1: \chapter{The membrane}
2: \label{ch4}
3: In the previous chapters we discussed the classical geometry of the UHM metric and the corrections due to NS $5$-brane instantons. In this chapter we compute the corrections arising from membranes wrapped along $3$-cycles of the internal rigid CY $3$-fold.\\
4: This will be done in a different way from the semiclassical instanton calculation of chapter \ref{ch3}. The reason is that the most general $4$-dimensional quaternionic metric with (at least) one isometry is known and can be used to describe the membrane corrected metric. As it turns out, this metric is governed by the \emph{Toda equation}, a complicated nonlinear differential equation in three variables. We will construct solutions to this equation and demonstrate that they correspond to membranes by comparing to results from string theory. \\
5: Having obtained these membrane corrections (to the effective action) we can proceed by gauging the isometry, which corresponds to a certain \emph{flux compactification}. This gauging will induce a scalar potential in the low energy effective action. It turns out that the membrane corrections lead to a (meta-stable) de Sitter vacuum.
6:
7: \section{Toda equation and the UHM}
8: \label{todasec1}
9: Let us recall some results from the previous chapters. We saw in equation \R{VI} that the simplest membrane action is given by
10: \equ{
11: S_{\inst}=\frac{2|Q_1|}{g_s} + \I \gf^s Q_1^s \label{eq:act1} \km
12: }
13: with $\gf^s Q_1^s$ given either by $\gvf Q_1^\gvf$ or $\gc Q_1^\gc$. This means that there are two distinct membrane instantons depending on whether one dualizes over $\gvf$ or $\gc$ (see also sections \ref{mem1} and \ref{secDTM1}). We shall confront this with string theory. We note again the factor of $2$ appearing in front of $|Q_1|$, we will come back to this in section \ref{stcal1}.\\
14: The presence of the theta angles in \R{act1} breaks the shift symmetry in $\gvf$ (or $\gc$) to a discrete subgroup. Furthermore, we have seen that the theta angle term of the NS $5$-brane is proportional to the axion $\gs$, see \R{boun2}, which means that the continuous shift symmetry in $\gs$ remains exact as long as we do not switch on $5$-branes. In other words, in the absence of fivebrane instantons, the quantum corrected UHM moduli space will be a quaternionic manifold with a shift isometry in $\gs$. Such manifolds have been classified in terms of a single function.\\
15: In \cite{Przanowski:1991ru} Przanowski derived the general form of $4$-dimensional quaternionic manifolds with at least one isometry, this was later re-derived by Tod \cite{Tod:1995vm}. The Przanowski-Tod (PT) metric reads, in adapted coordinates,
16: \equ{
17: ds^2=\frac{1}{r^2}\left[f dr^2 + f e^h(du^2 + dv^2) + \frac{1}{f}(dt+\gTh)^2\right] \label{eq:tod1} \km
18: }
19: where the isometry acts as a constant shift on $t$. The function $h=h(r,u,v)$ is determined by the $3$-dimensional continuous \emph{Toda equation}
20: \equ{
21: (\p{u}^2 +\p{v}^2)h + \p{r}^2 e^h=0 \label{eq:tod2} \pt
22: }
23: The function $f=f(r,u,v)$ is given in terms of $h$ as follows
24: \equ{
25: f=-\frac{3}{2\gL}(2-r\p{r}h) \label{eq:tod3} \pt
26: }
27: The $1$-form $\gTh=\gTh(r,u,v)=\gTh_r dr+ \gTh_u du + \gTh_v dv$ is a solution to the equation
28: \equ{
29: d\gTh=(\p{u} f dv-\p{v}f du)\wedge dr + \p{r}(f e^h) du\wedge dv \label{eq:tod4} \pt
30: }
31: Manifolds with such a metric are Einstein with anti-selfdual Weyl tensor and thus\footnote{Conversely one can check that starting with \R{tod1} and demanding that the metric is Einstein and has an anti-selfdual Weyl tensor forces $h, f$ and $\gTh$ to obey (\ref{eq:tod2}, \ref{eq:tod3}, \ref{eq:tod4}) respectively.} quaternionic (see section \ref{hypergeo}). The constant $\gL$ in \R{tod3} is the `cosmological' constant of the target-space, $R_{AB}=\gL G_{AB}$.\\
32:
33: To apply the above results to our situation we have to recast the line element \R{hf4t1} into the PT form \R{tod1}, i.e., we have to identify the PT coordinates with the UHM moduli.
34: If we consider the Heisenberg algebra of isometries \R{Heis-alg2}, we see that we can either identify $t$ with $\gs$ or with $\gvf$. In the first case the shift is generated by $\ga$ and in the second case by $\gb$. This leads to two possible representations of the PT metric that describe the same moduli space. We call these bases the membrane and the fivebrane bases respectively. In the membrane basis, which is the relevant basis, we identify the coordinate $t$ with $\gs$ such that the $\ga$-shift symmetry is manifest. The reason is the absence of $5$-brane instantons, which ensures a continuous $\ga$-shift symmetry. The coordinates can then be chosen as
35: \equ{
36: t=\gs \qquad r=e^\gf \qquad u=\gc \qquad v=\gvf \label{eq:fmap1} \pt
37: }
38: In this basis the classical moduli space metric of the UHM \R{hf4t1}\footnote{This metric has $\gL=-3/2$.} corresponds to the solution $e^h=r$, which gives $f=1$ and $\gTh=u dv$. Note that $\gTh$ is only defined up to an exact form.\\
39: Previously (see section \ref{vze} for instance) we have mentioned that there are perturbative ($1$-loop only) corrections to the UHM \cite{Strominger:1997eb, Antoniadis:2003sw}. We can easily incorporate those in the PT framework by using the following observation. The Toda equation is insensitive to constant shifts in $r$ (or in $u$ and $v$ for that matter) which means that $h(r+c,u,v)$ is also a solution to the Toda equation, $c\in \mR$. Applying this to the classical solution $r=e^h$, gives
40: \equ{
41: e^h=r+c \qquad f=\frac{r+2c}{r+c} \qquad \gTh=u dv \label{eq:pertu1} \pt
42: }
43: This exactly reproduces the $1$-loop corrected metric (in the string frame) of \cite{Antoniadis:2003sw} provided we identify
44: \equ{
45: c=-\frac{4\gz(2)\gc(Y_3)}{(2\gp)^3}=\frac{h^{(1,2)}-h^{(1,1)}}{6\gp} \nn \km
46: }
47: where $\gc(Y_3)$ is the Euler number of the CY $3$-fold on which the type IIA string theory has been compactified (see \R{euler1}). We consider only rigid Calabi-Yau's so $h^{(1,2)}=0$, which gives the bound
48: \equ{
49: c<0 \label{eq:c2} \pt
50: }
51: The PT coordinate $r$ is related to $\gr$ in \cite{Antoniadis:2003sw} through $r=\gr^2-c=e^\gf$, the relation between the fields and PT coordinates receives no (perturbative) quantum corrections.\\
52: Note that for $c<0$ the metric becomes negative-definite because $f$ becomes negative for $r<2|c|$. As a result, we have to restrict ourselves to the open interval $2|c|<r<\infty$.
53:
54: \subsubsection{Relation with other work}
55: The $1$-loop correction to the classical UHM was derived in \cite{Antoniadis:2003sw} by searching for deformations that respect both the Heisenberg group and the quaternionic structure of the UHM. This can be done systematically because the most general form of a $4$-dimensional quaternionic manifold with (at least) two commuting isometries has been constructed in \cite{Calderbank:2001uz}. This metric is expressed in terms of a single function which is determined by a linear differential equation. This is different from the PT metric that is governed by a nonlinear differential equation \R{tod2}.\\
56: We should also mention work of Ketov \cite{Ketov:2001gq, Ketov:2001ky, Ketov:2002vr} in which membrane (and five-brane) solutions are discussed. An important difference with our work is that Ketov works on the basis that the membrane instantons preserve two isometries, which allows him to use \cite{Calderbank:2001uz}. However, we have seen explicitly that membrane instantons preserve only one continuous isometry. That is, we can either use the shift in $\gvf$ or $\gc$ to dualize to the DTM for which we can construct membrane instanton solutions, with action \R{act1}. In section \ref{stcal1} we shall see that this corresponds to either wrapping the membrane over the `$\cA$-cycle' or `$\cB$-cycle'. In the end one should sum them both up to obtain the effect of the two membrane\footnote{So if one considers only one membrane instanton there are two isometries left, although they are not the same as the one Ketov uses.} (instantons), to correlation functions for instance. In that situation both isometries are broken to a discrete subgroup and only the shift in $\gs$ is left, in the absence of $5$-brane instantons at least.
57:
58: \subsubsection{The $5$-brane}
59: By now the reader might wonder why we have not used the PT metric to calculate $5$-brane corrections to the UHM moduli space. This would require that we identify $t$ with $\gvf$ (or $\gc$). However, then the mapping of the other PT coordinates to the classical (plus one-loop) UHM metric cannot be consistently realized to describe the fivebrane.\\
60: Current investigations indicate that this is because of the fact that four-dimensional quaternionic metrics with (at least) one isometry fall into two classes, see \cite{Przanowski:1991ru}. The one class constitutes the PT framework as described above and the other class involves a different differential equation, instead of the Toda equation. This class should probably be used to describe the fivebrane.
61:
62:
63:
64: \section{Solutions to the Toda equation}
65:
66: We would like to find a solution that mimics an (infinite) series of exponential corrections describing membrane instantons. We have seen that the real part of the action \R{act1} is inversely proportional to $g_s$. In PT coordinates this real part is given by $2\sqrt{r}$. Therefore we make a fairly general ansatz of the form
67: \equ{
68: e^h=r+\sum_{n\geq 1}\sum_m f_{n,m}(u,v)r^{-m/2 +\ga}e^{-2n \sqrt{r}} \label{eq:ansatz1} \km
69: }
70: where we expand in the ($1/g_s$ part of the) instanton action: $(e^{-S_\inst})^n$. A slightly more general ansatz is given by replacing the exponent with $e^{-\gd n \sqrt{r}}$ where $\gd$ is a constant. However, it turns out that this ansatz only satisfies the Toda equation (with certain functions $f$ that we will construct shortly) if $\gd=2$.
71: There exist interesting solutions to the Toda equation, see \cite{Calderbank:1999ad, Bakas:1996gf}, but with the mapping of PT coordinates to the fields that we use \R{fmap1}, these do not describe physically interesting solutions. In particular, they do not describe a membrane instanton expansion.
72: As observed earlier, we can always shift $r$ with a constant to produce a new solution. This means that we can first find a solution for $e^h(r,u,v)$ and then take $r\to r+c$ in this solution, thus incorporating the one-loop correction in one go.
73: The ansatz, apart from the classical piece, consists of a double sum. There is the sum over $n$ which gives an expansion in $e^{-2/g_s}$, meaning that $n$ is the instanton number. For each $n$ there is an expansion in $m$. This is a perturbative (loop) expansion in a given instanton sector, since it goes as $r^{-m/2}=(g_s)^m$ and the sum $m$ runs over integers. To account for the possibility that $m$ does not\footnote{As is the case in \cite{Ooguri:1996me} for instance.} run over integers, we have added a parameter $\ga \in [0,\2)$. We will later show that the Toda equation is satisfied for $\ga=0$ only, see appendix \ref{toda1}.
74: At each instanton level $n$ there is, by definition, a lowest value $m_n$ that defines the leading term in the expansion
75: \equ{
76: f_{n,m}(u,v)=0 \qquad \textrm{for}\quad m<m_n\nn\pt
77: }
78: Because we have spelled out the $r$ dependence explicitly, the Toda equation gives a series of differential equations for the $f_{m,n}(u,v)$. These equations can be solved iteratively, order by order in $n$ and $m$ to any desired order. To examine the Toda equation \R{tod4} in detail we first write it in the equivalent form
79: \begin{equation} \label{eq:Toda2}
80: e^h \big( \partial_u^2 + \partial_v^2 + e^h \partial_r^2 \big) e^h - (\partial_u e^h)^2
81: - (\partial_v e^h)^2 = 0\pt
82: \end{equation}
83: Using ansatz \R{ansatz1} equation \R{Toda2} can be written as
84: \begin{align}\label{eq:4.5}
85: 0 = \sum_{n,m} &\ r^{-m/2}\, e^{-2n\sqrt{r}}\, \Big\{ (\Delta +
86: n^2)\, f_{n,m+2} + n\, a_{m+2}\, f_{n,m+1} + b_{m+2}\, f_{n,m}
87: \notag \\[-2pt]
88: & + \sum_{n',m'} e^{-2n'\!\sqrt{r}}\, \big[ 2n\, a_{m'+1}\, f_{n',
89: m-m'-1} + 2 b_{m'+2}\, f_{n',m-m'-2} \notag \\[-6pt]
90: & \mspace{112mu} + f_{n',m-m'}\, (\Delta + 2n^2) - \nabla f_{n',m-m'}
91: \cdot \nabla \big] f_{n,m'} \notag \\[4pt]
92: & + \sum_{n',m'} \sum_{n'',m''} e^{-2(n'+n'')\sqrt{r}}\, f_{n,m'}
93: f_{n',m''}\, \big[ n^2 f_{n'',m-m'-m''-2} \notag \\[-8pt]
94: & \mspace{94mu} + n\, a_{m'+1}\, f_{n'',m-m'-m''-3} + b_{m'+2}\,
95: f_{n'',m-m'-m''-4} \big] \Big\} \km
96: \end{align}
97: where $\nabla\equiv (\partial_u,\partial_v)$, $\Delta\equiv \nabla^2$ and
98: \begin{equation}\label{eq:4.6}
99: a_m \equiv \half\, (2m-1)\ ,\qquad b_m \equiv \quart\, m(m-2)\pt
100: \end{equation}
101: Equation \R{4.5} is organized in terms of single-, double- and triple sums. We can write a generic exponential in \R{4.5} as $e^{-2N\sqrt{r}}$.
102: In the $(N=1)$-instanton sector only the single-sum terms contribute. The double- and triple-sums have to be taken into account beginning with the $(N=2)$- and $(N=3)$-instanton sectors, respectively. We will save some details for appendix \ref{toda1} where we prove that $m_n\geq-2$ for all $n$ and $\ga=0$.\\
103:
104: \subsection{The one-instanton sector}
105: \label{1insts}
106: We shall start with the one-instanton sector, $N=1$, in which case equation \R{4.5} reduces to
107: \begin{equation} \label{eq:f1m}
108: (\Delta + 1)\, f_{1,m} + a_m\, f_{1,m-1} + b_m\, f_{1,m-2} = 0\pt
109: \end{equation}
110: If for simplicity we consider first a one-dimensional truncation, we can prove that the general solution is given by
111: \begin{equation} \label{eq:solf1m}
112: f_{1,m}(x) = \Re \sum_{s\geq 0}\, \frac{1}{s!\,(-2)^s}\, k_{1,m}(s)\,
113: G_s(x) \km
114: \end{equation}
115: where $x\in \{u,v\}$. The coefficients are defined recursively as
116: \begin{equation}\label{eq:recurse-rel}
117: k_{1,m}(s+1) = a_m k_{1,m-1}(s) + b_m k_{1,m-2}(s) \pt
118: \end{equation}
119: The $G_s(x)$ are complex functions related to spherical Bessel functions of the third kind, see appendix \ref{toda1}. The coefficients $k_{1,m}(s)$ have as their lowest values in $s$
120: \equ{
121: k_{1,m}(0)=A_{1,m} \nn \km
122: }
123: which are the complex integration constants originating from the homogeneous part of \R{f1m}. By definition we have $f_{1,m}=0$ for $m<m_1$, which implies that $k_{1,m}(s)=0$ and in particular $A_{1,m}=0$ for $m<m_1$. The recursion relation \R{recurse-rel} furthermore implies $k_{1,m}(s>m-m_1)=0$, which means that the highest monomial in $x$ contained in $f_{1,m}(x)$ is of order $m-m_1$. This is nicely illustrated by the first two solutions
124: \begin{align}
125: & f_{1,m_1}(x) = \Re \big\{ A_{1,m_1} e^{\I x} \big\}\, , \notag
126: \\[2pt]
127: & f_{1,m_1+1}(x) = \Re \big\{ A_{1,m_1+1} e^{\I x} + \half a_{m_1
128: +1} A_{1,m_1}\, \I x\, e^{\I x} \big\}\pt \label{eq:solf1m0}
129: \end{align}
130: To solve \R{f1m} for a $(u,v)$-dependent solution one can proceed by separation of variables, which yields a basis of solutions. The most general solution is then obtained by superposition and can be written as
131: \bea
132: && f_{1,m}(u,v) = \int\! d\lambda\, \Re \sum_{s\geq 0}\, \frac{1}{s!\,
133: (-2\, \omega^2)^s}\, k_{1,m}(s, u; \lambda)\, G_s(\omega v)\ ,
134: \notag \\[1.5mm]
135: &&k_{1,m}(s+1, u; \lambda) = a_m k_{1,m-1}(s, u; \lambda) + b_m k_{1,
136: m-2}(s, u; \lambda)\km\label{eq:solf2m}
137: \eea
138: where
139: \begin{equation}
140: k_{1,m}(0, u ; \lambda) = B_{1,m}(\lambda)\, A_{1,m}
141: (\lambda) e^{\I\lambda u} \pt \nn
142: \end{equation}
143: The basis is parametrized by a continuous real parameter $\gl$ which is integrated over in \R{solf2m} and $\go\equiv \sqrt{1-\gl^2}$. $A_{1,m}(\gl)$ and $B_{1,m}(\gl)$ are arbitrary complex integration constants which determine the `frequency spectrum' of the solution.
144:
145: The $(u,v)$-dependent solution \R{solf2m} is too general for our physical problem, the reason being that the general form with its integration over $\gl$ has products of exponents in $\I u$ and $\I v$. These correspond in the supergravity description to the theta-angle-like terms. However as we were reminded of by \R{act1}, we can have either the $\gc$ or the $\gvf$ theta-angle and corresponding charges switched on. One either has the `$u$-instanton' ($\gc$) or the `$v$-instanton' ($\gvf$), not both simultaneously. Consequently we must restrict \R{solf2m} to reflect this property. \\
146: This can be done by letting the coefficient functions peak around $\gl=0$ and $\gl=1$. If we set
147: \begin{equation}
148: A_{1,m}(\lambda) = A_{1,m}\ ,\qquad B_{1,m}(\lambda) = \delta
149: (\lambda)\km \nn
150: \end{equation}
151: we obtain \R{solf1m} as a function\footnote{We will refer to this configuration as the `v-instanton'. The v-anti-instanton has opposite signs in the exponentials: $e^{-\I v}$ instead of $e^{\I v}$. Similarly for the `u-instanton'.} of $v$: $f_{1,m}(v)$ with $k_{1,m}(0)=A_{1,m}$.
152: Similarly one can produce a solution independent of $v$ ($x=u$) by taking
153: \equ{
154: A_{1,m}(\gl)=\gd(\gl-1) \qquad B_{1,m}(\gl)=B_{1,m} \nn \km
155: }
156: which produces $f_{1,m}(u)$ with $k_{1,m}(0)=B_{1,m}$ as integration constants. Naturally, the sum of these two solutions is a solution to \R{f1m} as well. The one-instanton sector $e^h$ can therefore completely be expressed in terms of the one-dimensional solutions as
157: \bea
158: &&\exp [h(r,u,v)] = \exp [h_\text{pert}(r)] + \exp [h_\text{1-inst}(r,u)
159: ] + \exp [h_\text{1-inst}(r,v)] + \ldots \nn \\ [2.9mm]
160: && \exp [h_\text{pert}(r)] = r + c \nn \\[2mm]
161: && \exp [h_\text{1-inst}(r,u)] = e^{-2 \sqrt{r+c}}\, \sum_{m\geq m_1}
162: f_{1,m}(u)\, (r+c)^{-m/2} \label{eq:9.1}
163: \eea
164: and similarly for $h_\text{1-inst}(r,v)$. We see that the $u$- and the $v$-instanton contribute separately to $e^h$ and thus to the line element \R{tod1}. The ellipses denote the higher $(n>1)$ instanton contributions. We have applied the shift in $r$: $r\to r+c$ to incorporate the one-loop result.
165: For later reference, we also give the leading order expression for
166: $e^h$ in the regime $r\mg 1$ (small string coupling). To leading order
167: in the semi-classical approximation, the instanton solution \R{9.1}
168: reads
169: \begin{equation} \label{eq:eh-exp}
170: e^h = r + c + \frac{1}{2}\, r^{-m_1/2}\, \big( A_{1, m_1}\, e^{\I v}
171: + A_{1, m_1}^\star\, e^{-\I v} + B_{1, m_1}\, e^{-\I u} + B_{1, m_1}^\star\,
172: e^{\I u} \big)\, e^{-2\sqrt{r}} + \ldots
173: \end{equation}
174: Note that we need to include both instantons and anti-instantons to
175: obtain a real solution.
176: To find the leading-order instanton corrected hypermultiplet metric, we
177: first compute the leading corrections to $f$ defined in \R{tod3}:
178: \bea
179: && f = \frac{r + 2c}{r+c} + \label{eq:finst} \\[1.5mm]
180: && \frac{1}{2}\, r^{-(m_1 + 1)/2}\, \big( A_{1,
181: m_1}\, e^{\I v} + A_{1,m_1}^\star\, e^{-\I v} + B_{1,m_1} \, e^{-\I u}
182: + B_{1,m_1}^\star\, e^{\I u} \big)\, e^{-2\sqrt{r}} + \ldots \nn
183: \eea
184: Substituting this result into \R{tod4} yields the leading
185: corrections to the $\Theta$ $1$-form. Setting
186: \begin{equation}
187: \Theta = u\, \d v\ + \Theta_\text{inst}\km
188: \end{equation}
189: these are given by
190: \begin{equation}\label{eq:Tinst}
191: \Theta_\text{inst} = r^{-m_1/2}\, e^{-2 \sqrt{r}}\, \Im \{ A_1\,
192: e^{\I v}\, \d u + B_1\, e^{-\I u}\, \d v \}\ + \ldots\pt
193: \end{equation}
194: The leading order corrections to the hypermultiplet scalar metric are
195: then obtained by substituting these expressions into the PT metric
196: \R{tod1}.\\
197:
198: \subsection{The two-instanton sector}
199: We now briefly discuss the $N=2$ sector. The Toda equation requires at
200: this level
201: \begin{align}\label{eq:N=2sector}
202: 0 & = (\Delta + 4)\, f_{2,m} + 2a_m\, f_{2,m-1} + b_m\, f_{2,m-2}
203: \notag \\[2pt]
204: & \tab + \sum_{m'} \big[ f_{1,m-m'-2} + a_{m'+1}\, f_{1,m-m'-3}
205: + b_{m'+2}\, f_{1,m-m'-4} \notag \\[-6pt]
206: & \mspace{76mu} - \nabla f_{1,m-m'-2} \cdot \nabla \big] f_{1,m'}\ ,
207: \end{align}
208: where we have used \R{f1m} for $\Delta f_{1,m'}$ in the double sum.
209: We have not derived the general solution to these equations in closed
210: form. The one-dimensional truncation is straightforward to
211: solve order by order in $m$. At lowest order\footnote{In appendix
212: \ref{toda1} we show that $-2\leq m_n\leq m_{n'}$ for $n\geq n'$.}
213: $m_2$ we have
214: \begin{equation} \label{eq:2m0eq}
215: (\Delta + 4)\, f_{2,m_2} + \gd_{m_2,-2}\, \big[ (f_{1,m_2})^2 -
216: (\nabla f_{1,m_2})^2 \big] = 0\ .
217: \end{equation}
218: Note that the inhomogeneous term is present only for the lowest possible
219: value $m_2=-2$. The one-dimensional truncation yields the equation
220: \begin{equation}
221: (\partial_x^2 + 4)\, f_{2,m_2}(x) + \gd_{m_2,-2}\, \Re \big\{ A_{1,m_2}^2
222: e^{2\I x} \big\} = 0\ ,
223: \end{equation}
224: where we have inserted the solution \R{solf1m0} for $f_{1,m_1}(x)$.
225: The general solution then reads
226: \begin{align}
227: f_{2,m_2}(x) & = \Re \big\{ A_{2,m_2} e^{2\I x} + \tfrac{1}{4}
228: \gd_{m_2,-2}\, A_{1,m_2}^2\, \I x\, e^{2\I x} \big\} \notag
229: \\[1mm]
230: & = \Re \big\{ A_{2,m_2} G_0(2x) - \tfrac{1}{8} \gd_{m_2,-2}\, A_{1,
231: m_2}^2 G_1(2x) \big\}\ ,
232: \end{align}
233: $A_{2,m_2}$ being a further complex integration constant.
234:
235: The solution for $m>m_2$ can now be constructed by solving the
236: appropriate equation arising from \R{N=2sector}. Based on
237: \R{solf2m} we can also construct the general $(u,v)$-dependent
238: solution for $f_{2,m_2}(u,v)$. The idea is to decompose the products of
239: $\cos(\lambda_1 u)\,\cos(\lambda_2 u)$, etc., appearing in the
240: inhomogeneous part of \R{2m0eq} into a sum of $\cos$ and $\sin$
241: terms using product formulae for two trigonometric functions. We can
242: then construct the full inhomogeneous solution by superposing the
243: inhomogeneous solutions for every term in the sum. We refrain from
244: giving the result, since it is complicated and not particularly
245: illuminating.
246:
247: We conclude this subsection by giving an argument that the iterative
248: solution devised above indeed gives rise to a consistent solution of the
249: Toda equation. The general equations which determine a new $f_{n,m}(u,
250: v)$ are two-dimensional Laplace equations to the eigenvalue $n^2$
251: coupled to an inhomogeneous term, which is completely determined by the
252: $f_{n,m}(u,v)$'s obtained in the previous steps of the iteration
253: procedure. These equations are readily solved, e.g., by applying a
254: Fourier transformation. It then turns out that the iteration procedure
255: is organized in such a way that any level in the perturbative expansion
256: \R{4.5} determines one `new' $f_{n,m}(u,v)$, i.e., there are no
257: further constraints on the $f_{n,m}(u,v)$ determined in the previous
258: steps. This establishes that the perturbative approach indeed extends to
259: a consistent solution of the Toda equation \R{tod2}.
260:
261: \subsection{The isometries}
262: \label{sect4.3}
263: Based on the Toda solution \R{9.1} we now discuss the breaking of
264: the Heisenberg algebra \R{Heis-alg2} in the presence of membrane
265: instantons to a discrete subgroup. We start with the shift symmetry in the axion $\sigma
266: \rightarrow\sigma-\alpha$. By identifying $t=\sigma$, this shift
267: corresponds to the isometry of the PT metric, so that it cannot
268: be broken by the instanton corrections.
269:
270: Analyzing the $\beta$ and $\gamma$-shifts is more involved. Under the
271: identification \R{fmap1} the $\beta$-shift acts as $v\rightarrow
272: v+\beta$. Taking the leading order one-instanton solution
273: \R{eh-exp}-\R{Tinst}, we find that $e^h$ as well as the
274: resulting functions $f$ and $\Theta$ appearing in the metric depend on
275: $v$ through $e^{\pm\I v}$ or $dv$ only. These theta-angle-like terms
276: break the $\beta$-shift to the discrete symmetry group\footnote{This agrees with earlier observations made in
277: \cite{Becker:1999pb}.}
278: $\mZ$. Going beyond the leading instanton corrections by taking
279: into account higher loop corrections around the single instanton will
280: generically break the $\beta$-shift completely. This is due to the
281: appearance of polynomials in $v$ multiplying the factors $e^{\pm\I v}$.
282: We point out that by setting the integration constants
283: multiplying the terms odd in $v$ to zero, there is still an unbroken
284: $\mZ_2$ symmetry. This symmetry is defined by $v\rightarrow-v$, $t\rightarrow-t$, which amounts to
285: interchanging $v$-instantons and anti-instantons.
286:
287: To deduce the fate of the $\gamma$-shift, $u\rightarrow u+\gamma$,
288: $t\rightarrow t-\gamma v$, we first observe that $t\rightarrow t-\gamma
289: v$ implies that the combination $dt+udv$ is invariant.
290: Applying the same logic as for the $\beta$-shift above, we then find
291: that the one-loop corrections of a single $u$-instanton break the
292: $\gamma$-shift to the discrete symmetry $\mZ$, which will be
293: generically broken by higher order terms appearing in the loop
294: expansion. Similar to the $\beta$-shift, however, we can arrange the
295: constants of integration appearing in the solution in such a way that
296: there is also a $\mZ_2$ symmetry. We expect that these two
297: $\mZ_2$ symmetries could play a prominent role when determining
298: (some of) the coefficients appearing in the solution \R{9.1} from
299: string theory.
300:
301:
302:
303: \section{String theory}
304: So far we have calculated membrane effects from a supergravity point of view. We already noted that this construction by itself was too general and needed to be tailored to the physical problem. Therefore we would like to compare these results with a microscopic string derivation of membrane effects.
305: In \cite{Becker:1995kb} Becker, Becker and Strominger did precisely that, i.e., they considered the compactification of M-theory on a CY $3$-fold (and a circle) with membranes wrapped along $3$-cycles. These membranes give rise to certain corrections to $4$-fermion correlation functions. This is not surprising: we already saw in section \ref{unbroken} on page \pageref{mem4a} that the membrane preserves one half of the supersymmetries. Conversely, it breaks one half of the supersymmetries. The resulting four fermionic zero modes lead to nonvanishing $4$-fermion correlators. \\
306: We will review the analysis of \cite{Becker:1995kb} and explicitly evaluate their result for the $4$-fermion correlation functions in the case of a rigid Calabi-Yau manifold. Then we shall compute the same object, but starting with the results obtained in the previous sections. We will find that these results agree beautifully. \\
307:
308: \subsection{The string calculation}
309: \label{stcal1}
310: In section \ref{11d} we gave the $11$-dimensional supergravity action. Let us write it down again in Euclidean signature with the conventions of \cite{Becker:1995kb}
311: \equ{
312: S_{11}=\frac{1}{2\gp^2}\int d^{11}x \sqrt{g} \big[ -R + \frac{1}{48} (dA_3)^2\big] + \frac{\!\I}{12\gp^2}\int A_3\wedge dA_3 \wedge dA_3 \label{eq:11dsugra} \km
313: }
314: where $A_3$ is the $3$-form potential, as in \R{11dsugra1}, note that we leave out the hat. We work in units in which the $11$-dimensional Planck length equals one.\\
315: The membrane solutions can effectively be described (for scales large compared to the thickness of the membrane) by its worldvolume action. This action is completely fixed by supersymmetry \cite{Bergshoeff:1987cm}\footnote{See also \cite{Bergshoeff:1987qx, Duff:1990xz, Hughes:1986fa}.} to be
316: \bea
317: S_3&=&\int d^3 \gs \sqrt{h}\big[ \2 h^{\ga\gb} \p{\ga}X^M \p{\gb}X^N g_{MN} \nn \\[1mm]
318: && -\frac{\I}{2} \bgTh \gG^\ga \nabla_\ga \gTh + \frac{\I}{3!}\ge^{\ga\gb\gg}A_{MNP}\p{\ga}X^M \p{\gb}X^N \p{\gg}X^P + \ldots\big] \nn \km
319: \eea
320: where $\gs^\ga$ with $\ga,\gb, \gg=1,2,3$ are the worldvolume coordinates and $h_{\ga\gb}$ is the (auxiliary) worldvolume metric with Euclidean signature. The $X^M(\gs)$ with $M,N=1,\ldots,11$ describe the membrane configuration: the maps embedding the membrane into the $11$-dimensional space. $\gTh$ is an $11$-dimensional Dirac spinor. Only the leading terms in powers of the fermions are given, ellipses denote higher order terms. \\
321: The global fermionic symmetries of $S_3$ act on the membrane fields as
322: \equ{
323: \gd_\ge \gTh=\ge \qquad \gd_\ge X^M=\I \bge \gG^M \gTh \nn \km
324: }
325: where $\ge$ is a constant spinor in $11$ dimensions. The worldvolume theory is also invariant under the so-called $\gk$ symmetries, local fermionic transformations
326: \equ{
327: \gd_\gk \gTh=2 P_+ \gk(\gs) \qquad \gd_\gk X^M=2\I \bgTh \gG^M P_+ \gk(\gs) \label{eq:kappa1} \km
328: }
329: with $\gk$ a spinor in $11$ dimensions. The projection operators
330: \equ{
331: P_\pm=\2\left(1\pm \frac{\I}{3!}\ge^{\ga\gb\gg}\p{\ga}X^M\p{\gb}X^N\p{\gg}X^P\gG_{MNP}\right) \nn \km
332: }
333: obey (see \cite{Becker:1995kb})
334: \equ{
335: P_{\pm}^2=P_\pm \qquad P_+ P_-=0 \qquad P_+ + P_-=1 \label{eq:kappa2} \pt
336: }
337: A general bosonic membrane configuration $X(\gs)$ breaks all the global supersymmetries generated by $\ge$. Some unbroken supersymmetries remain only if they are compensated by a $\gk$-transformation such that
338: \equ{
339: \gd_\ge \gTh + \gd_\gk \gTh=0 \nn \pt
340: }
341: If we apply $P_-$ and use \R{kappa1} we find that the \susy parameter which leaves the configuration invariant must obey
342: \equ{
343: P_-\ge=0 \label{eq:kappa3} \pt
344: }
345: Such spinors are left invariant under $P_+$, that is, $P_+\ge=\ge$ as can be seen using \R{kappa2}. On the other hand, if we have a spinor $\ge_0$ which obeys $P_+\ge_0=0$, \R{kappa2} gives $P_-\ge_0=\ge_0$. So this spinor breaks \susy and will therefore generate zero modes, compare with the discussion in section \ref{ubs}. \\
346:
347: Condition \R{kappa3} must be examined in the case when we compactify on a CY $3$-fold $Y_3$ while the membrane is wrapped along its $3$-cycles. To this end introduce complex coordinates $X^m, X^\bn$ $m,\bn=1,2,3$ for the Calabi-Yau manifold. Furthermore, let $\ge_+=(\ge_-)^\star$ be two covariantly constant $6$-dimensional spinors %\footnote{In an imaginary representation of the $6$-dimensional gamma matrices $\gg_m$.}
348: with opposite chirality, $\ge_+$ has by definition positive chirality. Their normalization can be chosen such that
349: \equ{
350: \gg_{mnp}\ge_+=e^{-\cK}\gO_{mnp}\ge_- \qquad \gg_{\bm np}\ge_+=2\I K_{\bm [n}\gg_{p]} \ge_+ \qquad \gg_{\bm}\ge_+=0 \label{eq:5H} \km
351: }
352: with $K$ the \Kh form, see section \ref{kahlergeo}. We have introduced
353: \equ{
354: \cK\equiv\2 \left(\cK_V-\cK_H\right) \label{eq:Kpot1} \km
355: }
356: where $\cK_H$ is the \Kh potential on the moduli space of complex structures\footnote{This potential was already introduced in equation \R{kahmod1}.} and $\cK_V$ the potential on the \Kh moduli space
357: \equ{
358: \cK_H=-\ln\left(\I\int_{Y_3} \gO\wedge \bgO\right) \qquad \cK_V=-\ln\left(\frac{4}{3}\int_{Y_3} K\wedge K\wedge K\right) \label{eq:Kpot2} \pt
359: }
360: If we split up the spinor $\ge$ in a $6$- and a $5$-dimensional\footnote{Later we will go to four spacetime dimensions, for now it does not matter.} part as $\ge_\gth\otimes \gl$ with $\gl$ the $5$-dimensional spinor and
361: \equ{
362: \ge_\gth\equiv e^{\I\gth}\ge_+ + e^{-\I \gth}\ge_- \nn \km
363: }
364: $P_-\ge=0$ implies $P_-\ge_\gth=0$. We can examine this condition using \R{kappa2} and \R{5H} which give two conditions:
365: \bea
366: && \p{[\ga}X^m \p{\gb]} X^\bn K_{m\bn}=0 \label{eq:condi1} \\[2mm]
367: && \p{\ga}X^m\p{\gb}X^m \p{\gs}X^p \gO_{mnp}=e^{\I \gvf}e^\cK \gve_{\ga\gb\gg} \label{eq:condi2} \km
368: \eea
369: with $\gvf\equiv 2\gth + \gp/2$. Thus the pullback of the \Kh form onto the membrane must vanish and the membrane volume element has to be proportional to the pullback of $\gO$ onto the membrane. If these two conditions are met, the map $X(\gs)$ is called a \emph{supersymmetric cycle} because then $P_-\ge=0$ is satisfied and supersymmetry is (partially) preserved.\\
370:
371: We will consider the case where the membrane is wrapped around such a supersymmetric cycle, denoted by $\cC_3$, such that we have four broken and four unbroken supersymmetries.\\
372: As in the previous chapter, the broken supersymmetries will give rise to corrections to the low energy effective action. We saw that the broken \susy parameters are those $\ge_0$ such that $P_+\ge_0=0$. With these spinors we can generate zero modes for the fermions in the hypermultiplets\footnote{The hypermultiplet sector is (almost) identical in four and five spacetime dimensions, see page \pageref{spinors}.} of the low energy effective theory. In \cite{Becker:1995kb} they are denoted by $\gc^I$; these are symplectic-Majorana spinors (see page \pageref{spinors}). The correlator
373: \equ{
374: \langle(\bgc_I\gc_J)(\bgc_K \gc_L)\rangle_\inst \label{eq:corf4a} \km
375: }
376: gives rise to membrane instanton corrections to a symmetric tensor $\cR_{IJKL}$ with indices $I,\ldots, L=1,\ldots,2n$, with $n=h^{(1,2)}+1$. By constructing vertex operators for the spinors $\gc^I$ the correlator can be evaluated giving the correction
377: \equ{
378: \gD_{\cC_3}\cR_{IJKL}=N e^{-S_\inst}\int_{\cC_3}d_I\int_{\cC_3}d_J\int_{\cC_3}d_K\int_{\cC_3}d_L \label{eq:corrs1} \km
379: }
380: where $N$ is some normalization factor which includes an unknown determinant which may very well depend on $g_s$. The instanton action $S_\inst$ is given by
381: \equ{
382: S_\inst =e^{-\cK}\left|\int_{\cC_3}\gO\right| + \I \int_{\cC_3}A_3 \label{eq:535b} \km
383: }
384: where $A_3$ is the $3$-form potential from $11$-dimensional supergravity, see \R{11dsugra}. The $d_I$ form a real symplectic basis of $H^3(Y_3,\mZ)$ which obeys
385: \equ{
386: \int d^I\wedge d^J=\gve^{IJ} \nn\km
387: }
388: with $\gve^{IJ}$ the invariant antisymmetric tensor\footnote{Remember that we can use this tensor to raise and lower indices of the fermions, in this case of the $\gc_I$.} of $Sp(2h^{(1,2)}+2)$ and $I,J=1,\ldots,2h^{(1,2)}+2$. Note that this is the same as \R{basissym1}, although for a different basis.
389: Equation \R{corrs1} is the object we have to evaluate explicitly. Introduce a basis of real $3$-cycles $(\cA^a, \cB_a)$ of $H_3(Y_3,\mZ)$ satisfying
390: \bea
391: &&\int_{\cA^a}\ga_b=\int_{Y_3}\ga_b\wedge \gb^a=\gd^a_b \nn \\[1mm]
392: &&\int_{\cB_b}\gb^a=\int_{Y_3} \gb^a\wedge \ga_b=-\gd^a_b \nn\\[1mm]
393: &&\int_{\cA^a}\gb^b=\int_{\cB_a}\ga_b=0 \label{eq:c5b} \pt
394: \eea
395: We have used the basis of harmonic $3$-forms $(\ga_a, \gb^b)$ with $a,b=0,\ldots,h^{(1,2)}$ as in \R{basissym1}. Note that to avoid confusion with the notation of \cite{Becker:1995kb} we have used the indices $a,b$. \\
396: For rigid Calabi-Yau manifolds $h^{(1,2)}=0$ and because we have only the one index $a=0$, we might as well omit it. As in section \ref{CY} we then proceed to define the periods of the holomorphic $3$-form $\gO$ as
397: \equ{
398: Z=\int_\cA\gO \qquad \cG=\int_\cB\gO \nn
399: }
400: and thus $\gO=Z\ga-\cG(Z)\gb$. The $Z$ are projective coordinates and we have only one $Z$, so we can take it to be equal to $1$. We choose the normalization such that\footnote{This choice of $\cG$ also ensures that $\cK_H$ is real.} $\cG=-\I Z^2$, $\gO$ then becomes
401: \equ{
402: \gO=\ga+\I\gb \label{eq:tc5c} \pt
403: }
404: Together with the conventions in \R{c5b} this results in the volume of $Y_3$ being normalized to $1$. In general we have, see \cite{Bodner:1990zm}, for a CY $3$-fold
405: \bea
406: \int\gO\wedge\bgO &=&\left(\frac{1}{3!}\right)^2\int d^6\gx \,\gO_{ijk}\bgO_{\bl\bm\bn}\gve^{ijk}\gve^{\bl\bm\bn}\nn\\
407: &=&-\I \left|\!\left|\gO\right|\!\right|^2 \int d \textrm{Vol}(Y_3) \nn \\
408: &=&-\I \left|\!\left|\gO\right|\!\right|^2 \textrm{Vol}(Y_3) \label{eq:volf2} \km
409: \eea
410: with $\gx$ some complex coordinates on $Y_3$. This means that the volume form equals
411: \equ{
412: d \textrm{Vol}(Y_3)=\frac{\gO\wedge\bgO}{-\I \left|\!\left|\gO\right|\!\right|^2 }=\ga\wedge\gb \label{eq:volf1} \km
413: }
414: where we have used \R{tc5c} and $\left|\!\left|\gO\right|\!\right|^2\equiv \frac{1}{3!}\gO_{ijk}\bgO^{ijk}=2$ . If we now use \R{c5b} we see that Vol$(Y_3)=\int\ga\wedge\gb=1$.\\
415: Our next job is to find the supersymmetric cycle $\cC_3$ for the rigid Calabi-Yau manifold. We shall demonstrate that the real $3$-cycles $\cA$ and $\cB$ are themselves supersymmetric cycles. \\
416: The first condition \R{condi1} states that the pullback of the \Kh form of $Y_3$ onto the membrane, that is onto $\cA$ or $\cB$, has to vanish.\\
417: To prove this, let us generalize the argument given in section $2.2$ of \cite{Becker:1995kb}. Suppose $Y_3$ has an antiholomorphic involution: an isometry $\gth: Y_3 \to Y_3$ such that $\gth^2=\Id$ and $\gth^\star J=-J$ with $J$ the complex structure. The effect on the \Kh form will then be $\gth^\star K=-K$, see \R{khform1}. The fixed points Fix$(\gth)$ of $\gth$ constitute a special Lagrangian submanifold\footnote{For more information on special Lagrangian manifolds and calibrated geometry see \cite{Joyce:2001xt, Gauntlett:2003di}. In \cite{Harvey:1982xk} it was shown that the two conditions (\ref{eq:condi1}, \ref{eq:condi2}) are actually equivalent.} and the pullback of the \Kh form behaves as
418: \equ{
419: K\big|_{\textrm{Fix}(\gth)} =\gth^\star K\big|_{\textrm{Fix}(\gth)}=-K\big|_{\textrm{Fix}(\gth)} \nn \km
420: }
421: which holds if $K\big|_{\textrm{Fix}(\gth)}=0$. Note that this (discrete) isometry is nothing but complex conjugation: interchanging barred and unbarred coordinates
422: \equ{
423: \gth: X^m\to X^\bm \nn \pt
424: }
425: We see from \R{compls} that (the action of) $J$ indeed reverses sign under this isometry. The fixed points of $\gth$ are then the real submanifolds, specifically the real $3$-cycles $\cA$ and $\cB$. \\
426:
427: The second condition \R{condi2}, which states that the pullback of $\gO$ must be proportional to the volume form of $\cC_3$, is also satisfied for $\cA$ and $\cB$. From equation \R{tc5c} we see that the pullback of $\gO$ onto cycle $\cA$ or $\cB$ is either (proportional to) $\ga$ or $\gb$ respectively. Equation \R{volf1} states that the volume form on $Y_3$ is given by $\ga\wedge \gb$. The Riemann bilinear identity
428: \equ{
429: \int_{Y_3}\gPs\wedge\gX=\sum_a \left(\int_{\cA^a}\gPs\int_{\cB_a}\gX -\int_{\cB_a}\gPs\int_{\cA^a}\gX\right) \nn \km
430: }
431: for arbitrary $3$-forms $\gPs$ and $\gX$, can be applied to the volume integral giving
432: \equ{
433: \textrm{Vol}(Y_3)=\int_{Y_3}\ga\wedge\gb=\int_{\cA}\ga\int_\cB\gb \nn\pt
434: }
435: Thus $\ga$ and $\gb$ correspond to the induced volume forms on $\cA$ and $\cB$ respectively. We can therefore conclude that $\cA$ and $\cB$ are the supersymmetric cycles.\\
436: Suppose we take a linear combination of the cycles $\cA$ and $\cB$:
437: \equ{
438: \cC_3=m\cA+n\cB \label{eq:susyc1} \km
439: }
440: with $m$ and $n$ integers, then this is not a supersymmetric cycle. The reasoning is as follows. We concluded that the (induced) volume forms on $\cA$ and $\cB$ are $\ga$ and $\gb$ respectively. The volume of $m$ times $\cA$ then becomes
441: \equ{
442: \int_{m\cA}\ga=m\int_\cA \ga=m \nn \km
443: }
444: using \R{c5b}, similarly for $n\cB$. Together they determine the volume of the $\cC_3$ cycle to be $m+n$. On the other hand, integrating (the pullback of) $\gO$ gives
445: \equ{
446: \int_{\cC_3}\gO=m-\I n \label{eq:OonC}
447: }
448: which is not proportional to $m+n$. The two are only proportional if either $m$ or $n$ is zero.
449: Consequently, a membrane wrapping $\cA$ and $\cB$ simultaneously is not a supersymmetric configuration and will not contribute to $\cR_{IJKL}$. The membrane needs to be wrapped either on $\cA$ or on $\cB$ to be supersymmetric.\\
450:
451: Having identified the supersymmetric cycles, we can evaluate \R{535b} and \R{corrs1}.
452: The integral over $\gO$ in $S_\inst$ \R{535b} is easily evaluated using \R{OonC}
453: \equ{
454: \left|\int_{\cC_3}\gO\right|=\sqrt{m^2+n^2} \label{eq:go1} \pt
455: }
456: To do the integral of $A_3$ over $\cC_3$ in \R{535b} we must expand the $3$-form potential $A_3$ as
457: \equ{
458: A_3=C_3 + v\ga +u\gb \km\label{eq:spacetimec1}
459: }
460: where $C_3$ is the spacetime $3$-form potential dual to $e_0$. $v$ and $u$ are functions of the four dimensional spacetime coordinates, we left out the vector multiplet terms. This is the same expansion as in \R{3form1e} (see also the discussion on page \pageref{hereI}). Evaluating this $A_3$ on the general cycle $\cC_3$, see \R{susyc1}, results in
461: \equ{
462: \I\int_{\cC_3}A_3=\I mv-\I nu \label{eq:temp5ac} \pt
463: }
464: In calculating $\cK$, see \R{Kpot1}, we must be a bit careful. The $\cK_H$ part is easy, using \R{volf1} and \R{volf2} we obtain
465: \equ{
466: \I\int_{Y_3}\gO\wedge\bgO=2 \label{eq:eqi}\pt
467: }
468: To compute $\cK_V$ we use a different (but equivalent) formula for the volume of a Calabi-Yau manifold:
469: \equ{
470: \textrm{Vol}(Y_3)=\frac{1}{3!}\int_{Y_3} K\wedge K\wedge K \nn\pt
471: }
472: This formula is not directly applicable because we have to compactify\footnote{We could have done this in an earlier stage, but it does not matter for the decomposition \R{spacetimec1} since that is with respect to the Calabi-Yau. This is why we left out the hat on $A_3$ from the beginning, compare also with \R{hgh1}.} one dimension, since we are still in $5$ dimensions and we would like to compute the corrections to the geometry of the UHM in $4$ spacetime dimensions. In the string frame this is achieved (compare with \R{stf1}) by decomposing the $10$-dimensional metric as
473: \equ{
474: ds^2_{11}=e^{-2\gf/3}(dx_{11}+A_m dx^m) + e^{\gf/3}ds^2_{10} \label{eq:compa1} \km
475: }
476: where the $10$-dimensional part incorporates $Y_3$. For the \Kh form it has the consequence that it picks up a dilaton dependence:
477: \equ{
478: K=e^{\gf/3}K_{\textrm{s.f.}} \nn \km
479: }
480: where $K_{\textrm{s.f.}}$ is the \Kh form in the string frame. This means that the we can write
481: \equ{
482: \frac{4}{3}\int_{Y_3}K\wedge K\wedge K=8\textrm{Vol}(Y_3)e^\gf=8e^\gf \nn\km
483: }
484: as measured in the string frame. Using this result and \R{eqi} we calculate
485: \equ{
486: e^{-\cK}=2e^{\gf/2} \nn \pt
487: }
488: The total result for $S_\inst$ using the PT variable $r$ can thus be written as
489: \equ{
490: e^{-S_\inst}=e^{-2\sqrt{m^2+n^2}\sqrt{r}}e^{-\I mv+\I nu} \label{eq:stinsta1} \pt
491: }
492: Observe that we really can interpret the $m$ and $n$ as instanton charges. Keep in mind though that for a supersymmetric configuration only one instanton can be switched on at a time, as we already anticipated in section \ref{mem1}.
493: Comparing the result for $S_\inst$ in \R{stinsta1}, which originated from \R{535b}, with the supergravity result \R{act1}, we find agreement. In \R{stinsta1} we should consider only $m$ or $n$ unequal to zero. Similarly, in \R{act1} we have either the membrane charge due to dualizing with respect to $\gvf$ or to $\gc$. We see that in the string picture the charges ($n$ and $m$) are quantized, i.e., they are integers. Furthermore, not only is the $g_s$($=\sqrt{r}$) behaviour correct, but the numerical factor of $2$ matches as well.\\
494: Lastly there is the tensorial structure of \R{corrs1}. The supersymmetric cycle is either $\cA$ or $\cB$ which means that the $d^I$ ($I=1,2$) should all be equal, $\cA$ giving corrections to $\cR_{1111}$ and $\cB$ to $\cR_{2222}$. \\
495: Summarizing, we can write for the corrections to $\cR$ due to the two supersymmetric cycles:
496: \bea
497: && \gD_\cA \cR_{1111}=N e^{-2|m| \sqrt{r}}e^{-\I mv} \nn \\[1.5mm]
498: && \gD_\cB \cR_{2222}=N e^{-2|n| \sqrt{r}}e^{+\I nu} \label{eq:Rres1} \km
499: \eea
500: the other combinations are zero.
501: In the next subsection we will make the comparison between the string theory result and the results in the PT framework more precise.
502:
503:
504:
505: \subsubsection{A note on the spinors}
506: \label{spinors}
507: In \R{corf4a} the correlator is written down which corrects the tensor $\cR$ as in \R{corrs1}. This expression is valid for five spacetime dimensions, with Lorentzian signature. We have evaluated \R{corrs1} giving the result \R{Rres1}. In the process we went down to four spacetime dimensions, by compactifying on a circle \R{compa1}. In five spacetime dimensions the $2n$ spinors $\gc^I$ are \emph{symplectic-Majorana} spinors. The point is that in four space and one time dimension, one cannot impose a Majorana condition. But if one has an even number of Dirac spinors $\gc^I$, one can impose a reality condition:
508: \equ{
509: \gc^{I\star}=\gO_{IJ}\cB \gc^J \label{eq:temp5a} \km
510: }
511: with $I=1,\ldots,2n$. The matrix $\gO$ is real and satisfies
512: \equ{
513: \gO_{IJ}\,\gO_{JK}=-\gd_{IK} \nn \pt
514: }
515: Equation \R{temp5a} is consistent if $\cB^\star \cB=-1$. The $\cB$ matrix has to do with the fact that if the set $\{\gG^a\}$ represents the Clifford algebra, in a certain number of dimensions with a certain signature, then $\{\pm\gG^{a\star}\}$ does so as well. The $\cB$ matrices then provide a similarity transformation relating the $\gG^a$ and $\gG^{a\star}$
516: \equ{
517: \cB_{\pm}\gG^a\cB_\pm^{-1}=\pm \gG^{a\star} \nn \pt
518: }
519: The existence of the $\cB_\pm$ depends on the number of dimensions and the signature. In five spacetime dimensions they do exist and obey $\cB^\star_\pm \cB_\pm=-1$, so we can either use $\cB_-$ or $\cB_+$ in \R{temp5a}, which is why we denoted it by $\cB$.
520: The $\cB_\pm$ matrices are related to the charge- and the Dirac conjugation matrix. For a nice account see \cite{Ortin:2004ms}.\\
521: If one then goes to four spacetime dimensions, one can impose the normal Majorana condition, which reads in this notation
522: \equ{
523: \gc^{I\star}=\cB\gc^I \nn \pt
524: }
525: The reduction of the spinors can be obtained by the following identification
526: \equ{
527: \bgl^\ba \sim (1+\gg_5)\gc^I \qquad \gl^a\sim (1-\gg_5)(\gO\gc)^I \nn \km
528: }
529: where $a,\ba=1,\ldots,2n$ as in chapter \ref{ch3}. With this notation we mean to indicate that a single $\gc^I$ gives two spinors: $\gl^a$ and $\bgl^\ba$. The projection operators $(1\pm\gg_5)$ thus give us chiral spinors. Because under complex conjugation the indices are interchanged we indeed have a system of $2n$ Majorana spinors. For more information on the relation between the geometry of hypermultiplets in five and four spacetime dimensions and the reduction, see \cite{Rosseel:2004fa}.
530:
531:
532:
533: \subsection{Comparing to the PT metric}
534: We explicitly constructed the corrections to the completely symmetric tensor (see \R{Rres1}) and we must compare this with the instanton corrections to a completely symmetric tensor in the four-dimensional $N=2$ supergravity theory.
535: This tensor has been constructed in \cite{Bergshoeff:2004nf, Bergshoeff:2002qk}. The symmetric tensor $\cW_{abcd}$ (where the indices are $Sp(2n,\mR)$ indices, with $n=1$ in our case) can be obtained from the curvature decomposition
536: \begin{equation} \label{eq:r.1}
537: R_{\hA\hB\hC\hD} = \nu (R^{SU(2)})_{\hA\hB\hC\hD} + \frac{1}{2}\, {L_{\hD\hC}}^{a
538: b}\, \cW_{abcd}\, {L_{\hA\hB}}^{cd}\pt
539: \end{equation}
540: In our conventions, see appendix \ref{geopt}, we have $\gn=-1/2$.
541: $R_{\hA\hB\hC\hD}$ is the curvature tensor of the quaternionic manifold, so $\hA,\ldots,\hD=1,\ldots,4n$ and the $SU(2)$ part of the tensor is given by
542: \begin{equation}
543: (R^{SU(2)})_{\hA\hB\hC\hD} = \frac{1}{2}\, g_{\hD [\hA}\, g_{\hB ]\hC} + \frac{1}{2}\,
544: K^r_{\hA\hB}\, K^r_{\hD\hC} - \frac{1}{2}\, K^r_{\hD[\hA}\,
545: K_{\hB ]\hC}^r\km \nn
546: \end{equation}
547: with
548: \begin{equation}
549: {L_{\hA\hB a}}^b = V_{\hA ia}\, \bar{V}^{ib}_\hB\pt
550: \end{equation}
551: We can solve \R{r.1} for $\cW_{abcd}$ by using the inverse relation for ${L_{\hA\hB}}^{ab}$:
552: \begin{equation}
553: - \frac{1}{2}\, V^{i\hB}_c\, V^{\hA}_{id}\, {L_{\hA\hB}}^{a
554: b} = \gd^a_c \, \delta^b_d\pt \nn
555: \end{equation}
556: The resulting expression for $\cW_{abcd}$ then reads:
557: \begin{equation}
558: \cW_{abcd} = \frac{1}{2}\, \gve^{ij}
559: \gve^{kl}\, V_{id}^\hA\, V_{jc}^\hB\, V_{kb}^\hD\, V_{l
560: a}^\hC \big( R_{\hA\hB\hC\hD} - \nu (R^{SU(2)})_{\hA\hB\hC\hD} \big)\pt
561: \end{equation}
562: Starting from the general PT metric \R{tod1} we can compute the rather complicated curvature tensor and by using our expressions for the vielbeins and complex structures, see appendix \ref{geopt}, we can compute $\cW_{abcd}$. The result is
563: \begin{align} \label{r.2}
564: \cW_{1111} & = 4 r^2 f^{-3} e^{-h} \big[ f (\partial^2_{\bz} f - \partial_{\bz}
565: h\, \partial_{\bz} f) - 3 (\partial_{\bz} f)^2 \big] \notag \\[2pt]
566: \cW_{2111} & = r^2 f^{-3} e^{-h/2} \big[ 2 f\, \partial_r \partial_{\bz} f - 3
567: (f\, \partial_r h + 2 \partial_r f)\, \partial_{\bz} f + f^2 \partial_r \partial_{\bz} h \big]
568: \notag \\[2pt]
569: \cW_{2211} & = - r^2 f^{-3} \big[ f \big( r \partial^3_r h - (\partial_r h)^2
570: \big) - 4 e^{-h}\, \partial_z f\, \partial_{\bz} f + 2 (\partial_r f)^2 \big]
571: \notag \\[2pt]
572: \cW_{2221} & = - r^2 f^{-3} e^{-h/2} \big[ 2 f\, \partial_r \partial_z f - 3
573: (f\, \partial_r h + 2 \partial_r f)\, \partial_z f + f^2 \partial_r \partial_z h \big] \notag
574: \\[2pt]
575: \cW_{2222} & = 4 r^2 f^{-3} e^{-h} \big[ f (\partial^2_z f - \partial_z h\, \partial_z
576: f) - 3 (\partial_z f)^2 \big]\pt
577: \end{align}
578: We have introduced the complex variable $z\equiv u+\I v$ to keep the expressions compact\footnote{The fact that we have such a compact expression is absolutely delightful given the \emph{very} unwieldy result for the curvature tensor, which we will not give.}. Note that $\cW_{1111}=\cW_{2222}^\star$, $\cW_{2111}=-\cW_{1222}^\star$ and $\cW_{2211}$ is real.\\
579: Given this general form of $\cW_{abcd}$ we compute its instanton contributions. We find that at the perturbative level (the classical plus $1$-loop correction but no instantons) specified by \R{pertu1}, the only nonvanishing component is given by
580: \begin{equation}
581: \cW_{2211} = - \frac{r^3}{(r+2c)^3}\pt \nn
582: \end{equation}
583: As for the leading order instanton corrections: the $1$-instanton sector up to $1$-loop, specified by \R{eh-exp} and \R{finst}, gives the following contributions to $\cW_{abcd}$
584: \begin{align} \label{eq:r.3}
585: \Delta \cW_{1111} & = \hN\, \big[ A_1\, e^{\I v} + A_1^\star\, e^{-\I v}
586: - B_1\, e^{-\I u} - B_1^\star\, e^{\I u} \big] \notag \\[2pt]
587: \Delta \cW_{1112} & = \hN\, \big[ A_1\, e^{\I v} - A_1^\star\, e^{-\I v}
588: + \I (B_1\, e^{-\I u} - B_1^\star\, e^{\I u}) \big] \notag \\[2pt]
589: \Delta \cW_{1122} & = \hN\, \big[ A_1\, e^{\I v} + A_1^\star\, e^{-\I v}
590: + B_1\, e^{-\I u} + B_1^\star\, e^{\I u} \big] \notag \\[2pt]
591: \Delta \cW_{1222} & =\hN\, \big[ A_1\, e^{\I v} - A_1^\star\, e^{-\I v}
592: - \I (B_1\, e^{-\I u} - B_1^\star\, e^{\I u}) \big] \notag \\[2pt]
593: \Delta \cW_{2222} & = \hN\, \big[ A_1\, e^{\I v} + A_1^\star\, e^{-\I v}
594: - B_1\, e^{-\I u} - B_1^\star\, e^{\I u} \big]\pt
595: \end{align}
596: We have defined $\hN\equiv (r+c)^{(1-m_1)/2} e^{-2\sqrt{r+c}}$. Comparing the $r$-dependence of $\hN$ with the $r$-dependence of $N$ in \R{Rres1} then fixes the value of $m_1$. In particular, if $N$ is $r$-independent $m_1=1$, in order to match $\hN$ to $N$.\\
597: Let us repeat \R{Rres1} for the case of the the single instanton:
598: \equ{
599: \gD_{\cA}\cR_{1111}=N e^{-2\sqrt{r}}e^{-\I v}
600: \qquad \gD_{\cB}\cR_{2222}=N e^{-2\sqrt{r}}e^{+\I u} \label{eq:Rres1b} \pt
601: }
602: This indicates that the integration coefficient $A_1$ must be associated to the $\cA$-cycle and $B_1$ to the $\cB$-cycle. \\
603: Yet \R{r.3} seems very different from the string result \R{Rres1b}.
604: However, the latter result expresses something different: it is the result from \emph{one} membrane wrapping \emph{one} supersymmetric cycle, either $\cA$ or $\cB$. In contrast, \R{r.3} includes both types of instantons and anti-instantons. Moreover, the fermion frame used by \cite{Becker:1995kb} is different from ours. So result \R{Rres1b} is in a different frame from ours. Still, these frames can differ at most by a local $SU(2)$ ($Sp(2,\mR)$) rotation.
605: The rotation group has to be compatible with the
606: reality condition imposed on the pair of symplectic Majorana spinors
607: coupling to $\cW_{abcd}$ and preserve fermion bilinears. These conditions then
608: lead to the fact that the most general transformation is given by $SU(2)$, this is discussed in \cite{Cortes:2003zd}.
609: We can parametrize such a $SU(2)$ rotation by
610: \begin{equation}
611: U = \lp e^{\I\xi} \cos\eta & e^{\I\rho} \sin\eta \\[1mm]
612: -e^{-\I\rho} \sin\eta & e^{-\I\xi} \cos\eta \rp\km \nn
613: \end{equation}
614: where the parameters $\eta$, $\xi$, $\rho$ can depend on the scalars, although we will find that this will not be necessary.
615: Consider the contribution to $\gD\cW$ arising only from the $\cB$-instanton. This requires setting $A_1=0$. Upon performing a global $SU(2)$ rotation of the fermion frame with parameters $\get=\gp/4$ and $\gx=\gr+\gp/2$ we obtain
616: \equ{
617: \gD\tilde{\cW}_{1111}=-4 \hN B^\star_1 e^{\I u} \qquad \gD \tilde{\cW}_{2222}=-4 \hN B_1 e^{-\I u} \km \label{eq:trW1}
618: }
619: with the other components vanishing identically. Note that the remaining free parameter $\gr$ in the transformation only induces a phase on the components of $\gD \tilde{\cW}_{abcd}$. We can set it to zero for convenience. This reproduces\footnote{Take $u\to -u$, which corresponds to an orientation reversal of the cycle, see \R{temp5ac}.} \R{Rres1b}, which is due to one instanton without anti-instantons. The correction \R{trW1} incorporates both the instanton and anti-instanton, which necessarily appear in different sectors of the theory\footnote{Just as in the case of the NS $5$-brane for which the zero modes in the fermions due to the $5$-brane instanton were in a different sector than those due to the $5$-brane anti-instanton see \R{brfermions1}.}.\\
620: Likewise, we can consider the contribution of the $\cA$-instanton by setting $B_1=0$. In this case the transformation $\get=\gp/4$ and $\gx=\gr=0$ leads to the correction
621: \equ{
622: \gD \tilde{\cW}_{1111}=4\hN A_1 e^{\I v} \qquad \gD \tilde{\cW}_{2222}=4 \hN A^\star_1 e^{-\I v} \label{eq:trW2} \km
623: }
624: with the other components vanishing identically. This is again of the form predicted by string theory \R{Rres1b}, although in a different fermionic frame from \R{trW1}. Summing these two corrections involves rotating some of the contributions into the proper fermionic frame. Hence, our result precisely agrees with the one obtained in \cite{Becker:1995kb}.\\
625: In order to sum the two contributions, we go to the $\cB$-instanton frame, i.e., the frame in which we obtained \R{trW1}, but now we also include $A_1$. The corrections to the $\cW$-tensor are then given by
626: \bea
627: && \gD \tilde{\cW}_{1111}=-\hN\left(A_1 e^{\I v} + A^\star_1 e^{-\I v} + 4 B^\star_1 e^{\I u}\right) \nn \\[1mm]
628: && \gD \tilde{\cW}_{1112}=\hN\left(A_1 e^{\I v} - A^\star_1 e^{-\I v}\right) \nn \\[1mm]
629: && \gD \tilde{\cW}_{1122}=-\hN\left(A_1 e^{\I v} + A^\star_1 e^{-\I v}\right) \nn \\[1mm]
630: && \gD \tilde{\cW}_{1222}=\hN\left(A_1 e^{\I v} - A^\star_1 e^{-\I v}\right) \nn \\[1mm]
631: && \gD \tilde{\cW}_{2222}=-\hN\left(A_1 e^{\I v} + A^\star_1 e^{-\I v} + 4 B^\star_1 e^{\I u}\right) \label{eq:trW4} \pt
632: \eea
633: This result implies that the four fermionic zero modes $\gps_1, \gps_2$ arising from a membrane wrapping the $\cA$ and the $\cB$ cycle, respectively, are not orthogonal.
634: %If they were orthogonal $\gD \tilde{\cW}_{1111}$ would only depend on both $u$ and $v$.
635: If in the $\cB$-frame we denote the two zero modes giving rise to the $e^{\I u}$ corrections by $\gps_1$, then the corrections proportional to $e^{\I v}$ arise from the zero modes $\gps_1-\gps_2$. This zero mode configuration produces all the signs appearing in \R{trW4}, since the $\cA$-anti-instanton has its zero modes in $\gps_1+\gps_2$.
636:
637:
638:
639: \section{Compactification with fluxes}
640: \label{fluxes}
641: In chapter \ref{ch2} we sketched the compactification from $10$ to $4$ dimensions, \emph{without} fluxes. However, in a compactification also more complicated field strengths can be considered. This introduces more parameters and is referred to as compactification \emph{with} fluxes.\\
642: Ideally one would like to compactify while turning on all possible fluxes, because this gives the most general effective action in four dimensions. It is a rather extended subject and we will refer to the literature for more information, specifically \cite{Louis:2002ny, Kachru:2004jr, Grana:2005jc} and references therein.\\
643: We will illustrate the mechanism by considering again the $2$-form potential from section \ref{fourd}. In \R{Bexp1} the $10$-dimensional $2$-form $\hB_2$ was expanded as $\hB_2=B_2 + b^i \go_i$, where the $\go_i$ are $(1,1)$-forms and the $b^i$ scalar fields in four dimensions. The corresponding field strength is defined as
644: \equ{
645: H_3\equiv d B_2 + db^i \go_i \nn \pt
646: }
647: Remember that $d\go_i=0$, so this quantity obeys the Bianchi identity $dH_3=0$. More generally we can include fluxes in the definition of the field strength:
648: \equ{
649: H_3=dB_2 + db^i \go_i + H_3^{\mathrm{flux}} \label{eq:flux1} \km
650: }
651: where $H_3^{\mathrm{{flux}}}$ is a $3$-form on the Calabi-Yau $3$-fold
652: \equ{
653: H_3^{\mathrm{flux}} \equiv p^A \ga_A + q_A \gb^A \nn
654: }
655: and the $(\ga_A, \gb^A)$ again form a real basis of $H^3(Y_3)$. This gives rise to extra flux parameters $p^A$ and $q_A$ which have to be constant. They are obtained by integrating $H_3^{\mathrm{flux}}$ over the $3$-cycles of the Calabi-Yau. In the $4$-dimensional effective action these parameters gauge certain fields, namely the scalars $\gx^a$ and $\tgx_a$ from the hypermultiplets \R{hyperm1} with $a=1,\ldots,h^{(1,2)}$. Concretely, the action will now contain terms
656: \equ{
657: D_\gm \gx^a\equiv \p{\gm}\gx^a -p^a A_\gm \qquad D_\gm \tgx^a\equiv \p{\gm}\tgx^a -q^a A_\gm \km \nn
658: }
659: where $A_\gm$ is the graviphoton, similarly for the axion (see \cite{Louis:2002ny}). In addition there will be a potential containing these parameters. The parameters gauge certain isometries of the scalar manifold \R{scm1}. The scalar manifold\footnote{We will ignore the vector multiplets, and thus $\cM_V$, for the moment.} possesses a number of isometries. For instance, we have seen in section \ref{instsol} that the quaternionic target space of the classical (and perturbative) UHM has a number of shift symmetries acting on the scalars, see \R{heis1}. These are collectively denoted by
660: \equ{
661: \gd \gf^\hA=\gL^I k^\hA_I(\gf) \nn \km
662: }
663: where the $\gf^\hA$ are the $4$ scalars in the UHM\footnote{The generalization to more hypermultiplets is straightforward.}. $k^\hA_I$ are the Killing vectors and $\gL^I$ their parameters, the $I$ denotes the various isometries. For example, the action of the $\ga$ shift on $\gs$ in \R{heis1} can be written as
664: \equ{
665: \gd \gs=\gL^I k_I^\gs =\gL^\ga k_\ga^\gs \nn \km
666: }
667: with $\gL^\ga=-\ga$ and $k^\gs_\ga=1$. \\
668: To gauge such isometries one introduces gauge covariant derivatives
669: \equ{
670: D_\gm \gf^\hA \equiv \p{\gm}\gf^\hA-k^\hA_I(\gf) A^I_\gm \label{eq:gaugecov1} \km
671: }
672: where $A^I_\gm$ are the vectors from the vector multiplets plus the graviphoton. In addition one has to gauge the various connections, which give rise to gauged curvatures. Furthermore, a potential has to be added to the action in order to preserve supersymmetry.\\
673: The part of the potential which involves quaternionic quantities is of the form
674: \equ{
675: V=e^{\wK} X^I\bX^J\left(2\gk^{-2} G_{\hA\hB} k_I^\hA k_J^\hB -3 P_I^r P_J^r\right) \label{eq:potI} \pt
676: }
677: The \Kh potential $\wK$ is given by \R{kahlerpot1} and the $X^I$ ($I=0,\ldots,n_v$ with $n_v$ the number of vector multiplets) are holomorphic functions of the scalars\footnote{These scalars are defined as $t^i\equiv b^i+\I v^i$, see the footnote on page \pageref{spk2}.} $t^i$ in the vector multiplets, precisely the sections from \R{Osection1}.
678: %The scalars of the vector multiplets parametrize a special \Kh manifold \R{scm1}.
679: The Killing vectors $k^\hA_I(\gf)$ satisfy the Killing equations which in $N=2$ supergravity can be solved in terms of a triplet of Killing prepotentials $P^r_I$, $r=1,2,3$, also called `moment maps'. See appendix \ref{geopt} for more information on the potential and moment maps.\\
680:
681: By sketching the effect of turning on a flux in $H_3$, \R{flux1}, we have introduced two different topics: compactification with fluxes and gauged supergravity. The reason for introducing the latter is that the effect including fluxes has on the $4$-dimensional effective theory, can be reproduced by gauging the $4$-dimensional \sugra theory one obtains by compactification without fluxes. This means that the powerful machinery of gauged supergravity can be used to study the effect of fluxes in the internal Calabi-Yau manifold on the effective theory. For more details on $N=2$ gauged supergravity we refer to \cite{D'Auria:1990fj, Andrianopoli:1996cm, Ceresole:2000jd, deWit:2001bk} and references therein.\\
682: There are a few points that must be made. In gauged supergravity, the hypermultiplet scalars can acquire charges under vectors, as in \R{gaugecov1}. This means, contrary to the ungauged case, that the hypermultiplets and vector multiplets no longer decouple. However, the structure of the Lagrangian is encoded in the same data as before the gauging with the addition of the Killing vectors encoding the isometries of the scalar manifold. In spite of the presence of a potential, no mixing of the metrics is allowed. Therefore the metric of the vector multiplets still receives no string loop corrections. \\
683: Furthermore, the presence of fluxes in general distorts geometry of the Calabi-Yau causing it to change into something else. In such cases the appropriate internal manifold is a generalized Calabi-Yau manifold, see \cite{Behrndt:2005vi, Louis:2005te}.\\
684:
685: It turns out that the case we are considering corresponds to a very simple type of flux compactification. Remember that we are in a situation in which membranes break all the continuous isometries of the UHM save one. The $4$-dimensional quaternionic scalar manifold of the UHM only has the shift in the axion $\gs$ left, or in PT coordinates $t$. As demonstrated in \cite{Louis:2002ny}, the presence of a nontrivial spacetime part in the $3$-form $A_3$ gauges this shift. In other words, the presence\footnote{Note that in \cite{Bodner:1990zm} the spacetime filling $C_3$ was set to zero.} of $C_3$ gauges the $\ga$-shift in $t$, see \R{fmap1} and \R{spacetimec1}.\\
686: The reason is that the field strength $d C_3$ of a $3$-form in four spacetime dimensions is dual to a $0$-form field strength and carries no local degrees of freedom. One can therefore eliminate it from the action by dualizing it to a constant. We will denote this constant by $e_0$ (see also \cite{Kachru:2004jr}). In the $4$-dimensional action the derivative of the axion will then become a covariant derivative:
687: \equ{
688: \p{\gm}\gs\to D_\gm \gs\equiv \p{\gm}\gs + 2e_0 A_\gm \nn \km
689: }
690: where $A_\gm$ is again the graviphoton. Comparing to \R{gaugecov1} we see that $k^\hA=-2 e_0 \gd^{\hA\gs}$, where we have left out the index $I$ because we are considering just one isometry. The precise form of this Killing vector is not important, as we shall see, and is taken equal to $e_0 \gd^{\hA\gs}$. Since we will only be gauging this one isometry, we only need one vector field and thus can suffice with the graviphoton. No vector multiplets need to be used.
691: In the $10$-dimensional supergravity theory we have the field strength corresponding to $A_3$: $F_4=dA_3$, see \R{3form1e}. This is (Hodge) dual to a $6$-form field strength $F_6=\!\! \,^\star F_4$. Consequently, the nontrivial spacetime part $dC_3$ is equivalent to turning on $F_6$ flux in the Calabi-Yau manifold. Conversely one could say that turning on $F_6$ flux can be dealt with after reducing on the Calabi-Yau by considering a nontrivial spacetime part $C_3$. This means that the $F_6$ flux does not backreact on the internal Calabi-Yau geometry. \\
692: In string theory the inclusion of fluxes can lead to tadpoles, but it was shown in \cite{Kachru:2004jr} that only by turning on $F_0$ and $H_3$ flux simultaneously in a type IIA compactification tadpoles can arise, so that is not an issue here.
693:
694: The potential \R{potI} simplifies considerably in our case. As we are considering the UHM with only one isometry, we have only nontrivial components of $P^r_I$ for $I=0$. Furthermore, there are no vector multiplets present and the potential simplifies to
695: \equ{
696: V=4G_{\hA\hB} k^\hA k^\hB -3P^r P^r \nn \km
697: }
698: as in appendix \ref{geopt}. The labels $\hA,\hB$ run over the PT coordinates. Substituting the results for the the moment maps \R{moment2} and the Killing vector \R{Kv2} we obtain
699: \equ{
700: V=\frac{1}{r^2}\left(\frac{4}{f} -3\right)e_0^2 \label{eq:potA} \pt
701: }
702: We see that the constant $e_0$ which multiplies the Killing vector appears as an overall factor, hence we will set it to one.
703:
704: \section{The potential}
705: We are now in a position to study the effect of membranes on the $4$-dimensional theory in a very concrete manner: through the potential. This potential will consist of three parts:
706: \equ{
707: V=V_{\class} +V_\text{1-loop} + V_\textrm{inst} \label{eq:Vtot} \pt
708: }
709: Using \R{pertu1} we calculate
710: \equ{
711: V_\class=\frac{1}{r^2} \qquad V_\textrm{1-loop}=-\frac{4c}{r^2(r+2c)} \nn \km
712: }
713: we will discuss $V_\textrm{inst}$ shortly, first we shall discuss $V_\class$ and $V_\text{1-loop}$.
714:
715: \subsubsection{The perturbative potential}
716: $V_\class$ is a positive monotonically decreasing function with a typical `runaway' behaviour in $r$. For $r\to 0$ or $g_s \to \infty$ (remember that $g_s =1/\sqrt{r}$) it diverges. $V_\class$ is displayed in figure \ref{eins}, it has no vacua except the trivial one at $r=\infty$.
717: \begin{figure}[t]
718: \renewcommand{\baselinestretch}{1}
719: \epsfxsize=0.48\textwidth
720: \begin{center}
721: \leavevmode
722: %\epsfxsize=0.37\textwidth
723: \epsffile{Vclass1.eps} \,
724: \epsfxsize=0.48\textwidth
725: \epsffile{Vpert1c.eps}
726: \end{center}
727: \parbox[c]{\textwidth}{\caption{\label{eins}{\footnotesize The scalar potential $V_\class$ (left) and
728: $V_\class + V_\textrm{1-loop}$ (right) for $c=-10$. Including the $1$-loop correction
729: with $c<0$ leads to a pole at $r=-2c$.}}}
730: \end{figure}\\
731: If we add the $1$-loop contribution the behaviour changes drastically. In figure \ref{eins} $V_\class + V_\textrm{1-loop}$ is displayed for the `generic' value $c=-10$. For different values of $c$ the location of the pole shifts but the (qualitative) properties of the potential remain the same. We found in \R{c2} that $c$ is always smaller than zero. Since the region $0<r<|2c|$ has to be discarded, see the discussion below \R{c2}, we also have to discard this region of the potential. For $r>|2c|$ the behaviour of the potential is similar to the behaviour of $V_\class$. \\
732: These two pieces of the potential are independent of $(u,v,t)$ which means that these scalars parametrize flat directions. The absence of $u$ and $v$ is just a consequence of their absence in \R{pertu1}. The fact that the potential does not depend on $t$ is a different matter. Because we have gauged the isometry in $t$, gauge invariance implies that the potential cannot depend on $t$.
733: Consequently we only have to stabilize $u$ and $v$ in the potential to lift the flat directions, since $t$ is stabilized no matter what. We now demonstrate that we can stabilize $u$ and $v$ by taking into account the membrane contributions.
734:
735: \subsubsection{The potential including membrane effects}
736: $V_\textrm{inst}$ is the part containing the contributions from the infinite instanton expansion. If we limit ourselves to small string coupling, that is $r\mg 1$, a good approximation is given by (\ref{eq:eh-exp}, \ref{eq:finst}).
737: We will consider the leading contributions coming from the one-instanton sector. Furthermore, we will consider the effect of the $v$-instanton sector while switching off the $u$-instanton sector. We can do this because both sectors enter separately in exactly the same manner in the potential, as can be seen from (\ref{eq:9.1}, \ref{eq:eh-exp}, \ref{eq:finst}). Therefore we can first study the effect of one sector and then draw similar conclusions for the other sector. \\
738: To leading order the instanton part of the potential becomes
739: \equ{
740: \label{eq:8.5}
741: V_\textrm{1-inst} = -4\, r^{-(m_1+5)/2}\, \big( \hat{A}_{1,m_1} \cos(v)
742: - \tilde{A}_{1,m_1} \sin(v) \big)\, e^{-2 \sqrt{r}}\km
743: }
744: where we have used \R{finst}. We have defined $A_{1,m_1}=\hat{A}_{1,m_1}+\I\tilde{A}_{1,m_1}$ and have set $B_{1,m_1}=0$. $V_\inst$ has no poles except at $r=0$. In fact, this
745: is an artefact of the expansion in \R{8.5}. If we would not have expanded the
746: denominator which contained $(r+2c)$, $V_\textrm{1-inst}$ also would develop a
747: singularity at $r=-2c$, which even dominates over the one in
748: $V_\textrm{1-loop}$, and the potential is no longer bounded from below.
749: Resolving this singularity presumably requires resumming the entire
750: instanton expansion to obtain expressions which are valid at small
751: values of $r\le -2c$. It is unclear how to resum the expansion and we will continue to work with the expanded
752: expressions \R{8.5}. Note that resolving singularities
753: by nonperturbative effects has been shown to work in the context of
754: the Coulomb branch of three-dimensional gauge theories with eight
755: supercharges \cite{Seiberg:1996nz, Dorey:1997ij, Dorey:1998kq, Chalmers:1996xh}. In these cases the moduli space is
756: hyperk\"ahler instead of quaternionic.\\
757: Furthermore, since $V_\textrm{inst}$ is exponentially suppressed for large $r$, the large $r$ behaviour of $V$ is dominated by $V_\class + V_\textrm{1-loop}$. This means that for $r\to\infty$ $V$ still approaches $0$ from above. The total potential is bounded from below and diverges to $+\infty$ for $r\downarrow |2c|$. Due to the transcendental nature of the potential we have to analyze the vacuum structure numerically. Furthermore, since we do not know anything about the $g_s$ dependence of the determinant in the background of instantons, contained in the factor $N$ in \R{corrs1}, we will choose $m_1=-2$, see \R{r.3}. This is of little importance since the qualitative behaviour of the potential is not very sensitive to this value. We can obtain similar results for different values of $m_1$.\\
758:
759: We can simplify this potential by imposing the symmetry which interchanges $v$-instantons with $v$-anti-instantons: $v\to -v$, $t\to -t$, see section \ref{sect4.3}. This means that we have to take $\tA_{1,m_1}=0$. Furthermore, we can take $\hA_{1,m_1}$ positive because negative values merely correspond to shifting $v\to v+\gp$. The local minimum in the $v$-direction is then located at $v=0$. Knowing the minimum in the $v$-direction, we can seek out the minimum in the $r$-direction. This depends on the value of the one remaining parameter $\hA_{1,m_1}$. For a given value of $c$ we can distinguish three different cases characterized by two values for $\hA_{1,m_1}$: $A_\textrm{min}$ and $A_\textrm{max}$. These have to be determined numerically for each value of $c$. If $\hA_{1,m_1} < A_\textrm{min}$ there is no local minimum in the $r$-direction except for the one at $r=\infty$. The (qualitative) picture is just that of $V_\class + V_\textrm{1-loop}$ in figure \ref{eins}. For $\hA_{1,m_1}>A_\textrm{max}$ we do find a minimum in the $r$-direction. The value of the potential at this point, let us call it $r_\textrm{AdS}$, is negative\footnote{In the next section we will explain the names for these minima.}. The most interesting situation occurs when $A_\textrm{min} < \hA_{1,m_1} < A_\textrm{max}$. In such a situation the minimum of the potential, occurring at $r_\textrm{dS}$, is positive, which is significant as we will learn in the next section. As in the previous case this stabilizes both the $v$ and the $r$ modulus. We are thus able to stabilize all the moduli of the UHM. This situation is depicted in figure \ref{zwei} for the value $c=-10$ and $\hA_{1,m_1}=9867$.
760: \begin{figure}[th]
761: \renewcommand{\baselinestretch}{1}
762: \epsfxsize=0.6\textwidth
763: \begin{center}
764: \leavevmode
765: \epsffile{Vinst1.eps} \,
766: \end{center}
767: \parbox[c]{\textwidth}{\caption{\label{zwei}{\footnotesize A view of $V(r,v)$ in the $r$-direction
768: along $v=0$.}}}
769: \end{figure}
770: Note that the only difference with the case $\hA_{1,m_1}>A_\textrm{max}$ is that the minimum is now positive.
771: The effect of increasing $\hA_{1,m_1}$ is to move the location of the minimum to the left: closer to the singularity. This means that the value for which the string coupling is smallest is obtained by taking $\hA_{1,m_1}=A_\textrm{min}$. In table \ref{t1} we give the relevant values for the parameters for the cases $c=-6/\gp \approx -1.9$ and $c=-10$.
772: \begin{table}
773: \centering
774: \begin{tabular}{|c|c|c|}
775: % after \\: \hline or \cline{col1-col2} \cline{col3-col4} ...
776: \hline
777: & $c = -1.9$ & $c = -10$ \\ \hline
778: $A_\textrm{min}$ & $53.0$ & $8180$ \\
779: $A_\textrm{max}$ & $60.8$ & $9900$ \\
780: $r_\mathrm{dS}$ & $7.4$ & $27.5$\\ \hline
781: \end{tabular}
782: \caption{{\footnotesize Illustrative values for
783: $A_\textrm{min}$, $A_\textrm{max}$, and $r_\textrm{dS}$ for the
784: $\wZ$-manifold $h^{(1,2)}=0$, $h^{(1,1)}=36$, $c\approx-1.9$) and
785: a fictional rigid CY$_3$ where $c=-10$, corresponding to $h^{(1,1)}=
786: \cO(100)$. For $A_\textrm{min}<\hat{A}_{1,m_1}<A_\textrm{max}$ the potential \R{Vtot}
787: has a meta-stable dS vacuum at which all hypermultiplet moduli are
788: stabilized.}}\label{t1}
789: \end{table}
790: The first case corresponds to the so-called $\wZ$-manifold (see for instance \cite{Candelas:1993nd}). The case $c=-10$ corresponds to a fictitious rigid CY $3$-fold with $h^{(1,1)}\approx \cO(100)$.\\
791: We see from table \ref{t1} that decreasing $|c|$ also decreases the values for $A_\textrm{min}$ and $A_\textrm{max}$ while their relative difference stays about the same. Decreasing $|c|$ also moves $r_\textrm{dS}$ closer to the singularity. This is a generic feature and does not depend on the particular values of $c$ in the table. \\
792:
793: To check that the minimum in the $r$-direction is also a minimum in the $v$-direction, we plot the total potential \R{Vtot} in the neighbourhood of the minimum, see figure \ref{drei}. To obtain this graph we used the values $c=-10$ and $\hA_{1,m_1}=9867$, as before.
794: \begin{figure}[th]
795: \renewcommand{\baselinestretch}{1}
796: \epsfxsize=1.0\textwidth
797: \begin{center}
798: \leavevmode
799: \epsffile{Vinst2gr1.eps} \,
800: \end{center}
801: \parbox[c]{\textwidth}{\caption{\label{drei} A detailed view of the meta-stable dS minimum of
802: the potential $V(r,v)$ in the $r$ and $v$ direction.}}
803: \end{figure}
804: If we would extend the graph in the $v$-direction we would see that it is periodical due to the cosine in \R{8.5}. This periodicity can be lifted by including higher order terms in $r$. One can check that by taking such higher order terms into account one can still construct a minimum with positive value. The values of the other minima (the `copies') are generically lowered to negative values.
805:
806:
807: \section{De Sitter space}
808: \label{Sitter}
809: De Sitter spacetime (dS) is a maximally symmetric solution to Einstein's equations with a positive cosmological constant $\gL$ (see \cite{Spradlin:2001pw, Hawking} for more details). We can describe $d$-dimensional de Sitter space as a hyperboloid embedded in flat $(d+1)$-dimensional Minkowski space $\cM^{d,1}$. If $\cM^{d,1}$ has coordinates $X^A, A=0, \ldots, d$ and metric $\get_{AB}=\textrm{diag}(-1,1,\ldots,1)$, the hypersurface
810: \equ{
811: \get_{AB}X^AX^B=R^2 \nn
812: }
813: defines a $d$-dimensional dS$_d$, with `radius' $R$. For the metric on dS$_d$ we can use the induced metric from the embedding space $\cM^{d,1}$. One can show that dS$_d$ is an Einstein space with positive scalar curvature and an Einstein tensor satisfying
814: \equ{
815: G_{\gm\gn} + \gL g_{\gm\gn}=0 \nn \km
816: }
817: where
818: \equ{
819: \gL=\frac{(d-2)(d-1)}{2R^2} \nn
820: }
821: is the cosmological constant. \\
822:
823: We have been considering an $N=2$ supergravity theory in four spacetime dimensions. We have discussed that gauging an isometry gives rise to a potential (the relevant terms are given in \R{app9t1a} and \R{app9t1b}). If we take the variational derivative with respect to $g_{\gm\gn}$ of the (hypermultiplet) action we find
824: \equ{
825: 0=\frac{1}{\sqrt{g}}\frac{\gd S}{\gd g_{\gm\gn}}=R^{\gm\gn}-\2 g^{\gm\gn} R + T^{\gm\gn} \nn \km
826: }
827: where $T^{\gm\gn}$ is the variational derivative of the part of the action containing the fields and the potential. This means that when considering the scalar fields at the minimum of the potential, $T^{\gm\gn}$ contributes to an effective cosmological constant term:
828: \equ{
829: g^{\gm\gn} V_{\textrm{min}}(r,u,v) \nn \pt
830: }
831: $V_{\textrm{min}}(r,u,v)$ is the constant value of the potential at its minimum (we denoted the fields as in the previous section).
832: When this value is larger than zero (if this is the only contribution to the vacuum energy) it can be associated to a de Sitter spacetime\footnote{Alternatively, if the minimum of the potential is less than zero it corresponds to an Anti-de Sitter space, hence the name $r_{\textrm{AdS}}$.}.\\
833: Supergravity cannot be realized in a de Sitter space. This means that the $N=2$ supergravity theory has no supersymmetry for a de Sitter vacuum (see e.g. \cite{Pilch:1984aw}).
834: We will not discuss the breaking of supersymmetry in the case of an AdS or Minkowski\footnote{This corresponds to a zero cosmological constant, so a vanishing value of the potential at its minimum.} vacuum. There is much more to say about de Sitter vacua in supergravity and supersymmetry breaking, see for instance \cite{Ferrara:1995gu, Fre:1996js, Fre':2003gd, Fre:2002pd}. \\
835: The reason that de Sitter space is relevant, is that astronomical observations indicate that our own universe is in a de Sitter phase, see \cite{Trodden:2004st, Peebles:2002gy} and references therein for more information. Therefore it is exciting that we can reproduce a de Sitter vacuum from our supergravity setup. Note that one really should compute the coefficient $\hA_{1,m_1}$ in string theory because in our supergravity approach it is still a free parameter. This means that only if the result is such that $A_\textrm{min} < \hA_{1,m_1}<A_\textrm{max}$ we can conclude that we can produce a de Sitter vacuum from string theory.\\
836: Naturally we are not the first to have searched for a de Sitter vacuum in string theory (see \cite{Fre:2002pd, Fre':2003gd}). There is, for instance, the work of `KKLT' \cite{Kachru:2003aw}, to which we will compare our results.\\
837: The work of KKLT consists of three parts. Firstly, all moduli, apart from the \Kh moduli, are fixed by the inclusion of fluxes. Secondly, the \Kh moduli (the volume modulus to be precise) are stabilized by nonperturbative instanton effects, but at an AdS vacuum. Thirdly, a positive energy contribution in the form of anti D$3$-branes\footnote{They work in type IIB string theory.} is added to lift the minimum of the potential to a positive value, i.e., a dS vacuum.\\
838: In our approach the inclusion of a single R-R $3$-form spacefilling flux provides a potential. Classically this is of the runaway type in the dilaton and both R-R scalars $\gc$ and $\gvf$ are flat directions. Only when we include the membrane corrections (plus the one-loop corrections) does this change, leading to a (meta-)stable vacuum. Depending on $\hA_{1,m_1}$ this is an AdS, Minkowski or dS vacuum. \\
839: Note that, in contrast to KKLT, we do not have to include additional positive energy contributions by hand. This contribution is already provided for by the background flux at the classical level. Furthermore, by making a suitable choice for the numerical parameters corresponding to the one-loop determinant around a one-instanton background, the moduli can be stabilized in a meta-stable de Sitter vacuum at small string coupling constant $g_s\ll 1$. This is consistent with the fact that we only took into account the terms dominating for small $g_s$ (\ref{eq:finst}). \\
840:
841: We have examined a simple model in which we have one hypermultiplet and no vector multiplets. As this hypermultiplet is present in any $N=2$ supergravity theory in four dimensions obtained by compactifying type IIA supergravity on a Calabi-Yau manifold, the results are expected to have validity for more general Calabi-Yau compactifications as well (that is $h^{(1,2)}\neq 0$).
842: A next step would then be to include vector multiplets and hypermultiplets and consider more general fluxes, e.g. $2$- and $4$-form fluxes related to the even homology cycles of the Calabi-Yau manifold, thus stabilizing the K\"ahler moduli\footnote{Remember that these moduli reside in the vector multiplets.}.
843: One could also try to include the nonperturbative effects coming from the $5$-brane, as calculated in the previous chapter, into the potential.
844:
845:
846:
847: \clearpage{\pagestyle{empty}\cleardoublepage}
848:
849:
850:
851:
852:
853:
854:
855:
856:
857:
858:
859:
860:
861:
862:
863:
864:
865:
866:
867:
868:
869:
870:
871:
872:
873:
874:
875:
876:
877:
878:
879:
880:
881:
882:
883:
884:
885:
886:
887: