1: \chapter{Calabi-Yau compactifications}
2: \label{ch2}
3: In the first part of this chapter we will schematically review the compactification of $11$-dimensional supergravity to $4$ dimensions, using a CY $3$-fold. Specific attention shall be given to the M$5$-brane and the M$2$-brane.\\
4: We start by compactifying the bosonic fields of $11$-dimensional supergravity on a circle, which gives the fields of type IIA supergravity.
5: We then outline the compactification of type IIA on a general Calabi-Yau manifold. This gives rise to a certain set of fields in four dimensions: the bosonic field content of $N=2$ supergravity coupled to vector multiplets and hypermultiplets.
6: We focus on the universal sector which arises as a tensor multiplet that does not depend on the details of the Calabi-Yau manifold. Next we will investigate this tensor multiplet by considering its alternative formulations: the universal hypermultiplet and the double-tensor multiplet. We will focus on this sector because we will restrict ourselves for simplicity to compactifications on rigid Calabi-Yau manifolds.
7:
8: A point worth emphasizing is that although we speak about `branes' frequently, no worldvolume actions or boundary states will be used. They will always be dealt with from the point of supergravity. The branes will always appear as solitonic objects and consequently will be treated in a `macroscopic' manner.
9:
10: \section{$11$ dimensions}
11: \label{11d}
12: In this section we follow the conventions of \cite{Becker:1999pb}.\\
13: The bosonic part of the $11$-dimensional \sugra action \cite{Cremmer:1978km}, is given by
14: \equ{
15: S_{11}=\2 \int d^{11}x \sqrt{-\hg} \hR -\4 \int \left(\hF_4 \wedge ^\star \hF_4 -\3 \hA_3 \wedge \hF_4\wedge\hF_4\right) \label{eq:11dsugra1} \pt
16: }
17: We have set the $11$-dimensional gravitational coupling constant $\gk_{11}$ to $1$.
18: The \sugra action contains a $3$-form potential $\hA_3$ with field strength $\hF_4=d \hA_3$ which obeys the Bianchi identity
19: \equ{
20: d \hF_4=0 \label{eq:bianchi1}
21: }
22: and the field equation
23: \equ{
24: d \,^\star \hF_4 + \2 \hF_4^2=d\left( ^\star \hF_4 +\2 \hA_3 \wedge \hF_4\right)=0 \label{eq:eom1} \pt
25: }
26: This equation gives rise to the conservation of an `electric' type charge\footnote{Actually, these charges are quantized, as we will see later explicitly, in a slightly different setting in section \ref{stcal1}.}
27: \equ{
28: q_e = \int_{\p{}\cM_8} \left( ^\star \hF_4 +\2 \hA_3 \wedge \hF_4\right) \label{eq:electric1} \km
29: }
30: the integral is over the boundary at infinity of some spacelike $8$-dimensional subspace of $11$-dimensional spacetime, \cite{Page:1984qv, Duff:1990xz, Gueven:1992hh}. On the other hand, \R{bianchi1} gives a `magnetic' conserved charge
31: \equ{
32: q_m=\int_{\p{}\cN_5} \hF_4 \label{eq:magnetic1} \km
33: }
34: where we now integrate over the boundary (at infinity) of a spacelike $5$-dimen\-sional subspace. \\
35: Just as the Maxwell $1$-form couples naturally to the worldvolume (a worldline in fact) of a charged particle, $\hA_3$ couples naturally to the worldvolume of a $2$-dimensional charged extended object. The charge is given by \R{electric1} and represents the charge of a \emph{membrane} with a $3$-dimensional worldvolume. The integration surface $\p{} \cM_8$ is a surface surrounding the membrane.\\
36: The second way to come up with an extended object is to think about the dualized field strength $^\star \hF_4$, a $7$-form. Suppose one would find an underlying potential for $^\star \hF_4$, then this potential would couple to a $6$-dimensional worldvolume. This `solitonic' or `magnetic' object is called the \emph{fivebrane}\footnote{M$5$-brane really: the $5$-brane from M-theory, of which the low energy effective theory is presumably given by $11$-dimensional supergravity.}, which means that $\p{}\cN_5$ in \R{magnetic1} actually surrounds this $5$-brane. \\
37: One procedure to go from $11$ to $10$ dimensions is to compactify on a circle, i.e., $x_{11}\sim x_{11}+2\gp$. By compactifying $11$-dimensional supergravity on a circle one obtains type IIA supergravity in $10$ dimensions\footnote{For more information on this procedure see \cite{Polchinski:1998rr}.}. The Kaluza-Klein ansatz for the metric is
38: \equ{
39: ds_{11}^2=e^{-2\gf/3} g_{AB} dx^A dx^B + e^{4\gf/3} \left(dx^{11} -A_A dx^A\right)^2 \label{eq:stf1} \km
40: }
41: where $A_A$ is a vector field, $A=0,\ldots, 9$, $\gf$ is the $10$-dimensional dilaton. The $11$-dimensional $3$-form $\hA_3$ gives the $10$-dimensional objects
42: \equ{
43: A_{ABC}=\hA_{ABC} \qquad B_{AB}=\hA_{AB 11} \label{eq:hgh1} \km
44: }
45: with $10$-dimensional field strengths $H_3 = dB_2$, $F_2=dA$ and $F_4 = dA_3$. \\
46: Using the same reasoning as before, the $5$-brane charge in $10$ dimensions is associated to
47: \equ{
48: \int_{\p{} \cN_4} H_3\label{eq:magnetic2}
49: }
50: and the membrane charge to
51: \equ{
52: \int_{\p{}\cM_7} \left(e^{\gf/2} ~^\star F_4 + A_3 \wedge H_3\right) \label{eq:electric2} \pt
53: }
54: Let us take a closer look at these objects. Consider, for instance, the $5$-brane, the magnetic dual of the fundamental string (in $10$ dimensions). This object was originally discovered as a consistent background for supergravity, see \cite{Duff:1990wv, Callan:1991ky, Stelle:1998xg}. In the case of type IIA \sugra it is characterized by the following equations:
55: \equ{
56: e^{2\gf}=e^{2\gf_\infty} + \frac{Q}{r^2} \qquad H_{\gm\gn\gr}=\gve_{\gm\gn\gr}^{\,\,\,\,\,\,\,\,\,\gs} \p{\gs} \gf \label{eq:nsbrane2} \km
57: }
58: where $Q$ is the charge of the $5$-brane and $e^{2\gf_\infty}$ is an integration constant. The (conformally) flat $4$-dimensional space transversal to the $5$-brane is parametrized by $x^\gm$. The transversal distance $r=|x|$ is measured from a point $x$ in transversal space to the brane which is located for simplicity at the origin. The longitudinal space can taken to be flat. Actually, since the geometry of this solution is simply a direct product, we can insert more complicated geometries in the direction of the $5$-brane, as long as they satisfy the Einstein equations. The transversal geometry will remain the same. For example, one can take the geometry of a CY $3$-fold.
59: However, note that we must perform a Wick rotation such that the brane becomes a Euclidean brane before wrapping it around the Calabi-Yau. The result is then a $5$-brane which is wrapped around the CY, thus appearing in the transversal space as an instanton located at the origin.
60: We can construct a similar setup for the membrane, but in this case the membrane cannot wrap around the whole of the CY, but it has to wrap a $3$-cycle.
61:
62:
63:
64: \section{$4$ dimensions}
65: \label{fourd}
66: One can compactify type IIA supergravity to four dimensions on a $4$-dimensional Minkowski space times an internal Calabi-Yau $3$-fold, i.e., $\cM^{1,3} \times Y_3$. The result is an $N=2$ supergravity theory coupled to vector multiplets and hypermultiplets. It is instructive to consider the amount of supersymmetry preserved by such a compactification first.\\
67: A compactification specifies a background for the $10$-dimensional \sugra theory. This background preserves a certain amount of supersymmetry. The precise amount preserved by this groundstate is determined by the number of independent supersymmetry parameters for which the fermionic \susy variations vanish.
68: If one specifies the background by setting the background values for the fermionic fields in the theory to zero, the only \susy variation which is not trivially zero is the variation of the gravitino:
69: \equ{
70: \gd \gps_A^i=\nabla_A \ge^i \nn \km
71: }
72: with $A$ a 10-dimensional index.
73: In type IIA \sugra in $10$ dimensions, there are two independent \susy parameters $\ge^i\,,\, i=1,2$ and two gravitino's. We see that the variation only vanishes if there are covariantly constant spinors in our Kaluza-Klein background. \\
74: For the compactification ansatz of $4$-dimensional Minkowski space times a Calabi $3$-fold, the spinors $\ge^i$ decompose as
75: \equ{
76: \ge^i \propto \bget^i\otimes \gz(y) + \get^i \otimes \gz^\dagger(y) \nn \km
77: }
78: where $\gz(y)$ is a covariantly constant spinor on the CY $3$-fold, $\gz^\dagger(y)$ its complex conjugate and the $\get^i$ are complex $4$-dimensional Weyl spinors.\\
79: A CY $n$-fold admits one covariantly constant spinor, see the discussion in section \ref{CY}. This means that in $4$ spacetime dimensions we are left with the two \susy parameters $\get^1$ and $\get^2$, together providing $8$ supercharges, i.e., $N=2$ supergravity. For more information about the details and consistency of this Kaluza-Klein program see \cite{Lust:1997kx, Duff:1986hr, Pope:1987ad, Duff:1989cr}.\\
80: In chapter \ref{ch3} we will elaborate on the notion of Killing spinors and perform a detailed analysis on the amount of supersymmetry preserved when we choose a non-trivial bosonic background in $4$ dimensions, namely the instanton background. \\
81:
82: Next we illustrate how the scalar fields of the hypermultiplets and the gravitational and vector multiplets of the $N=2$ supergravity theory in $4$ dimensions arise.\\
83: We follow the Kaluza-Klein program and expand the $10$-dimensional fields in harmonics on the internal CY $3$-fold $Y_3$. If we collectively denote the fields by $\gF^{(I)}(x,y)$, where $x^\gm$ are the Minkowski spacetime coordinates and $y^m$ (or $(z^i, \bz^\bi)$) are the internal (complex) coordinates, we can write
84: \equ{
85: \gF^{(I)}(x,y)=\sum_\ga \gF^{(I)}_\ga(x) Y^\ga_{(I)}(y) \nn \km
86: }
87: with $Y^\ga_{(I)}(y) $ the appropriate harmonics.\\
88: The light fields appearing in $4$ dimensions are those coefficient functions $\gF^{(I)}_0(x)$ whose harmonics $Y^0_{(I)}(y)$ are zero modes of the internal Laplacian $\gD_{(I)}$ which acts on them in the field equations. \\
89: For instance, consider a $2$-form field $\hB_2$ in $10$ dimensions with action
90: \equ{
91: \int d\hB \wedge ^\star d \hB \nn \pt
92: }
93: Decomposing the $10$-dimensional space as $\cM^{1,3} \times Y_3$, the $10$-dimensional Laplacian decomposes as $\gD_{10} =\gD_4 + \gD_6$. The equation of motion $d ^\star d \hB_2=0$ together with the gauge condition $d ^\star \hB_2=0$ can be written as $\gD_{10} \hB_2=0$, see \R{Laplacian1}. Upon decomposing this becomes
94: \equ{
95: \left( \gD_4 + \gD_6\right)\hB_2 =0 \nn \pt
96: }
97: A massless $B_2$ field in $4$ dimensions should obey $\gD_6 \hB_2=0$.
98: This means that we must expand $\hB_2$ in harmonics of $\gD_6$.
99: We know, see \R{A}, \R{B} and \R{C}, that such harmonics correspond to the various cohomology groups on $Y_3$. Therefore the following reductions of $\hB_2$ can be associated to the cohomology groups, in complex coordinates
100: \bea
101: && B_{\gm \gn} \leftrightarrow H^{(0,0)}(Y_3) \nn \\
102: && B_{\gm i} \,\leftrightarrow H^{(1,0)}(Y_3) = \emptyset \label{eq:Bred1} \\
103: && B_{i j} \,\leftrightarrow H^{(2,0)}(Y_3) =\emptyset \nn \\
104: && B_{i\bj} \,\leftrightarrow H^{(1,1)}(Y_3) \nn \pt
105: \eea
106: Explicitly this means that we can expand $\hB_2$ as
107: \equ{
108: \hB_2=B_2 + b^i \go_i \label{eq:Bexp1} \km
109: }
110: where the $\go_i$ are $(1,1)$-forms, $i=1, \ldots, h^{(1,1)}$. $B_2$ is a $2$-form in $4$ dimensions and
111: the $b^i$ are real scalar fields in $4$ dimensions: $b^i(x)$.\\
112: The other fields are treated in a similar manner. Reducing to $10$ dimensions gave us the following NS-NS fields:
113: \equ{
114: B_2 \, ,\quad g_{AB}\, , \quad A_A \quad \textrm{and} \quad \gf \nn
115: }
116: and the R-R field $A_3$, $B_2$ has just been discussed. The metric field $g_{AB}(x,y)$ can be decomposed as $g_{\gm\gn}(x)$ (the $4$-dimensional graviton) $g_{\gm m}$ and $g_{mn}$, or $g_{m i}$, $g_{ij}$, $g_{i\bj}$ and their complex conjugates. In section \ref{CY} we have discussed the zero modes of $g_{ij}$ and $g_{i\bj}$ and we have seen that they correspond to deformations of the \Kh class and the complex structure, see \R{mixed} and \R{pure}, which means that they can be expanded as
117: \equ{
118: \I \gd g_{i\bj} =\sum_{l=1}^{h^{(1,1)}} v^l(x) \go^l \label{eq:mixed2}
119: }
120: and
121: \equ{
122: \gd g_{ij}=\sum_{a=1}^{h^{(1,2)}} z^a(x) \bb_{a ij} \label{eq:pure2} \pt
123: }
124: The $\bb_{a ij}$ are related to the $(1,2)$-forms $\gg_a \,\,\, a=1, \ldots , h^{(1,2)}$ (not in a real basis this time) by
125: \equ{
126: \bb_{aij}=\frac{\I}{|\!|\gO|\!|^2} \gO_i^{\, \bl \bk} \bgg_{a; \bl \bk j} \label{eq:12form} \km
127: }
128: where $|\!|\gO|\!|^2 \equiv \frac{1}{3!} \gO_{ijk} \bgO^{ijk}$. Furthermore $g_{\gm i}=0$ because $h^{(1,0)}=0$, which we used in \R{Bred1} as well. Similarly, the $1$-form $A$ only gives a $1$-form in $4$ dimensions: the graviphoton $A^0$. The $10$ dimensional dilaton gives a $4$-dimensional dilaton, $\gf(x)$. Lastly, the R-R $3$-form $A_3$ is expanded as
129: \equ{
130: A_3=C_3 + A^i \wedge \go_i + \gx^A \ga_A + \tgx_B \gb^B \label{eq:3form1e} \km
131: }
132: where $C_3(x)$ is a $3$-form on $\cM^{1,3}$, the $h^{(1,1)}$ $A^i(x)$'s are $1$-forms. As before, see \R{basissym1}, the $(\ga_A, \gb^A)$ form a real basis of $H^3(Y_3)$, consequently $A_3$ gives us $2(h^{(1,2)}+1)$ real scalars. \\
133: These fields organize themselves into the following multiplets: the gravitational multiplet
134: \equ{
135: \left(A^0, g_{\gm\gn}\right) \label{eq:gravm1a}
136: }
137: and the vector multiplets
138: \equ{
139: \left(A^i, v^i, b^i\right) \nn \km
140: }
141: containing $h^{(1,1)}$ vectors and twice as many real \label{spk2} scalars\footnote{These scalars are often treated in one go by \emph{complexifying the \Kh cone}: write $B_2 + \I J=(b^i + \I v^i) \go_i \equiv t^i \go_i$, where we now have $h^{(1,1)}$ complex scalars $t^i$.}. The hypermultiplets
142: \equ{
143: \left(z^a, \gx^a, \tgx_a\right) \label{eq:hyperm1} \km
144: }
145: contain the $h^{(1,2)}$ complex scalar fields $z^a$ from \R{pure2} and in total $4h^{(1,2)}$ real scalar fields. \\
146: Finally there is the \emph{tensor} multiplet, TM for short:
147: \equ{
148: \left(\gx^0, \tgx_0, \gf, B_2\right) \label{eq:tenm1a} \km
149: }
150: a multiplet which is always present and does not depend on the details of the CY $3$-fold since $B_2$ and $\gf$ are related to the $0$-form of the $CY$ and $(\gx^0, \tgx_0)$ are the expansion coefficients of $A_3$ w.r.t. the unique $\gO$. This means that if we were to consider a \emph{rigid} ($h^{(1,2)}=0$) CY this multiplet would still be present.\\
151: Often the tensor multiplet is dualized to an additional hypermultiplet since a $2$-form field in $4$ dimensions is dual to a (pseudo) scalar field, generally called the axion. The resulting multiplet is called the \emph{universal hypermultiplet} (UHM) because it must be present in any CY-compactification. \\
152: Due to supersymmetry \cite{deWit:1984px} (see also \cite{Aspinwall:2000fd} for a geometrical argument) the total target manifold parametrized by the various scalars factorizes as a product of vector and hypermultiplet manifolds:
153: \equ{
154: \cM_{\textrm{scalar}}=\cM_V\otimes \cM_H \label{eq:scm1} \pt
155: }
156: $\cM_V$ is a special \Kh manifold and $\cM_H$ is a quaternionic manifold.\\%, \cite{Ferrara:1989ik}. \\
157: This is a very important result, since it means that string corrections, whether perturbative or nonperturbative, only affect the hypermultiplets, since the dilaton belongs to a hypermultiplet and $g_s\equiv e^{-\gf/2}$. The vector multiplet geometry remains unaffected\footnote{Just as the hypermultiplet metric receives no $\ga'$ corrections, in type IIA that is. For a general discussion on the corrections to the moduli spaces of both type IIA and type IIB, see \cite{Aspinwall:2000fd}.}. \\
158:
159:
160: \label{hereI}
161: The $3$-form $C_3$ is dual in $4$ dimensions to a constant $e_0$. It turns out that taking this space-filling $C_3$ along in the compactification (or conversely $e_0$) has the effect of gauging the axion, see \cite{Louis:2002ny}. The graviphoton $A^0$ acts as the gauge field and the gauging gives rise to a potential, this will be discussed at a later stage.
162:
163: In the above we have schematically derived the bosonic field content of ($N=2$) supergravity in $4$ dimensions. Similarly one can reduce the whole $10$-dimensional action to obtain the $4$-dimensional one and relate in this way all the terms appearing in the $4$-dimensional action to the original fields and the geometric data of the CY $3$-fold. This is done in the seminal paper \cite{Bodner:1990zm}, see also the useful appendices in \cite{Louis:2002ny}.
164:
165: Conversely, knowing that the reduction gives one \sugra multiplet, $4(h^{(1,2)} +1)$ hypermultiplets and $h^{(1,1)}$ vector multiplets, one could construct such an action from first principles. One could also choose to leave out the vector multiplets, since the moduli space factorizes anyway. Such an action has been constructed in \cite{Bagger:1983tt}, in which the most general action for hypermultiplets coupled to $N=2$ supergravity is constructed. This action is a non-linear sigma model\footnote{See \cite{deWit:2002vz} and the references therein for non-linear sigma models in supergravity.} where the Lagrangian contains all kinds of complicated non-linear terms, specifically the kinetic terms which provide a metric for the target space. Supersymmetry constrains these `complicated terms' such that the hypermultiplet sector of the theory parametrizes a quaternionic manifold. This means that if we assume supersymmetry is unbroken in the $4$-dimensional supergravity theory, quantum corrections cannot give new terms in the effective action.
166: The various quantities, such as the metrics for the kinetic terms, may change, but only in such a way that the supersymmetry relations are still obeyed. Geometrically speaking, the scalars keep parametrizing a quaternionic manifold, although perhaps a different one. \\
167: This is a very powerful and useful concept which will be put to use in chapter \ref{ch4}. In that chapter we will construct perturbations of the classical hypermultiplet space (corresponding to the classical UHM action). These perturbations respect the quaternionic geometry of the target-space and describe, as it will turn out, effects that must be attributed to membrane instantons. \\
168: The conclusion of the above discussion is that the possible effects the $5$-brane and the membrane can have, must show up in the (universal) hypermultiplet(s) sector of the effective action, which is why we will be neglecting the vector multiplets from now on.
169:
170: %So far we have only dealt with bosonic fields, but it is instructive to consider the fermionic side of the story for a brief moment.
171: %Compactifying as we did in the above, specifies a background for the $10$-dimensional \sugra theory. This background preserves a certain amount of supersymmetry. The precise amount preserved by this groundstate is determined by the number of independent supersymmetry parameters for which the fermionic \susy variations vanish.
172: %If one specifies the background by setting the background values for the fermionic fields in the theory to zero
173: %, the only \susy variation which is not trivially zero is the variation of the gravitino:
174: %\equ{
175: %\gd \gps_A^i=\nabla_A \ge^i \nn \km
176: %}
177: %with $A$ a 10-dimensional index.
178: %In type IIA \sugra in $10$ dimensions, there are two independent \susy parameters $\ge^i\,,\, i=1,2$ and two gravitino's. We see that the variation only vanishes if there are covariantly constant spinors in our Kaluza-Klein background. \\
179: %For the compactification ansatz of $4$-dimensional Minkowski space times a Calabi $3$-fold, the spinors $\ge^i$ decompose as
180: %\equ{
181: %\ge^i \varpropto\bget^i\otimes \gz(y) + \get^i \otimes \gz^\dagger(y) \nn \km
182: %}
183: %where $\gz(y)$ is a covariantly constant spinor on the CY $3$-fold, $\gz^\dagger(y)$ its complex conjugate and the $\get^i$ are complex $4$-dimensional Weyl spinors.\\
184: %We know that (see the discussion in section \ref{CY}) a CY $n$-fold admits one covariantly constant spinor. This means that in $4$ spacetime dimensions we are left with the two \susy parameters $\get^1$ and $\get^2$, together providing $8$ supercharges, i.e. $N=2$ supergravity. For more information about the details and consistency of this Kaluza-Klein program see \cite{Lust:1997kx, Duff:1986hr, Pope:1987ad, Duff:1989cr}.\\
185: %In chapter \ref{ch3} we will elaborate on the notion of Killing spinors and perform a detailed analysis on the amount of supersymmetry preserved when we choose a non-trivial bosonic background in $4$ dimensions, namely the instanton background. \\
186:
187: The aim of the next sections is to investigate the $4$-dimensional theory for which we will compute instanton corrections in chapters \ref{ch3} and \ref{ch4}.
188:
189: \section{The universal hypermultiplet}
190: \label{instsol}
191: The easiest situation to consider is compactifying type IIA supergravity on a rigid CY $3$-fold, because we then only have to deal with the universal tensor multiplet. The vector multiplets can be consistently truncated to zero for the purposes of this chapter. The resulting four-dimensional low energy bosonic effective Lagrangian is given by
192: \bea
193: && e^{-1} \cL_T =-R -\2 \pu{\gm}\gf \p{\gm}\gf + \2 e^{2\gf} H^\gm H_\gm -\4 F^{\gm \gn} F_{\gm\gn} \!\!\quad \textrm{(the NS-NS sector)} \nn \\
194: &&\,\,\, -\2 e^{-\gf}\left(\pu{\gm}\gc \p{\gm}\gc + \pu{\gm}\gvf \p{\gm}\gvf\right) -\2 H^\gm \left(\gc \p{\gm} \gvf -\gvf \p{\gm}\gc\right) \quad \textrm{(the R-R sector)} \nn
195: \eea
196: The R-R scalars $\gvf$ and $\gc$ were formerly known as $\gx^0$ and $\tgx_0$, compare with \R{tenm1a}. Notice that both $\gc$ and $\gvf$ have constant shift symmetries. $H^\gm=\frac{1}{2}\gve^{\gm\gn\gr\gs}\p{\gn}B_{\gr\gs}$ is the dual NS $2$-form fields strength (i.e. $H_3=d B_2$). The tensor $B_2$ also has a shift symmetry and is dual to the axion.
197: $F^{\gm\gn}$ is the field strength associated to the graviphoton $A_\gm$, previously denoted by $A^0$ (see \R{gravm1a}). As before the dilaton is denoted by $\gf$.\\
198:
199: We can transform this Lagrangian into the Lagrangian of the UHM by dualizing\footnote{This will be explained later after giving some information about the isometries of the UHM.} $B_2$ to the axion $\gs$ giving
200: \bea
201: e^{-1} \cL_{UH} &=&-R -\2 \pu{\gm}\gf \p{\gm}\gf -\4 F^{\gm \gn} F_{\gm\gn} \quad \label{eq:UHM1} \\
202: &&\,\,\, -\2 e^{-\gf}\left(\pu{\gm}\gc \p{\gm}\gc + \pu{\gm}\gvf \p{\gm}\gvf\right) -\2 e^{-2\gf}\left(\p{\gm} \gs + \gc \p{\gm}\gvf\right)^2 \nn \km
203: \eea
204: which is the classical UHM at string tree-level. It is a non-linear sigma model \cite{Cecotti:1988qn, Ferrara:1989ik} with a quaternionic target space corresponding to the coset space
205: \equ{
206: \frac{SU(1,2)}{U(2)} \label{eq:uhmgeo1} \km
207: }
208: where $SU(1,2)$ is the isometry group\footnote{See for instance \cite{Ketov:2001gq, Besse} for some alternative parametrizations of this metric.} of the Lagrangian.
209: This space is actually \Kh as well, as can be seen by going to the frequently used parametrization
210: \bea
211: && e^\gf=\2\left( S+\bS -2C\bC\right) \quad \,\qquad \gc = C+\bC \nn \\
212: && \gs =\frac{\I}{2}\left(S-\bS + C^2 -\bC^2\right) \qquad \gvf=-\I\left(C-\bC\right) \label{eq:realb1} \pt
213: \eea
214: In these coordinates the \Kh potential is given by
215: \equ{
216: \wK=-\ln\left(S+\bS -2 C\bC\right) \nn
217: }
218: and the line element on the target space by
219: \equ{
220: ds^2=e^{2\wK}\left(dS d\bS -2C dS d\bC -2\bC d\bS dC + 2(S+\bS)dC d\bC\right) \nn \km
221: }
222: the pullback of which, to $\cM^{1,3}$, precisely gives the Lagrangian. The $8$ isometries which leave the Lagrangian invariant were analyzed in \cite{deWit:1990na, deWit:1992wf}. The shift isometries of the R-R scalars and the axion mentioned above form a subgroup of the isometry group. This group is known as the Heisenberg group and acts on the fields as
223: \bea
224: && S\to S+\I \ga + 2 \bge C + |\ge|^2 \nn \\[1mm]
225: && C\to C+\ge \label{eq:heis1} \km
226: \eea
227: where the finite $\ga$ and $\ge$ are respectively real and complex. In the real basis \R{realb1} the Heisenberg group acts as
228: \equ{
229: \gf\to\gf \, \quad \gc\to \gc+\gg \,\quad \gvf\to \gvf + \gb \,\quad \gs \to \gs-\ga-\gg \gvf \label{eq:heis2} \km
230: }
231: $\gb=\I (\ge-\bge)$ and $\gg=(\ge+\bge)$.
232: The generators of the Heisenberg group obey the commutation relations
233: \equ{
234: [\gd_\ga, \gd_\gb]=[\gd_\ga, \gd_\gg]=0 \qquad [\gd_\gb, \gd_\gg]=\gd_\ga \nn \pt
235: }
236:
237: Furthermore there is a $U(1)$ symmetry which acts as a rotation on $\gvf$ and $\gc$, together with a compensating transformation on $\gs$. Its finite transformation can be determined from the results in \cite{Becker:1995kb, Davidse:2003ww} and reads
238: \bea
239: && \gvf \to \cos(\gd) \,\gvf + \sin(\gd) \,\gc \qquad \gc \to \cos (\gd) \,\gc -\sin (\gd)\, \gvf \km \nn \\ [1.5mm]
240: && \gs \to \gs -\4 \sin (2 \gd) \left(\gc^2 -\gvf^2\right) + \sin^2 (\gd)\, \gc \gf \label{eq:u1tr} \pt
241: \eea
242: Four of the eight isometries have now been specified.\\
243: The isometries of the Heisenberg group are expected to survive at the perturbative level. The reasons for expecting this are as follows. The Heisenberg group does not act on the dilaton, i.e., on $g_s$. This means that we can examine the Heisenberg group order by order in perturbation theory. Furthermore, the shifts in $\gc$ and $\gvf$ are preserved because they originated from the gauge transformation of the $3$-form $A_3$ in ten dimensions, see \R{3form1e}. The field strength of this $3$-form will involve its derivative, $dA_3$. The gauge symmetry is therefore given by $\gd A_3=d\gL_2$, where $\gL_2$ is a $2$-form. If one works out $dA_3$ using \R{3form1e} one finds (among others) the terms $d\gx^0 \ga_0 +d\tgx_0 \gb^0$ for which the gauge symmetry becomes
244: \equ{
245: \gx^0\to \gx^0 + \gg \qquad \tgx_0 \to \tgx_0+ \gb \nn \pt
246: }
247: We gave the constant shifts the same name as in \R{heis2}, to which these transformations correspond. We can think about the shifts in the axion in a similar way. \\
248: Finally there are the $4$ remaining isometries we do not give here. They involve non-trivial transformations on the dilaton and hence will change the string coupling constant. We will not consider these isometries nonperturbatively, in fact it is not even known what happens to them perturbatively.
249: Although the isometries of the Heisenberg group (\ref{eq:heis1}, \ref{eq:heis2}) are unaffected by quantum corrections, which appear at one-loop only in the string frame see \cite{Antoniadis:2003sw, Strominger:1997eb, Antoniadis:1997eg, Anguelova:2004sj}, nonperturbative effects will quantize these isometries. This will be discussed in chapter \ref{ch3}.
250:
251:
252: %As stated above, the one-loop calculation will be done in a supergravity setting, in which we can perform these semi-classical (for weak string coupling) instanton calculations in a fairly conventional field-theoretic manner. Although such an approach will only give the leading \sugra corrections, one might hope that knowledge of string theory and the isometries of the UHM target manifold provides a more `complete' answer\footnote{Such a program has worked successfully in the context of supersymmetric field theories in three dimensions with eight supercharges, where the hypermultiplet moduli space is hyperk\"ahler, \cite{Seiberg:1996nz, Dorey:1997ij, Dorey:1998kq}, see also \cite{Ooguri:1996me, Ketov:2001gq, Ketov:2001ky} for related issues.}.
253: %Instanton calculations are necessarily done in the Euclidean formulation where the instanton solutions appear as minima of the action. One convenient way to find such minima is by way of Bogomol'nyi equations. Solutions to such equations satisfy the equations of motion and hence give a minimum of the action. But whereas equations of motion are generically of second order, Bogomol'nyi equations are of first order and therefore of much more practical use\footnote{There is of course an immense literature on instanton calculations, see
254: %for instance \cite{Dorey:2002ik, Belitsky:2000ws}, and the references therein, for instantons in the context of supersymmetric theories.}.
255: %What one therefore would like to do is to look for a minimum of the UHM, utilizing such Bogomol'nyi equations for instance, and then perform a semiclassical expansion around this minimum. This would give the desired nonperturbative corrections to the UHM metric, and to the action of course. \\
256:
257: \section{The double-tensor multiplet}
258: \label{secDTM1}
259: Instead of working with the UHM we will find it convenient to work with the so-called \emph{double-tensor multiplet}, or DTM for short.
260: As the name suggests, the DTM contains $2$ tensors and arises by dualizing a R-R scalar \mbox{($\gvf$ or $\gc$)} from the tensor multiplet to an additional tensor. We shall start by discussing the DTM in Lorentzian signature and the actual dualization process. In the next section we shall give its Euclidean formulation and the reason why we choose the Euclidean DTM over the Euclidean UHM.\\
261: The bosonic part of the DTM Lagrangian is given by
262: \equ{
263: e^{-1}\cL_{DT}=-R -\2 \pu{\gm}\gf\p{\gm}\gf -\4 F^{\gm\gn}F_{\gm\gn} -\2 e^{-\gf} \pu{\gm}\gc \p{\gm}\gc + \2 M^{IJ} H_I^\gm H_{\gm J} \label{eq:DTM1} \km
264: }
265: where we have dualized $\gvf$ into another $H$. Comparing with the UHM gives $\gf^I=(\gvf, \gs) \leftrightarrow H^I=(H^1, H^2)$. In effect we can go from the UHM to the DTM, the reason is that the UHM possesses two commuting isometries over which we can dualize, for instance the shift symmetries in $\gvf$ and $\gs$ corresponding to $\ga$ and $\gb$. The DTM also possesses such symmetries since $B_1$ and $B_2$ only appear through their derivatives. \\
266: The kinetic term (metric) for the $3$-form field strengths $H_I^\gm=\2 \gve^{\gm\gn\gr\gs}\p{\gn}B_{\gr\gs I}$ is given by
267: \equ{
268: M^{IJ}=e^\gf \left(%
269: \begin{array}{cc}
270: 1 & -\gc \\
271: -\gc & e^\gf + \gc^2 \\
272: \end{array}%
273: \right) \label{eq:tensormetric1} \pt
274: }
275: The remaining two scalars $\gf$ and $\gc$ parametrize the coset $Sl(2,\mR)/O(2)$, see appendix \ref{appdtm1} for the various target manifolds. However, the two tensors break this $Sl(2,\mR)$ symmetry to a $2$-dimensional subgroup generated by rescalings of the tensors and by the remaining generator of the Heisenberg algebra \R{heis2} acting on $\gc$ and $B_1$ as
276: \equ{
277: \gc \to \gc + \gg \qquad B_1 \to B_1 + \gg B_2 \label{eq:tensorisom1} \km
278: }
279: $\gf$ and $B_2$ are invariant. Summarizing: in going from the UHM to the DTM two of the shift symmetries of the Heisenberg group, namely $\ga$ and $\gb$ (see \R{heis2}), have been used to dualize to two tensors, which now in turn have shift symmetries. The remaining shift symmetry $\gg$ now acts on $\gc$ and $B_1$ as in \R{tensorisom1}.\\
280: We will sometimes use the `Heisenberg invariant'
281: \equ{
282: \hH_1\equiv H_1 -\gc H_2 \label{eq:heisinv1} \km
283: }
284: which, as the name suggests, is invariant under \R{tensorisom1}.\\
285:
286: Let us now finally be more specific about the process of dualization, for more details see \cite{Theis:2003jj}. In general one can dualize the scalars of a bosonic sigma model into tensors if they appear only through their derivatives. This means that the target manifold has a set of Abelian Killing vectors\footnote{Actually, the Killing vectors have to leave the complex structures invariant and the isometries have to commute with supersymmetry. For a quaternionic manifold this is automatically the case. For more details see \cite{deWit:2001bk}.},
287: \equ{
288: \gd_\gTh \gf^{\hA} =\gTh^I k_I^\hA(\gf) \qquad [k_I, k_J]=0 \nn \pt
289: }
290: We can choose coordinates $\gf^\hA=(\gf^A, \gf^I)$, where $\gf^A=(\gf, \gc)$, such that these transformations act as constant shifts on the $\gf^I$ and leave the $\gf^A$ invariant. When writing $\gf^\hA$ we mean to denote all the scalars in the UHM, i.e. $\gf^\hA=(\gf,\gc,\gvf,\gs)$. $\gf^I$ are the scalars in the UHM which are dualized into the two tensors of the DTM and $\gf^A$ are the scalars in the DTM. As the notation suggests, tensors can be dualized as well if they appear solely through their field strengths. Furthermore they guarantee commuting isometries in the corresponding scalar sigma model, so if we go from the DTM to the UHM we have the two commuting shift isometries discussed earlier. \\
291: This procedure also works at the perturbative level, as mentioned before, since the Heisenberg group is preserved, \cite{Antoniadis:2003sw}. At the nonperturbative level things are slightly more subtle, we will discuss this. \\
292: To dualize from the UHM to the DTM one replaces the derivatives on $\gf^I$ by covariant derivatives. Add Lagrange multipliers $B_{\gm\gn I}$ with $H^\gm_I=\frac{1}{2}\gve^{\gm\gn\gr\gs}\p{\gn}B_{\gr\gs I}$. Integrating out these multipliers sets the gauge fields to zero and gives the original action back. Alternatively one can integrate out the gauge fields, giving the dual action. Explicitly: consider the bosonic part of the UHM,
293: \equ{
294: e^{-1}\cL_{UH} =-\2 G_{\hA\hB} \pu{\gm}\gf^{\hA} \p{\gm}\gf^\hB -R -\4 F^{\gm\gn}F_{\gm\gn} \nn \km
295: }
296: in which $G_{\hA\hB}$ is the scalar (quaternionic) metric one can read off from \R{UHM1}. Now we gauge two isometries by replacing $\p{\gm}\gf^I \to \p{\gm}\gf^I -A_{\gm}^I$ and by adding the Lagrange multiplier term $H_I^\gm A_\gm^I \,,\,I=1,2$. The Lagrangian density becomes
297: \bea
298: e^{-1}\cL &=&-\2 G_{AB}\p{\gm}\gf^A \pu{\gm}\gf^B -G_{AI} \p{\gm}\gf^A\left(\pu{\gm}\gf^I -A^{\gm I}\right) -R -\4 F^{\gm\gn}F_{\gm\gn} \nn \\[1mm]
299: &&\, -\2 G_{IJ}\left(\p{\gm}\gf^I -A_\gm^I\right) \left(\pu{\gm}\gf^J -A^{\gm J}\right) -H^\gm_I A_\gm^I \label{eq:temp2a} \km
300: \eea
301: remember that $A=1,2$.
302: The next step is to integrate out the gauge field, the equation of motion for $A_\gm^I$ is
303: \equ{
304: A^I_\gm=A^I_A\p{\gm}\gf^A + \p{\gm}\gf^I -H_\gm^I \nn \km
305: }
306: \label{sec:page1} where $A^I_A\equiv M^{IJ} G_{JA}$. Inserting this back into \R{temp2a} gives
307: \equ{
308: e^{-1}\cL_{DT}=-\2 \cG_{AB}\pu{\gm} \gf^A \p{\gm} \gf^B + \2 M^{IJ} H_I^\gm H_{\gm J} -R-\4 F^{\gm\gn}F_{\gm\gn} \label{eq:DTM2}
309: }
310: where $M^{IJ}$ has been defined in \R{tensormetric1}, the relation with the UHM metric $G_{\hA\hB}$ is $M^{IJ}\equiv (G_{IJ})^{-1}$. The metric for the $2$ remaining scalars is also a subset of $G_{\hA\hB}$, namely
311: \equ{
312: \cG_{AB}=G_{AB} -G_{AI} M^{IJ} G_{JB} = \left(%
313: \begin{array}{cc}
314: 1 & 0 \\
315: 0 & e^{-\gf} \\
316: \end{array}%
317: \right)
318: \label{eq:GAB1} \pt
319: }
320: Together these components form the hypermultiplet metric, or the other way around: the hypermultiplet metric decomposes as
321: \equ{
322: \label{eq:dual-metric}
323: G_{\!\hat{A}\hat{B}} = \lp \cG_{AB} + A_A^I M_{IJ} A_B^J & A_A^K
324: M_{KJ} \\[2mm] M_{IK} A_B^K & M_{IJ} \rp \pt
325: }
326: We have obtained the DTM by dualizing two scalars from the UHM to tensors. Perturbatively, the DTM guarantees\footnote{Because the Heisenberg group is preserved in perturbation theory.} two commuting shift symmetries in the dual UHM description. Nonperturbatively, however, the duality also involves the constant modes of the dual scalars $\gvf$ and $\gs$ by means of theta-angle-like terms. These are surface terms which have to be added\footnote{The dualization process is defined up to such surface terms.} to the DTM Lagrangian and are non-vanishing in the presence of instantons and anti-instantons. In Euclidean space they can be written as integrals over $3$-spheres at infinity:
327: \equ{
328: S_{\textrm{surf}}^E=\I \left(\gvf \int_{S^3_\infty} H_1 + \gs\int_{S^3_\infty} H_2 \right)=\I \gvf Q_1 + \I \gs Q_2 \label{eq:boun1} \km
329: }
330: The charges $Q_1$ and $Q_2$ are related to the membrane and fivebrane instantons associated with the two tensors $H_1$ and $H_2$, as we will discuss in chapter \ref{ch2b} (see \R{instcharges1}). $\gvf$ and $\gs$ are now some parameters that play the role of coordinates on the moduli space in the dual UHM theory.
331: The dualization back to the UHM is performed by promoting $\gs$ and $\gvf$ to fields that serve as Lagrange multipliers enforcing the Bianchi identities on the tensors. This gives back the fields $\gvf$ and $\gs$ in the UHM, in the DTM there are no $\gvf$ and $\gs$ fields because they have been dualized to two tensors.\\
332: Boundary terms such as these are also added in $3$-dimensional gauge theories in the Coulomb phase where the effective theory can be described in terms of a vector which can be dualized into a scalar (the dual photon) in a similar manner as described above, see \cite{Seiberg:1996nz, Dorey:1997ij, Dorey:1998kq} and in particular \cite{Polyakov:1976fu, Polyakov:1987ez}. \\
333:
334: \section{Wick rotations and duality}
335: \label{wickrotI}
336: Having constructed the DTM we still need to analytically continue it. We use the Wick rotation to analytically continue (the details are given in appendix \ref{appwick}). The result is
337: \equ{
338: e^{-1}\cL^E_{DT}=\2 \pu{\gm}\gf\p{\gm}\gf + \2 e^{-\gf}\pu{\gm}\gc \p{\gm}\gc + \2 M^{IJ} H_I^\gm H_{\gm J} + R+\4 F^{\gm\gn}F_{\gm\gn} \label{eq:DTM3} \pt
339: }
340: In performing our Wick rotation we have assumed that we are dealing with a class of metrics for which we can perform this procedure. In general this is a subtle issue. However, in the next chapter we will limit ourselves to flat space, for which the rotation is well defined. We will do so because the instanton configurations we will discuss must have vanishing energy momentum tensor. This requirement will be satisfied by working in flat space. We see that in general the presence of the Einstein-Hilbert term causes our otherwise positive definite Euclidean action to be unbounded from below. Again, in flat space this is not an immediate problem since the Ricci scalar vanishes. However, gravitational fluctuations around the instanton configuration could ruin this. \\
341: From this Lagrangian two Bogomol'nyi equations can be derived, one corresponding to a NS $5$-brane and another to a membrane instanton, which will be the subject of the next chapter. However before continuing, we will discuss the reason for working with the Euclidean DTM, as promised.\\
342:
343: Apart from the Einstein-Hilbert term, the action corresponding to \R{DTM3} is positive definite and has real saddle points for the $3$-forms, which we will construct in the next chapter. In principle one could work with the UHM as well, but then one would have to deal with imaginary saddle points. To see this, let us reconsider the dualization process of a single (for simplicity) $3$-form $H$ to a scalar, the generalization is straightforward.\\
344: In Minkowski space the path integral involving a $2$-form is given by
345: \bea
346: &&\int [dH] e^{\I S_H} \prod_x \gd[dH(x)] \label{eq:D1}\\
347: && = \int [dH] \int [da] \exp\left( \I S_H + \frac{\I}{3!} \int d^4 x\, a \,\gve ^{\gm\gn\gr\gs}\p{\gm}H_{\gn\gr\gs}\right) \nn \pt \eea
348: The Bianchi identity is enforced by including a delta function in the path integral, which is given in its functional representation in the second line. Note that $a$ is just a (real) dummy variable at this point. The action is given by
349: \equ{
350: S_H=\frac{1}{3!}\int d^4 x \sqrt{-g} \left(-\2 H_{\gm\gn\gr}H^{\gm\gn\gr}\right) \nn \pt
351: }
352: The functional integral over $H$ is a Gaussian integral and can be explicitly performed. To do so shift the integration variable:
353: \equ{
354: H_{\gm\gn\gr}=\frac{1}{\sqrt{-g}}\,\gve_{\gm\gn\gr\gs}\pu{\gs}a + h_{\gm\gn\gr} \nn \km
355: }
356: where the first part is the familiar dualization relation between a $3$-form and a pseudoscalar in $4$ dimensions. The exponent in \R{D1} becomes
357: \equ{
358: -\frac{\I}{2}\int d^4 x \,\p{\gm}a\pu{\gm}a -\2\frac{\!\I}{3!}\int d^4 x \sqrt{-g}\,h_{\gm\gn\gr}h^{\gm\gn\gr} \nn \pt
359: }
360: The functional integral over the fluctuations $h_{\gm\gn\gr}$ can easily be performed since they appear undifferentiated, so we are left with the dualization relation
361: \equ{
362: H_{\gm\gn\gr}=\frac{1}{\sqrt{-g}} \, \gve_{\gm\gn\gr\gs}\pu{\gr}a \nn \pt
363: }
364: In Euclidean space, there is no factor of $\I$ in front of $S_H$, see \R{D1}. This has the consequence that one obtains an extra factor of $\I$ in the dualization relation:
365: \equ{
366: H^E_{\gm\gn\gr}=\frac{\I}{\sqrt{g}} \, \gve_{\gm\gn\gr\gs}\pu{\gr}a \nn \pt
367: }
368: This factor of $\I$ is crucial in ensuring that a positive $3$-form action in Euclidean space dualizes correctly to a positive scalar action. \\
369: Hence we see that performing a real saddle point calculation for the DTM in Euclidean space, corresponds to an imaginary saddle point calculation for the UHM, for more details see \cite{Burgess:1989da}.\\
370: Note that this factor of $\I$ one needs in the integral representation of the delta function in Euclidean space is exactly provided for by the way we define our Wick rotation. That is to say, after Wick rotating our Lagrange multiplier term precisely has the right factor of $\I$, see \R{tempa21}, and we thus have a consistent framework: dualizing commutes with the analytic continuation.
371:
372: \clearpage{\pagestyle{empty}\cleardoublepage}
373:
374:
375:
376:
377:
378:
379:
380:
381:
382:
383:
384:
385: