bar Q$ (but
1398:     still compatible with oscillations). }
1399:   \label{fig:num-solution-warp}
1400: \end{figure}
1401: 
1402: 
1403: \begin{comment}
1404:   
1405: 
1406:   \subsection{Positive curvature spacetimes}
1407:   \label{sec:posit-curv-spac}
1408: 
1409:   Using Eq.~\eqref{eq:Ric-bar} it is immediate to calculate the Ricci
1410:   scalar for the spacetime in Eq.~\eqref{eq:g-bar}. One thus has:
1411:   \begin{multline}
1412:     \bar R = \bar g^{\ssc{mn}} \overline{Ric}_{\ssc{mn}} = \bar g^{tt}
1413:     \overline{Ric}_{tt} + \bar g^{\mu \nu} \overline{Ric}_{\mu \nu} = d
1414:     \left(\ddot \sigma + \dot \sigma^2
1415:     \right) + e^{2 \sigma } R + d \left( d \dot \sigma^2 + \ddot \sigma \right) =\\
1416:     = e^{-2 \sigma} R + d \left( d + 1 \right) \dot \sigma^2 + 2 d \ddot \sigma
1417:   \end{multline}
1418:   or, using the equations of motion
1419:   \begin{equation}
1420:     \bar R = e^{-2 \sigma} R + d \left(1 - d \right) \Sigma^2 - 2 d V^\prime (\sigma) + 4 d \Sigma \dot \Phi   
1421:   \end{equation}
1422: 
1423:   Now, since by construction $g_{\mu \nu}$ and $B_{\mu \nu}$ solve the
1424:   equations of motion it is immediate to see that $R>0$. Moreover in
1425:   the special case $d=3$, $H$ is proportional to the volume form on
1426:   $g$. This implies that $H_{\mu \nu}^2 \propto g_{\mu \nu}$ and then the equations
1427:   reduce to
1428:   \begin{equation}
1429:     Ric_{\mu \nu} = \Lambda^2 g_{\mu \nu}
1430:   \end{equation}
1431:   which implies that $g_{\mu \nu}$ is to be the metric on an Einstein
1432:   three-manifold (the most simple case being a three-sphere).
1433: 
1434:   Let us study in more detail this last case when the space is a
1435:   three-sphere and the metric in Euler coordinates is written as:
1436:   \begin{equation}
1437:     \di s^2 = - \di t^2 + e^{2 \sigma(t)} \left( \di \vartheta^2 + \cos^2 \vartheta \di \phi^2 + \sin^2 \vartheta \di \psi^2  \right)  .
1438:   \end{equation}
1439:   Using the expression above, the Ricci scalar is found to be:
1440:   \begin{equation}
1441:     \bar R = 6 \left( e^{-2 \sigma} + 2  \dot \sigma^2 + \ddot \sigma \right) 
1442:   \end{equation}
1443: 
1444:   If $\sigma(t) $ is a monotonic function (on the linearized system this is
1445:   the case when $\bar Q^2 > 4 R /d = 8$), then $\bar R(t) $ is always
1446:   positive (de Sitter-like). If the initial value $Q (0)$ is small, on
1447:   the other hand, $\sigma (t) $ is an oscillating function and the
1448:   curvature can oscillate (in sign) for early times, before
1449:   asymptotically converging to $\bar R(t) \to R = 6$.
1450: 
1451:   Pictorially, this metric describes an isotropic space which
1452:   conformally oscillates in time and for late times converges to a
1453:   size fixed by the three-dimensional cosmological constant when the
1454:   spacetime becomes static.
1455: 
1456: \end{comment}
1457: 
1458: %%% Local Variables:
1459: %%% mode: latex
1460: %%% TeX-master: "Master"
1461: %%% End:
1462: