Brought to you by:
Paper

Homogeneous heterotic supergravity solutions with linear dilaton

Published 11 January 2012 © 2012 IOP Publishing Ltd
, , Citation Christoph Nölle 2012 J. Phys. A: Math. Theor. 45 045402 DOI 10.1088/1751-8113/45/4/045402

1751-8121/45/4/045402

Abstract

I construct solutions to the heterotic supergravity BPS equations on products of Minkowski space with a non-symmetric coset. All of the bosonic fields are homogeneous and non-vanishing, the dilaton being a linear function on the non-compact part of spacetime.

Export citation and abstract BibTeX RIS

1. Introduction

The aim of this paper is to present some homogeneous solutions to the heterotic supergravity BPS equations [8, 7]

Equation (1.1)

and the Bianchi identity

Equation (1.2)

where ε is a spinor on a ten-dimensional Lorentzian manifold M, H is a three-form, ϕ a function and F the curvature of a gauge field on M, and ∇± are two connections on the tangent bundle TM, involving H. In (1.1), γ is the map from forms to the Clifford algebra. These equations are to be thought of as first order equations that imply the heterotic equations of motion, which are second order partial differential equations. The full string theory equations also contain an infinite tower of higher order α'-corrections, that we have truncated at order (α')1. As will be discussed in section 2, solutions to (1.1) and (1.2) in general satisfy the equations of motion only up to (α')2-corrections, and should be viewed as perturbative approximations to a string background. Hence, a rigorous proof that the solutions presented below give rise to full heterotic string backgrounds would probably require a construction of the worldsheet conformal field theory. The generality and simplicity of the solutions, however, can already be taken as an indication of its existence.

The manifold M will be chosen of the form $M= \mathbb R^{p,1}\times G/K$ for a non-symmetric, naturally reductive coset G/K, with (mostly) simple compact Lie groups KG, equipped with the metric induced by the Killing form. This choice is canonical for the following reasons. First of all every such non-symmetric coset carries a G-invariant (or homogeneous) three-form, which we will identify with H. Upon proper normalization of H, the question whether the gravitino equation ∇ε = 0 has a solution ε turns into a simple representation-theoretical problem. Furthermore, upon this choice of H we also get a solution for F, namely the curvature R of the so-called canonical connection ∇, which also appears in the gravitino equation. It satisfies both γ(R)ε = 0 if ε solves the gravitino equation, and dH ∼ tr(R+R+RR), leading to a solution of all of the equations except the dilatino one, $\gamma ({\rm d}\phi -\frac{1}{12}H)\varepsilon =0$.

All this suggests that the heterotic supergravity equations are tailored to admit homogeneous solutions; in particular, the Bianchi identity allowing for non-trivial dH is an important deviation from the standard supergravity rule dH = 0, which would immediately rule out these spaces. The situation changes with the dilatino equation however. On a general coset G/K there are no homogeneous one-forms which could serve as dϕ, and if they exist they tend to be non-exact. Therefore, the only obvious choice would be to take dϕ = 0. This is not possible, however, as we have γ(H)ε ≠ 0, and there is also a simple no-go theorem excluding this type of solutions.

In [30] we proposed to circumvent this problem by allowing for non-trivial fermion condensates; in this paper, a solution based on purely bosonic backgrounds will be developed. The backdoor we are going to use is to introduce an additional $\mathbb R^p$ factor to our spacetime M, with p the rank difference of G and K (or p = 2 for equal rank groups). Then, we can take ϕ to be a linear function on $\mathbb R^p$, giving rise to a constant and thus homogeneous one-form dϕ. It will be shown below that upon this choice of ϕ it is often possible to solve also the last equation. The amount of supersymmetry preserved in the space orthogonal to $\mathbb R^p \times G/K$ is then at least $\mathcal N= 2^p -p$. For trivial K, one gets a Wess–Zumino–Witten (WZW) model coupled to a linear dilaton, and these models also exist in type II string theory. They were considered already in [16], as a certain limit of NS5 branes.

From a physical perspective, the linear dilaton certainly rules out these spaces as models for our universe; but see [27] for an intersecting brane scenario with chiral fermions. An interpretation of the type of solution considered here in terms of a decoupling limit of string theory is given in [3, 21].

On the other hand, a couple of homogeneous solutions to the above equations have been presented in recent years where the dilaton is actually constant [18, 33]. The method used in these works is somewhat different from ours, as they take as starting point Strominger's reformulation of the BPS equations [32]. Furthermore, they avoid the above-mentioned no-go theorem by choosing M to be a non-semisimple Lie group (or a finite quotient thereof) equipped with a metric of negative scalar curvature, whereas in our models the metric always comes from the bi-invariant one on G and has positive scalar curvature. Although many of the spaces we discuss allow for other homogeneous metrics as well, the solution of the equations becomes more involved with these. One advantage of not relying on Strominger's equations is that we are not restricted to compact spaces of dimension 6. In fact, we will find solutions with compact spaces of arbitrary odd dimension, and also six-dimensional ones.

After introducing the necessary tools for homogeneous spaces in section 3, we will in section 4 develop a method which allows us both to prove existence of ∇-parallel spinors on many of the considered spaces and to calculate the action of γ(H) on these spinors, thus enabling us to determine the linear dilaton needed to satisfy also the dilatino equation. In section 5, we discuss homogeneous Sasaki–Einstein manifolds, which are a particular class of spaces where this method can be applied. Section 6 has some examples treated in detail, based on the cosets

  • SU(n + 1)/SU(n) = S2n + 1
  • Sp(n + 1)/Sp(n) = S4n + 1
  • Sp(n)/SU(n)
  • SO(2n)/SU(n)
  • SO(n + 1)/SO(n − 1)
  • Spin(7)/G2 = S7
  • G2/SU(3) = S6
  • SU(3)/U(1)×U(1)
  • SO(5)/SO(3)max.

It is intriguing that all of those spaces admit one of the following structures:

  • nearly Kähler (in 6D),
  • nearly parallel G2 (in 7D),
  • Sasaki–Einstein (in odd dimension),
  • 3-Sasaki (in 4n + 3 dimensions),

although the metric we use in most cases differs from the one defining this structure. These are exactly the spaces whose cones admit parallel spinors, and they play an important role in other types of string theory as well [1, 12]. The amount of supersymmetry preserved depends on the geometric type of the manifold; nearly Kähler, G2, and Sasaki–Einstein generically have $\mathcal N=1$, whereas 3-Sasakian spaces preserve more supersymmetry.

It should be mentioned that the method presented does not generalize to symmetric spaces, like Sn = SO(n + 1)/SO(n) with its round metric; the equation ∇ε = 0 does not have a solution there.

2. Heterotic supergravity

The low-energy limit of heterotic string theory is given by 10D $\mathcal N=1$ supergravity coupled to super-Yang–Mills. The bosonic part of the effective action is [7]

Equation (2.1)

where we adopt the widely used convention to denote by tr a positive-definite form on a Lie algebra, in fact always minus the ordinary trace over tangent space in our examples. It leads to the following field equations (to order α'):

Equation (2.2)

The full action with fermions is invariant under supersymmetry, acting on the fermions as

Equation (2.3)

where ψ is the gravitino, λ the dilatino and χ the gaugino. The quantization map γ is explicitly

Equation (2.4)

where we use the convention {γμ, γν} = 2gμν. The requirement that these variations vanish ensures that a background preserves supersymmetry and is precisely the set of equations (1.1). Here the connections ∇± are related to the Lévi-Cività connection of g via

Equation (2.5)

In addition to the equations of motion or supersymmetry equations, one has to impose the Bianchi identity

Equation (2.6)

The BPS equations (1.1) together with the Bianchi identity (2.6) imply the equations of motion of heterotic supergravity up to terms of order (α')2 [25]. It has been shown in [25] that the (α')2-corrections do not vanish, unless one replaces the curvature form R+ in the Bianchi identity and the equations of motion by the curvature of an instanton gauge field on the tangent bundle. From a string theory point of view, such a replacement is not justified; however, instead one would have to take into account the full tower of α'-corrections to the equations to obtain a consistent theory, as explained in [7]. We will solve the BPS equations and the Bianchi identity exactly at order (α')1; hence, our solutions satisfy the equations of motion only up to (α')2 corrections, and should be viewed as perturbative approximations to an exact string background. But those equations which do not involve the gauge field will be satisfied (and the Yang–Mills equation for F as well). They are the H-equation d*e−2ϕH = 0 and the following combination of dilaton equation and trace of the Einstein equation:

Equation (2.7)

From this, we can derive a simple no-go theorem. Suppose the dilaton is constant and H has only spacelike components; then,

Equation (2.8)

and the scalar curvature must be non-positive. It should be mentioned that there is also a constraint on the cohomology class of H:

Equation (2.9)

if dH = 0, which leads to the quantization of the level in WZW models for instance. For dH ≠ 0, the requirement will be that a certain combination of H and the Chern–Simons forms of R+ and F defines an integer cohomology class, but we will simply ignore this condition in what follows, as most of the spaces we consider have $H^3(M,\mathbb Z)=0$ anyway.

3. Homogeneous vector bundles

Let G be a connected compact simple Lie group equipped with the bi-invariant Riemannian metric g (induced by minus the Killing form on its Lie algebra $\mathfrak g$), and K a naturally reductive subgroup. This means we have an orthogonal splitting of the Lie algebra $\mathfrak g = \mathfrak k\oplus \mathfrak m$ with ad$(\mathfrak k)\mathfrak m \subset \mathfrak m$, so that $\mathfrak m$ carries a representation of $\mathfrak k$. Let (V, ρ) be a representation of K, and E = G × KV the associated vector bundle over G/K, which consists of equivalence classes [g, v] with gG and vV, and identification [g, v] = [gk−1, ρ(k)v] for all kK. Its sections are in a 1–1 correspondence with maps f: GV satisfying

Equation (3.1)

G acts on the space of sections Γ(G/K, E) through (g · f)(h) = f(g−1h). The set of G-invariant sections (also called homogeneous sections) is thus given by the constant functions, and therefore in a 1–1 correspondence to the K-invariant elements of V.

Lemma 3.1. Let V carry a representation of K; then,

We will adopt the following index convention. Basis elements of $\mathfrak k$ will be denoted by Ik, Il, ..., those of $\mathfrak m$ by Ia, Ib, ..., and the full set of basis elements of $\mathfrak g$ by Iμ, Iν, .... The dual basis of left-invariant one-forms on G is denoted eμ, or ek and ea for those dual to Ik and Ia. The pull backs of these forms to G/K will be denoted by the same symbols, and they satisfy the Maurer–Cartan equations

Equation (3.2)

where fλμν are the structure constants of $\mathfrak g$, defined by [Iμ, Iν] = fλμνIλ. Our metric on $\mathfrak g$ will be minus the Killing form

Equation (3.3)

or in coordinates

Equation (3.4)

g is $\mathfrak g$-invariant, thus also $\mathfrak k$-invariant, and gives rise to a homogeneous metric on G/K.

The three-form.

Another important example of a homogeneous section is the following. Define $H \in \Lambda ^3 \mathfrak m^*$ through

Equation (3.5)

or in coordinates

Equation (3.6)

Then, H is K-invariant and gives rise to a three-form on G/K. In case K is chosen trivial, H becomes a generator of $H^3(G,\mathbb Z)=\mathbb Z$ upon proper normalization of the metric g. In general, H is a natural candidate for the three-form of heterotic string theory. For the Bianchi identity, we need to know the derivative of H, and for its equation of motion d*H (using the notation eabcd = eaebeced).

Lemma 3.2. We have

Equation (3.7)

whereas d*H = 0.

Proof. From the Maurer–Cartan equation, we have

Consider the last term. It follows from the Jacobi identity that fabcfadk splits into a part which is symmetric in b and d, and another part symmetric in c and d. Therefore, this term vanishes. Using again a Jacobi identity, we conclude that

d*H. We assume the Iμ to form an orthonormal basis, such that fμνλ is totally antisymmetric. Furthermore, we will not keep track of whether an index is up or down, but rather sum over any index appearing more than once. Then, we have

with the derivative

Equation (3.8)

The first term is easily seen to vanish: b and c only run over the values of a1, a2 and a3, giving contributions of the type

where the two f factors are symmetric in a3 and a4, and thus vanish. Now let us consider the second contribution in (3.8). We have

which vanishes due to the Jacobi identity again.

Connections

Due to the identification $T^*(G/K)= G\times _K\mathfrak m^*$, a connection on a homogeneous vector bundle G × KV can be considered as a map

satisfying additional properties. The simplest example is the so-called canonical connection ∇, acting as

Equation (3.9)

Here XL is the left-invariant vector field on G corresponding to X. In a trivialization of T(G/K) induced by a local map G/KG, which allows us to pull back the left-invariant one-forms on G to locally defined one-forms on G/K, its connection form is given by

Equation (3.10)

with dρe the differential of ρ: K → Aut(V) at the identity. As is clear from the definition, the parallel sections of ∇ correspond to constant functions, and thus to K-invariant elements of V.

Lemma 3.3. Let V carry a K representation; then, the parallel sections of G × KV wrt ∇ are in a 1–1 correspondence with K-invariant elements of V, and by lemma 3.1 are precisely the G-invariant sections.

As the notation suggests, we will identify ∇ with the connection appearing in the gravitino equation ∇ε = 0, and thereby translate the problem of solving this differential equation into a representation-theoretical one for the holonomy group K. Besides ∇, we have some further homogeneous connections on T(G/K), the Lévi-Cività connection ∇ of g, and the connection ∇+, given by

Equation (3.11)

Explicitly, one finds [30]

Equation (3.12)

The structure group for these connections is generically SO($\mathfrak m)$, but ∇ has the structure group K⊂ SO($\mathfrak m)$. Their curvatures, as elements of End($\mathfrak m)\otimes \Lambda ^2\mathfrak m^*$ and in coordinates, are

Equation (3.13)

with $\pi _\mathfrak k:\mathfrak g\rightarrow \mathfrak k$ the orthogonal projection. For the Bianchi identity we need to know tr(R+R+) and possibly tr(RR). These are given by

Equation (3.14)

where we introduced the (negative) Killing form $\langle \cdot .\cdot \rangle _\mathfrak k$ of the subalgebra $\mathfrak k$, and similarly

Using the result of lemma 3.2, we conclude that

Equation (3.15)

which is almost the Bianchi identity. It looks however as if we need to put α' = 4 to solve the Bianchi identity, but this is due to our arbitrary normalization of the metric on G/K. Note that the lhs of (3.15) is completely scale-independent, whereas the rhs scales with the same factor as the metric. Therefore, the Bianchi identity really fixes the scale in terms of α'.

Spinors

The spin bundle on a homogeneous manifold is constructed as follows. Ad-invariance of the Killing form implies that $\mathfrak k$ acts orthogonally on $\mathfrak m$, giving rise to an embedding

Equation (3.16)

which can be composed with the spin representation ${\rm d}S: \mathfrak {so}(\mathfrak m)\rightarrow \mathfrak {spin}(\mathfrak m)$, to give $\widetilde{{\rm ad}}:= {\rm d}S\,\circ \,$ad$_\mathfrak m$. We assume that this lifts to a representation of K, a sufficient condition for this being that K is simply connected. Denoting the spinor space over $\mathfrak m$ by S (also $S(\mathfrak m)$ occasionally), we get an associated bundle

Equation (3.17)

which is the spinor bundle over G/K. The connections we considered before give rise to connections on $\mathcal S$, and the parallel sections wrt ∇ correspond to K-invariant elements of S. To determine whether there exist parallel spinors, we thus need to know whether the trivial representation of K (or $\mathfrak k$) occurs in the decomposition of the spinor representation S over $\mathfrak m$ into irreducibles, which is a purely algebraic task.

Suppose then that ε is parallel wrt ∇, so that also Rε = 0. Then, it follows from the symmetry property Rabcd = Rcdab that R annihilates ε under the Clifford action as well, γ(R)ε = 0, which makes R a candidate for F solving the gaugino equation γ(F)ε = 0. We have seen before that it is also an excellent candidate to solve the Bianchi identity.

The following commutation relation between the quantized three-form and elements of $\mathfrak k$ acting on spinors over $\mathfrak m$ will be useful.

Lemma 3.4. For $X\in \mathfrak k$ we have

Equation (3.18)

Proof. A simple calculation in the Clifford algebra shows that

Equation (3.19)

but this is proportional to γ(ad(X) · H), where · denotes the action of $\mathfrak {so}(\mathfrak m)$ on $\Lambda ^3 \mathfrak m^\ast$, and we know that H is invariant under this action of $\mathfrak k$.

This means that γ(H) leaves the set of invariant elements in $S(\mathfrak m)$ invariant, so if there is only one invariant spinor, γ(H) maps it to a multiple of itself.

The dilaton

In the supergravity equations, only the differential dϕ occurs, and if we impose homogeneity again it gives rise to a K-invariant element of $\mathfrak m^*$. Often there do exist K-invariant elements in $\mathfrak m^*$ if the rank of K is smaller than the rank of G, which correspond to Cartan generators orthogonal to $\mathfrak k$, but the associated one-forms on G/K are not exact, and therefore not suitable for our purpose. We have to conclude that dϕ = 0 is the only admissible solution for the dilaton.

On the other hand, we have seen that a vanishing dilaton is not compatible with positive scalar curvature, which is why we will have to introduce a linear dilaton on an additional $\mathbb R^p$ factor of the total manifold to obtain a solution of all the supergravity equations.

Symmetric spaces

Suppose G/K is symmetric, meaning that $[\mathfrak m,\mathfrak m]\subset \mathfrak k$. Then, H = 0, and from the dilatino equation we also have that dϕ = 0. The equation ∇ε = 0 then tells us that there is a parallel spinor, implying that M is Ricci-flat [2], which is impossible for symmetric spaces. Hence, G/K must be non-symmetric.

4. Spinors on cosets

Representation-theoretic method.

Given a coset G/K, we need to determine whether the spin representation over $\mathfrak m=\mathfrak g/\mathfrak k$ contains the trivial $\mathfrak k$-representation as an irreducible component. Suppose for the moment that $\mathfrak k$ is simple, and denote the set of weights of $\mathfrak m$ by $\Omega (\mathfrak m)$, whereas $\Omega ^+(\mathfrak m)$ contains only the positive ones. Then, the weights that appear in $S(\mathfrak m)$ are of the form

Equation (4.1)

where all combinations of signs appear (this can be understood from lemma 4.1 below). Now only the dominant weights can be highest weights of an irreducible representation of $S(\mathfrak m)$, so it is often enough to determine all the dominant weights in $\Omega (S(\mathfrak m))$. Then, a couple of situations can occur. If the zero weight is not in $\Omega (S(\mathfrak m))$, then the trivial representation is not contained. If it is, one can sometimes conclude by dimensional reasoning that the trivial representation must or must not occur as a component. Sometimes the situation is even simpler.

Consider the coset SO(n + 1)/SO(n) = Sn. In this case, $\mathfrak m$ is simply the fundamental representation of $\mathfrak {so}(n)$, and it follows that $S(\mathfrak m)$ is the (Dirac) spinor representation, which is often reducible, but does not have invariant elements. Thus, Sn with its standard round metric does not admit a homogeneous solution, in accordance with our general result for symmetric spaces.

Despite its elegance, we will not employ the representation-theoretic method in the following, but use the more down-to-earth approach explained in the next paragraph. The reason for that is that the latter method allows us to determine how the three-form H acts on invariant spinors, and thereby how to choose the dilaton appropriately to also solve the dilatino equation $\gamma ({\rm d}\phi -\frac{1}{12} H)\varepsilon =0$. A drawback is that the method is not always applicable, as explained below.

Direct method for lower-rank subgroups.

Recall that the action of $\mathfrak k$ on $\mathfrak m$ defines an embedding $\mathfrak k\subset \mathfrak {so}(\mathfrak m)$. Here we give an explicit construction of the spinor space for the case that $\mathfrak k\subset \mathfrak {su}(\mathfrak m)$ for a well-chosen complex structure on $\mathfrak m$. We assume for the time being that rk$(\mathfrak k)< $ rk$(\mathfrak g)$, although for certain maximal rank subgroups we will be able to generalize our construction. A particular example where this is possible is the case where G/K is a six-dimensional nearly Kähler manifold.

A further assumption we want to make is that $\mathfrak k$ and $\mathfrak g$ admit a common root space decomposition, i.e. they are of the form

Equation (4.2)

where the Cartan algebra of $\mathfrak g$ is

Equation (4.3)

with r = rk$(\mathfrak g)>$ rk$(\mathfrak k)=r-p$, and the set of positive roots of $\mathfrak g$ is denoted by R+, whereas those of $\mathfrak k$ are contained in S+R+. The complement $\mathfrak m$ of $\mathfrak k$ in $\mathfrak g$ is then given by

Equation (4.4)

which is usually not a Lie algebra. Our final assumption is that the roots in R+S+ are higher than those in S+ (which roots are highest, i.e. more positive than others, is a convention, but it is not always possible to choose the roots in this way; typically if rk$(\mathfrak k)=$ rk$(\mathfrak g)$ then this ordering is not possible). By restriction, we will consider roots of $\mathfrak g$ as roots of $\mathfrak k$ as well. Further we assume a basis of root vectors Eα of the $\mathfrak g_\alpha$ and E−α of the $\mathfrak g_{-\alpha }$ to be chosen, obeying the commutation relation

Equation (4.5)

where Nαβ are constants, and the normalization can be chosen as [19]

Equation (4.6)

Now extend the metric of $\mathfrak m$ to one on $\mathfrak m^{\prime }:=\mathfrak m \oplus \mathbb R^p$ by using the standard metric on $\mathbb R^p$. Furthermore, we extend the action of $\mathfrak k$ on $\mathfrak m$ trivially to one on $\mathfrak m\oplus \mathbb R^p$, so that we still have

Equation (4.7)

and the spin representation ${\rm d}S: \mathfrak {so}(\mathfrak m^{\prime }) \rightarrow \mathfrak {spin}(\mathfrak m^{\prime })$ embeds $\mathfrak k$ into the spin algebra. A complex structure J on $\mathfrak m^{\prime }$ is obtained by defining holomorphic vectors to be given by the positive root vectors Eα, α ∈ R+S+, as well as the following combinations of Cartan vectors Hj and standard basis vectors ∂j of $\mathbb R^p$:

Equation (4.8)

whereas anti-holomorphic ones are given by the negative root vectors E−α and the $ \overline{E}_j = H_j + {\rm i} \partial _j$. The Hj are assumed to be elements of the real Lie algebra $\mathfrak g$, so that roots wrt them are purely imaginary. From the commutation relation (4.5) and our ordering of the roots, we deduce the following relations:

Equation (4.9)

where $\mathfrak m^+$ consists of holomorphic vectors in $\mathfrak m^{\prime }\otimes \mathbb C$, and $\mathfrak m^-$ of anti-holomorphic ones. Thus, our subalgebra $\mathfrak k$ commutes with the complex structure, and is contained in the unitary subalgebra of $\mathfrak {so}(\mathfrak m^{\prime })$:

Equation (4.10)

It follows from this that the complex structure J extends to an almost-complex structure on the manifold $\mathbb R^p\times G/K$, which is just a different way of saying that its structure group is contained in $U(\mathfrak m^{\prime })$. Let eα be the holomorphic one-forms dual to the Eα, and ej be dual to Ej, and denote by W the space of all holomorphic one-forms on $\mathfrak m^{\prime }$. Then, the spinor space is the exterior algebra

Equation (4.11)

The Clifford algebra Cl$(\mathfrak {m^{\prime }}^*)$ is the quotient of the tensor algebra over ${\mathfrak m^{\prime }}^*$ by the relation

Equation (4.12)

where g−1 is the metric induced by g on the dual ${\mathfrak m^{\prime }}^*$. As a vector space, the Clifford algebra is isomorphic to the exterior algebra $\Lambda \mathfrak {m^{\prime }}^*$, but not as an algebra of course. The vector space isomorphism is just the quantization map γ. Cl$(\mathfrak {m^{\prime }}^*)$ acts on S as follows; associate with every holomorphic one-form eα an operator ζα := γ(eα), an operator $\overline{\zeta }^\alpha =\gamma ({\rm e}^{-\alpha })$ to every anti-holomorphic one-form e−α, as well as ζj = γ(ej) to ej and $\overline{\zeta }^j=\gamma (\overline{e}^j)$ to $\overline{e}^j$. They act on S = ΛW as creation and annihilation operators:

Equation (4.13)

where $\breve{w}^{b_i}$ means leaving out the element, and the wa are elements of W. This action can be uniquely extended to one of the full Clifford algebra. Inside the Clifford algebra Cl(2n) := Cl($\mathbb R^{2n})$ sits the spin algebra $\mathfrak {spin}(2n)$, which is the image of the map

Equation (4.14)

where γa = γ(dxa). One can check that this action of the orthogonal Lie algebra restricts to the usual action of $\mathfrak {su}(n)$ on $\Lambda \mathbb C^{n*}$, in particular, the decomposition

is preserved by this subalgebra, and then even by $\mathfrak u(n)$, although the additional $\mathfrak u(1)$-component acts in a non-standard way. To show that $\mathfrak k$ leaves at least one spinor invariant, it is therefore enough to show that it is contained in $\mathfrak {su}(\mathfrak m^{\prime })$, as this acts trivially on S0Sn. We can easily determine the action of the Cartan generators of $\mathfrak k$ on spinors (recall $\widetilde{{\rm ad}}(X)=$ dS ○ ad(X) for $X\in \mathfrak k$).

Lemma 4.1. We have for the $H_j\in \mathfrak k$, j = p + 1, ..., r

Equation (4.15)

Proof. According to the definition of dS, we get

where we introduced [Hj, Eα] ≕ (Hj)βαEβ. But this just means that (Hj)βα = δβαα(Hj) and ${(H_j)^{\overline{\beta }} }_{\overline{\alpha }} = -{\delta ^\beta }_\alpha \alpha (H_j)$. Furthermore, $g_{\alpha ,\overline{\beta }}=g_{\alpha ,-\beta }=2\delta _{\alpha \beta }$.

Proposition 4.2. Under the assumptions of this section suppose that

Equation (4.16)

as roots of $\mathfrak k$. Then, $\mathfrak k$ acts trivially on the completely empty and completely filled states

where $m=\dim (\mathfrak m^{\prime })$.

Proof. We know already that $\mathfrak k\subset \mathfrak u(\mathfrak m^{\prime })$ leaves invariant the decomposition S = ⊕lSl. The subrepresentations on S0 and Sm are one dimensional, and according to lemma 4.1 the Cartan elements of $\mathfrak k$ act by multiplication with

If these vanish, the representations are trivial.

A closer inspection shows that the trivial representation occurs with multiplicity at least 2p + 2, as $\mathfrak k$ acts trivially on all of the holomorphic vectors Ej. The requirement that the supersymmetry generator in heterotic supergravity be Majorana–Weyl will reduce the amount of supersymmetry preserved in the space orthogonal to $\mathbb R^p\times G/K$ to $\mathcal N=2^{p}$, whereas the dilatino equation reduces it even further.

A different way to write condition (4.16) is in terms of the Weyl vector $\rho =\frac{1}{2} \sum _{\alpha \in R^+}\alpha$. Then, it reads

Equation (4.17)

As we mentioned already, it is equivalent to $\mathfrak k\subset \mathfrak {su}(\mathfrak m)$. The condition cannot be satisfied by a subalgebra of maximal rank, as this would imply $\sum _{\alpha \in R^+\setminus S^+} \alpha =0$ as roots of $\mathfrak g$, contradicting the positivity of all the roots in R+S+. In this case, one can sometimes choose the complex structure in a different way, i.e. such that holomorphic vectors do not coincide with positive root vectors, and apply the procedure presented here analogously. This will be demonstrated for G2/SU(3) below.

Summarizing the results of this section, we have proven that given a naturally reductive subalgebra $\mathfrak k\subset \mathfrak g$ of lower rank, admitting a common root space decomposition with $\mathfrak g$ of the form (4.2) such that the roots not belonging to $\mathfrak k$ are all higher than those of $\mathfrak k$, and such that their sum acts trivially on the Cartan generators of $\mathfrak k$, it admits invariant spinors over $\mathfrak m\oplus \mathbb R^p$, where p = rk($\mathfrak g)-$ rk($\mathfrak k$).

With lemma 3.3, we conclude that there are invariant, and thereby ∇-parallel spinors over $\mathbb R^p \times G/K$. Furthermore, we have a very explicit construction of these spinors, which will allow us to also solve the dilatino equation in this case.

The quantized three-form.

For the dilatino equation, we also need to know how the three-form H acts on the invariant spinors, i.e. we have to determine γ(H)ε for ε ∈ S0Sm. Using the commutation relations (4.5), we can determine H:

Equation (4.18)

with hj dual to Hj. Under the quantization map γ the last line eliminates S0 and Sm, and therefore we only need to consider

Equation (4.19)

By construction the one-form hj is quantized to

Equation (4.20)

and we conclude that

Equation (4.21)

so that

Equation (4.22)

Therefore, the dilatino equation $ \gamma ({\rm d}\phi - {\textstyle \frac{1}{12}}H) \varepsilon =0$ is not satisfied if we choose ϕ = 0 and ε ∈ S0Sm. To overcome this problem we can introduce a linear dilaton on the additional $\mathbb R^p$-factor. Take the real valued

Equation (4.23)

Its differential acts on $\mathcal S$ as

Equation (4.24)

which implies

Equation (4.25)

How do we then construct the supersymmetry generator ε? For the whole construction to be useful in heterotic string theory, the dimension of $\mathbb R^p \times G/K$ must be less than 10, i.e. we need

Equation (4.26)

then, we complete this to a ten-dimensional Lorentzian manifold $M= \mathbb R^{9-l+p ,1} \times G/K$. The spinor bundle over M is the tensor product of the one over $\mathbb R^{l-1,1}$ and the one over $\mathbb R^p \times G/K$, and there is an (anti-linear) charge conjugation C: SS mapping S0 to Sm times a complement which does not affect the action of forms over $\mathbb R^p \times G/K$. Let η be any constant, positive chirality spinor on $\mathbb R^{l-1,1}$. Then, ε is chosen proportional to 1⊗η + C(1⊗η) with $1 \in S^0\simeq \mathbb C$, and this is parallel wrt ∇, annihilated by $\gamma ({\rm d}\phi -\frac{1}{12} H)$, and Majorana–Weyl, as required by heterotic string theory.

There is a little subtlety related to the amount of supersymmetry preserved. Suppose that the dimension of $\mathbb R^p\times G/K$ is such that it admits Majorana spinors, and that $\mathcal S^0$ is invariant. Then, there are at least two invariant Majorana spinors, given by

Equation (4.27)

These generate just one supersymmetry however, as two Majorana spinors η + Cη and i(η − Cη) over $\mathbb R^{l-1,1}$ tensored with our invariant spinors lead to the same set of spinors over the total ten-dimensional space.

What is then the amount of supersymmetry preserved in $\mathbb R^{9-l,1}$? Consider all those spinors that are obtained from the completely empty one by acting with an arbitrary number of creation operators ζj, j = 1, ..., p. These generate a complex 2p-dimensional space $\tilde{S}=\tilde{S}^0\oplus \cdots \oplus \tilde{S}^p$, which is invariant under γ(H) and γ(dϕ), and we chose ϕ such that $\tilde{S}^0=S^0$ is annihilated by $\gamma ({\rm d}\phi - \frac{1}{12} H)$. On the higher subspaces $\tilde{S}^k$ the condition $\gamma ({\rm d}\phi - \frac{1}{12} H)\psi =0$ becomes one linear equation, so that generically there should be 2pp invariant elements. Indeed one can check that $\tilde{S}^p$ is not invariant, whereas $\tilde{S}^2$ has one invariant spinor, and so on. We conclude that our solutions have

Equation (4.28)

supersymmetry, with equality if there are no additional invariant spinors over $\mathbb R^p\times G/K$ to those we constructed explicitly. In particular for p = 0 and p = 1 this leads to $\mathcal N\ge 1$, and these are the cases we will consider in the examples below. For equal rank subgroups, we will have to add a factor $\mathbb R$ by hand on which the linear dilaton lives, and the orthogonal space will be $\mathbb R^{8-\dim G+\dim K,1}$.

5. Homogeneous Sasaki–Einstein manifolds

This is a particular class of examples where the assumptions on the root space decomposition of $\mathfrak k$ and $\mathfrak g$ are satisfied. A good general reference on Sasaki spaces is the book by Boyer and Galicki [13], and a simple definition is to call a Riemannian manifold Sasaki–Einstein if its metric cone is a Calabi–Yau. Similarly, 3-Sasakian manifolds by definition have hyperkähler cones. According to theorem 11.1.13 of [13], every homogeneous Sasaki–Einstein manifold is the total space of a principal U(1)-bundle over a so-called generalized flag manifold. These latter spaces are by definition cosets of a Lie group G by the centralizer of a torus, and they carry a Kähler–Einstein structure [4]. The classification and an exhaustive list of examples for non-exceptional G can also be found in Arvanitogeórgos' book [4]. We now give an explicit method to obtain the U(1)-bundles, which leads to many of the examples considered below.

Let G be a again a connected compact simple Lie group with Lie algebra $\mathfrak g$, and Cartan subalgebra $\mathfrak h$. Choose the positive roots, and let β be the highest simple root. Denote again by R+ the set of positive roots, and by S+R+ the set of roots which are linear combinations of all the simple roots except β. Then, $\mathfrak g$ has a root space decomposition of the form

Equation (5.1)

and we define the subalgebras $\mathfrak t\subset \mathfrak c\subset \mathfrak g$:

Equation (5.2)

Furthermore, we need the space

Equation (5.3)

These definitions imply the following commutation relations:

Equation (5.4)

Denote by C, TG the corresponding Lie groups (with T ≃ U(1)); then, C is the centralizer of T in G, and TC is a normal subgroup. Thus, K := C/T is a group again, and its Lie algebra is the orthogonal complement of $\mathfrak t$ in $\mathfrak c$ wrt the Killing form. Due to the relation (G/K)/TG/C, we get a U(1)-fibration

Equation (5.5)

where the base space carries a homogeneous Kähler structure, with complex structure induced by the choice of positive roots. Let

Equation (5.6)

It vanishes on the Cartan generators of $\mathfrak k$. Defining the one-form $A \in \Omega ^1(G/K; \, \mathfrak t$) through

Equation (5.7)

with $H_\mathfrak t\in \mathfrak t$ and h ∈ Ω1(G/K) the left-invariant one-form corresponding to the dual of $H_\mathfrak t$, we get a connection on the bundle G/KG/C [29], as A satisfies

Equation (5.8)

Here HL is the left-invariant vector field corresponding to $H_\mathfrak t$, and Rt: G/KG/K, [g]↦[gt]. The curvature $\Omega \in \Omega ^2(G/K;\mathfrak t)$ of this connection is calculated using the Maurer–Cartan equation, and given by

Equation (5.9)

with (δ, α) the inner product on weights induced by minus the Killing form. Let ρ be the minimal non-trivial representation of $\mathfrak t$ on $\mathbb C$ which exponentiates to an action of T; it gives rise to an associated line bundle LG/C, whose curvature $\Omega ^L\in \Omega (G/C, i\mathbb R)$ is the pull-back of ρ(Ω) under any set of local sections. We obtain a Kähler structure on G/C by the choice $\omega = \frac{i}{2\pi }\,{\Omega ^L}$ for the Kähler form ω, and according to [4] the Kähler metric is Einstein. It follows that L is ample, and the classification result theorem 11.1.13 in [13] (and its proof) imply that G/K carries a Sasaki–Einstein structure.

What is more important for us is that the result $\delta |_{\mathfrak k\cap \mathfrak h}=0$ implies that there are parallel spinors on $G/K \times \mathbb R$, and all the conditions from our previous section are satisfied. Thus, we find solutions to the heterotic supergravity equations on all of these spaces. It should be mentioned that the Sasaki–Einstein metric is never the one induced by the Killing form, which we are using.

Not all of the examples given below fall into this class however. Besides the Einstein–Sasaki manifolds, we will also find solutions on nearly Kähler, nearly parallel G2 and 3-Sasakian manifolds. These latter do carry an Einstein–Sasaki structure as well, but they allow for more parallel spinors, and besides the U(1)-fibration over a Kähler manifold, they also possess an Sp(1)-fibration over a quaternionic Kähler manifold. The only example of this kind we include is Sp(n + 1)/Sp(n) = S4n + 3, but probably all of them could be used. The 3-Sasakian metric is related to the Killing form in [10], and there is a complete list of the homogeneous examples in any of [2, 1013].

6. Examples

We will treat the cases listed in the introduction, and additionally SU(2) as an example of a group manifold. For the spaces SU(n + 1)/SU(n), Sp(n + 1)/Sp(n), Sp(n)/SU(n), SO(2n)/SU(n) and SO(n + 1)/SO(n − 1), it follows already from the general discussion about Sasaki–Einstein manifolds above that solutions exist. Nevertheless we give some details below, which allow us to say more about the amount of supersymmetry preserved and determine it in some cases, and furthermore to calculate the three-form explicitly. We will not take care to normalize the generators as in the general discussion above, but otherwise use the same conventions. In particular for a root α we denote by eα the one-form dual to the root vector Eα, whereas the dual of a Cartan generator Hj is denoted by hj. As there will always be just one Cartan generator in $\mathfrak m$, we drop the index and denote its dual simply by h. The generic form of the three-form on the Sasaki spaces is then

Equation (6.1)

unless relations of the type α + β = γ exist between three roots in R+S+, in which case we get further contributions from the last line in (4.18).

SU(2) and other group manifolds.

The simplest examples are the ones with K the trivial group. In this case, the existence of invariant spinors is trivial and need not be checked. The manifold on which our fields live is $\mathbb R^p \times G$ with p = rk(G), and the dilaton is a linear function on $\mathbb R^p$. What is further special about this case is that $H^3(G,\mathbb Z) = \mathbb Z$, so in order to satisfy the condition $[H]\in H^3(M,4\pi ^2 \alpha ^{\prime }\mathbb Z)$ we have to adjust the scale of G, whereas the Bianchi identity is trivially satisfied, with dH = 0 and R+ = R = 0. For cosets G/K on the other hand, the third cohomology may well be trivial, and it is the Bianchi identity which fixes the scale.

The models we obtain here are the low-energy description of Wess–Zumino–Witten models and well known. SU(2) is the simplest of the group manifolds, with rank 1. Thus, we have to consider $\mathbb R\times$SU(2). Choose a basis I1, I2, I3 of the Lie algebra $\mathfrak {su}(2)$ satisfying [Ii, Ij] = epsilonijkIk. Then, there is a corresponding basis of left-invariant vector fields on SU(2), which we denote by I1, 2, 3 again, and a dual basis of left-invariant one-forms e1, e2, e3. The metric on $\mathbb R\times$SU(2) is given by

Equation (6.2)

the H-field becomes proportional to the volume form of SU(2):

Equation (6.3)

and the dilaton is

Equation (6.4)

Contrary to the Lévi-Cività connection, both connections ∇+ and ∇ are flat, i.e. R+ = R = 0, and the Bianchi identity dH = 0 holds. So far all of this remains true for higher-dimensional groups as well. In this particular case, one can check that besides the supersymmetry equations also the heterotic equations of motion are satisfied (and the higher order corrections in α' vanish). This is not true for simple groups in general, but should not be expected either for groups of dimension ⩾8. The amount of supersymmetry preserved in $\mathbb R^{5,1}$ is $\mathcal N=1$.

The only condition we did not check so far was the quantization condition on H. In order to calculate the cohomology class of H, we pull it back to S3 along the map

Equation (6.5)

(where $\eta \in [0,\frac{\pi }{2}),\ \xi _1,\xi _2\in [0,2\pi )$) to calculate its integral

Thus, the cohomology class of H is

Equation (6.6)

To satisfy the integrality constraint, we have to rescale the metric as g' = μg, with

Equation (6.7)

which automatically rescales H' = μH (as H is defined in terms of the metric), and leaves invariant the supersymmetry equations. There is thus no continuous volume modulus in the game, as is well known for WZW models. For higher-dimensional simple groups, the analysis is very similar.

SU(n + 1)/SU(n) = S2n + 1.

Here the difference of the ranks is always 1, so that we get a model based on $\mathbb R\times S^{2n+1}$, where the metric on the sphere is different from the round one [9]. A Cartan basis of $\mathfrak {su}(n+1)$ is given by

Equation (6.8)

with positive root vectors Eij for j > i and (Eij)ab = δaiδbj. The corresponding negative root vectors are −Eji. We have

Equation (6.9)

so that the positive roots with respect to λ1, ..., λn are

Equation (6.10)

with 1 ⩽ a < bn + 1 (entries beyond the n components of the root are dropped). The Cartan generators λa are not orthogonal though a better basis is given by

Equation (6.11)

or more explicitly

Equation (6.12)

The roots wrt these generators are

Equation (6.13)

The upper-left corner $\mathfrak {su}(n)$ subalgebra of $\mathfrak {su}(n+1)$ has Cartan generators H1, ..., Hn − 1, whereas Hn generates the orthogonal torus. The roots belonging to $\mathfrak m$ are

Equation (6.14)

Their action on the Cartan generators of $\mathfrak {su}(n)$ is as follows. To every Ha with 1 ⩽ a < n we have one root αa + 1, n + 1 and a roots αj, n + 1 with 1 ⩽ ja such that

Equation (6.15)

Therefore, the relation

Equation (6.16)

holds, and the conditions of section 4 are satisfied. Thus, we can introduce a linear dilaton on $\mathbb R$ and find an invariant spinor (in both S0 and Sn + 1) satisfying all of the supersymmetry conditions, as well as the Bianchi identity. The amount of supersymmetry preserved is $\mathcal N=1$, which follows from the fact that $\mathfrak {su}(n)\subset \mathfrak {spin}(2n)$ has exactly two invariant Majorana spinors (if they exist), and $\mathfrak {so}(\mathfrak m)= \mathfrak {so}(2n+1)$, with $\mathfrak {su}(n)$ embedded in the upper-left $\mathfrak {so}(2n)$ in the standard way. As we have no relations between three roots of $\mathfrak m$, the three-form is simply

Equation (6.17)

with conventions explained at the beginning of the section.

Sp(n + 1)/Sp(n) = S4n + 3.

For the spheres S4n + 3 we have this alternative representation as cosets, again with rank-difference 1. Here $\mathfrak {sp}(n)$ denotes the compact real form of the Lie algebra Cn, not the symplectic Lie algebra which is a non-compact real form. The definition is

Equation (6.18)

where $\mathbb H$ are the quaternions, and † denotes the (quaternion) conjugate transpose. The quaternion conjugate is explicitly given by

A matrix satisfying X† + X = 0 has only imaginary entries on the diagonal and the off-diagonal entries are completely determined by the upper-triangular part, so that the real dimension of $\mathfrak {sp}(n)$ is 3n + 2n(n − 1) = 2n2 + n. A set of Cartan generators is given by the diagonal matrices Ha, a = 1, ..., n, with entry i at the (a, a)-th position, and zeros everywhere else. In particular, the rank is n. The root space decomposition takes the following form. Besides the Cartan generators, we have the basis elements

Equation (6.19)

for a = 1, ..., n, and

Equation (6.20)

i.e. (Ers)ab = δarδbs − δasδbr, etc for 1 ⩽ r < sn. In the following, we will consider the complexification of $\mathfrak {sp}(n)$, and denote the imaginary unit in this space by i, whereas the quaternionic number denoted by i so far will no longer occur explicitly. We have the commutation relations

Equation (6.21)

and choose the positive roots to be

Equation (6.22)

wrt the properly normalized Cartan generator $H^{\prime }_j= \frac{1}{\sqrt{2}}H_j$.

The coset. Again we choose the embedding of $\mathfrak {sp}(n)$ into the upper-left corner of $\mathfrak {sp}(n+1)$. Then, the orthogonal torus is generated by Hn + 1, and the 2n + 1 positive root vectors of $\mathfrak m$ are

Equation (6.23)

with corresponding positive roots

Equation (6.24)

The action of these roots on the Cartan generators Ha (a = 1, ..., n) of $\mathfrak k$ is

Equation (6.25)

from which we read off that

is again satisfied. There are n + 1 invariant spinors, as $\mathfrak {so}(\mathfrak m) =\mathfrak {so}(4n+1)$, with the standard embedding of $\mathfrak {sp}(n)$ into the upper left $\mathfrak {so}(4n)$. They lead to $\mathcal N=\frac{n}{2} +1$ SUSY for even n, and $\mathcal N=\frac{n+1}{2}$ SUSY for odd n. The three-form is somewhat more complicated now, due to the relations

Equation (6.26)

between roots in R+S+. We get

Equation (6.27)

where the constants Nα, β are defined by [Eα, Eβ] = Nα, βEα + β for Eα the root vector corresponding to α.

Sp(n)/SU(n).

We use the notation of the last example. Here the roots γrs, corresponding to matrices with entries $1,i\in \mathbb H$, belong to $\mathfrak {su}(n)$, whereas the other ones with entries j and k belong to $\mathfrak m$. The Cartan basis of $\mathfrak {su}(n)$ is spanned by the elements λi = Hi + 1Hi for i = 1, ..., n − 1, and there is an orthogonal torus generated by ∑iHi. Adding up the roots of $\mathfrak m$, we get

Equation (6.28)

in the basis H1, ..., Hn, and this vanishes on the Cartan algebra of $\mathfrak {su}(n)$, so that our condition is satisfied. The amount of supersymmetry could be larger than $\mathcal N=1$; we have $\dim \mathfrak m= n^2+n+1$, and $\mathcal N$ depends on the number of invariant spinors for the corresponding embedding $\mathfrak {su}(n)\subset \mathfrak {spin}(n^2+n)$. There are no relations involving only three roots of $\mathfrak m$, so that the three-form becomes

Equation (6.29)
SO(2n)/SU(n).

Again the rank difference one is 1. We define $\mathfrak {su}(n)$ as the subalgebra of $\mathfrak {so}(2n)$ leaving the subspace S0S in the spin representation of $\mathfrak {so}(2n)$ invariant. Upon introduction of the canonical complex structure on $\mathbb R^{2n}$ the subalgebra

Equation (6.30)

is determined by the equations

Equation (6.31)

(where Xab = δacXcb) and

Equation (6.32)

for 1 ⩽ a < bn. Denote by Ers (1 ⩽ r < s ⩽ 2n) the standard basis of $\mathfrak {so}(2n)$, with (Ers)ab = δraδsb − δsaδrb. A Cartan basis of $\mathfrak {so}(2n)$ is given by λa = E2a − 1, 2a for a = 1, ..., n or

Equation (6.33)

A basis of $\mathfrak {su}(n)$ is given by the Cartan generators

Equation (6.34)

and combinations of the type (1 ⩽ a < bn)

Equation (6.35)

where all signs are correlated. They satisfy

Equation (6.36)

The Cartan basis H1, ..., Hn − 1 can be extended to one of $\mathfrak {so}(2n)$ by adding $H_n= \sqrt{\frac{2}{n}}\sum _{a=1}^n \lambda _a$, which gives rise to an orthogonal torus again. A basis of $\mathfrak m$ is given by Hn plus

Equation (6.37)

They have the commutation relations

Equation (6.38)

so that the roots βabR+S+, corresponding to B+ab in the (λ1, ..., λn)-basis, are

Equation (6.39)

In the basis H1, ..., Hn − 1, they satisfy the rules as given in table 1.

Table 1. Action of the roots βab (a < b) of $\mathfrak m$ on the Cartan generators of $\mathfrak {su}(n)$. The number in the last column shows how often the corresponding case occurs, being irrelevant in the first case.

a, b∉{2c − 1, 2c}, or a = 2c − 1, b = 2c βab(Hc) = 0  
b = 2c − 1, or a = 2c − 1, b ≠ 2c βab(Hc) = +i n − 2 times
a = 2c, or a ≠ 2c − 1, b = 2c βab(Hc) = −i n − 2 times

We conclude that

Here, $\dim \mathfrak m=n^2-n +1$, and again we might have $\mathcal N>1$ SUSY, with $\mathcal N$ equal to the number of invariant spinors for $\mathfrak {su}(n)\subset \mathfrak {spin}(n^2-n)$. But for n > 3, this is certainly not relevant for heterotic string theory, and for n ⩽ 3 we have the exceptional isomorphisms $\mathfrak {so}(6)=\mathfrak {su}(4)$, $\mathfrak {so}(4) = \mathfrak {su}(2)\oplus \mathfrak {su}(2)$, so that we do not get any essentially new models. The three-form is simply

Equation (6.40)
SO(n + 1)/SO(n − 1).

We use the notation of the previous example. Contrary to $\mathfrak {so}(n)/\mathfrak {so}(n-1),$ this space admits invariant spinors and the common root space decomposition. Suppose first that n is odd, so that we consider $\mathfrak {so}(2k)/\mathfrak {so}(2k-2)$. Denote the common Cartan generators by λ1, ..., λk − 1, and the remaining one in $\mathfrak m$ by λk. The positive roots of $\mathfrak {so}(2k)$ in the λ-basis are

Equation (6.41)

for 1 ⩽ a < bk, and the ones belonging to $\mathfrak m$ are

Equation (6.42)

From

Equation (6.43)

it follows immediately that the condition $\sum _{\alpha \in R^+\setminus S^+ }\alpha \big |_{\mathfrak {so}(2k-2)}=0$ is satisfied. What is the amount of supersymmetry preserved? We have that $\mathfrak m$ is twice the fundamental representation of $\mathfrak {so}(2k-2)$, plus the trivial from the additional Cartan generator. The complex spin representation has the nice feature that for two even-dimensional representations V, W we have

Equation (6.44)

so that ignoring the additional Cartan generator which we have taken into account already, $S(\mathfrak m)$ is simply the tensor product of the spin representation with itself. This contains the trivial representation only twice [34]; upon introduction of a $\mathfrak {spin}(2n)$-invariant inner product and an orthonormal basis {ψi} of S the invariant elements are given by

Equation (6.45)

with Γ the chirality element. From the embedding $\mathfrak {so}(2k-2)\subset \mathfrak {su}(\mathfrak m),$ we knew there would be at least two spinors invariant under the complexified Lie algebra, and as there are no further ones we get indeed $\mathcal N=1$ SUSY. The situation with $\mathfrak {so}(2k+1)/\mathfrak {so}(2k-1)$ is very similar, the only additional positive roots appearing are the

Equation (6.46)

with γk the only new root in R+S+. But this one vanishes on all of the λj for j < k, so we have the condition satisfied here as well. The dimension of SO(n + 1)/SO(n − 1) is 2n − 1, together with the additional $\mathbb R$ factor supporting the linear dilaton this gives dimension 2n, and the only essentially new model with dimension low-enough to be of interest for heterotic string theory is the seven-dimensional SO(5)/SO(3). Below we will consider another embedding of SO(3) into SO(5) which also preserves $\mathcal N=1$. Here, H is

Equation (6.47)

for $\mathfrak {so}(2k)/\mathfrak {so}(2k-2)$, whereas on $\mathfrak {so}(2k+1)/\mathfrak {so}(2k-1)$ it has additionally the term ${\rm e}^{\gamma _{k}}\wedge \,{\rm e}^{-\gamma _{k}}\wedge h$.

Spin(7)/G2 = S7.

For this example, we get a solution as well, but our construction is not applicable; the concrete calculations below mainly serve as a preparation for the next example, G2/SU(3). Here we can use the following argument instead. We have dim$(\mathfrak m)=7$, and $\mathfrak m$ is not a sum of trivial representations of $\mathfrak g_2$. But then it has to be the irreducible seven-dimensional one, as this is the lowest possible dimension for irreducible $\mathfrak g_2$ representations. The spin representation of $\mathfrak g_2$ over $\mathfrak m$ then comes from the standard embedding into $\mathfrak {spin}(7)$, which will be given below, and it is well known that this leaves invariant exactly one Majorana spinor ψ. On the other hand γ(H) commutes with $\mathfrak g_2$ acting on spinors, so that ψ must be an eigenspinor:

Equation (6.48)

For a properly normalized linear function ϕ on $\mathbb R$, we then get γ(dϕ) ∼ id on the one-dimensional spinor space over $\mathbb R$, and from a suitable non-zero 1D spinor κ we can construct an eight-dimensional spinor κ⊗ψ such that

Equation (6.49)

This solution preserves $\mathcal N=1$ SUSY in the orthogonal two-dimensional space, and the same argument applies to other seven-dimensional spaces with nearly parallel G2-structure and one Killing spinor. We will work out one example in more detail below for the space SO(5)/SO(3) with a non-standard embedding. Other examples of this type are given by the Aloff–Wallach spaces N(k, l).

In the remainder of the paragraph, we work out the embedding of $\mathfrak g_2$ into $\mathfrak {spin}(7)$, and its root space decomposition. We define $\mathfrak g_2$ as the subalgebra of $\mathfrak {spin}(7)$ fixing a given Majorana spinor [20]. This condition amounts to seven equations, which we write in terms of the antisymmetric Xab := δacXcb, $X\in \mathfrak {so}(7)$:

Equation (6.50)

Denote by Eij the 7× 7-matrix with (Eij)ab = δaiδbj − δbiδaj. A Cartan basis for $\mathfrak {so}(7)$ is given by

Equation (6.51)

and we have

etc. We can choose a basis of $\mathfrak g_2$ with Cartan generators

Equation (6.52)

and the other basis elements

Equation (6.53)

etc (cf (6.50); the elements listed here are deduced from the first line in (6.50), the second and third line give rise to another eight analogous basis elements, so that we end up with 14 generators). Now we introduce the following complex linear combinations of these:

Equation (6.54)

The positive roots wrt H1, H2 are ${\rm i}/\sqrt{2}$ times:

Equation (6.55)

where the first line has the A-roots, and the second one the B-roots. The ordering of the roots is

In the orthogonal complement of $\mathfrak g_2$ in $\mathfrak {so}(7),$ let us introduce the third Cartan generator

Equation (6.56)

Now it turns out that [H3, A] = 0, but [H3, B] ≠ 0. Thus, the root space decomposition of $\mathfrak g_2$ cannot be lifted to one of $\mathfrak {spin}(7)$, and our construction does not apply to this coset, but the concrete realization of $\mathfrak g_2$ will still be useful for the following example.

G2/SU(3) =S6.

If we put Xa7 = 0 in (6.50), then we obtain the defining equations of $\mathfrak {su}(3)$ inside $\mathfrak {so}(6)$, and thereby realize $\mathfrak {su}(3)$ as a subalgebra of $\mathfrak g_2$. A basis of $\mathfrak {su}(3)$ is given by the Cartan generators H1, H2 and A1, A2, A3 plus their complex conjugates. Therefore, the roots α1, α2, α3 are in S+, whereas β1, 2, 3, are in R+S+. The roots of $\mathfrak g_2$ satisfy the following relations:

Equation (6.57)

contradicting our assumption that roots in R+S+ are higher than those of S+. In this case, this is not a matter of bad choice of positive roots, it is simply not possible to satisfy the condition. However, if we now choose the complex structure on $\mathfrak m$ as follows, by giving its +i eigenspace ((1,0)-vectors) and −i eigenspace ((0,1)-vectors):

Equation (6.58)

then (6.57) shows that we have again $\mathfrak k=\mathfrak {su}(3)\subset \mathfrak u(\mathfrak m)$. J is also compatible with the metric, but leads to a three-form H ∈ Ω(3, 0)⊕Ω(0, 3), instead of the usual H ∈ Ω(2, 1)⊕Ω(1, 2). Now it is not the positive roots in RS corresponding to creation and annihilation operators, but the roots β1, β2 and −β3. Therefore, also the spin representation of the Cartan generators $H_1,H_2\in \mathfrak {su}(3)$ changes slightly (the ζ3 part changes sign) and is now given by

Equation (6.59)

This leaves invariant S0Sm, being equivalent to β1 + β2 − β3 = 0. The three-form is (independently of the complex structure)

Equation (6.60)

It acts on the spinors through

Equation (6.61)

and in particular on $\mathcal S^0\oplus S^3$ as

Equation (6.62)

Here we do not have additional flat directions at our disposal to define the linear dilaton on, as the rank of $\mathfrak k$ is equal to rk$(\mathfrak g_2)$. Now we could try to introduce one by hand and consider $\mathbb R \oplus \mathfrak g_2/\mathfrak {su}(3)$ with ϕ linear on $\mathbb R$. It turns out that this does not lead to any solutions, which is probably due to the fact that the tensor product of the spinor spaces in dimensions 3 and 7 does not give the ten-dimensional spinor space, which in turn follows from dimensional reasons. Let us then try adding $\mathbb R^2$. In this case we can put ϕ(x1, x2) = 2x2, introduce a fourth complex spinor variable $w^4\simeq \frac{1}{2}({\rm d}x^1 +{\rm i}{\rm d}x^2)$ and obtain $\gamma ({\rm d}\phi ) = 2{\rm i}(\overline{\zeta }^4-\zeta ^4)$. Of the four invariant Majorana–Weyl spinors

Equation (6.63)

the first two are annihilated by $\gamma ({\rm d}\phi -\frac{1}{12}H)$, the other two are not. Thus, we get two solutions on $\mathbb R^2 \times G_2/$SU(3). As the dilaton depends only on one of the two additional directions, we should view the other $\mathbb R$ as belonging to the orthogonal space $\mathbb R^{2,1}$, with $\mathcal N=1$ supersymmetry preserved, noting that a three-dimensional Majorana spinor can be combined with η and μ in a unique way into a ten-dimensional Majorana–Weyl spinor.

The almost complex structure on G2/SU(3) we defined here makes it into a so-called nearly Kähler manifold. In six dimensions, only four manifolds of this type are known, and they are all cosets [15]:

The same construction we presented here for G2/SU(3) can be applied to any of them, but for SU(2)3/SU(2)diag it has to be modified slightly, with the two Cartan generators of $\mathfrak m$ combined into one complex vector, effectively treating SU(2)3 as a rank 1 group. Except for G2/SU(3), the nearly Kähler cosets have higher-dimensional generalizations, and one may wonder whether the construction applies to these.

From the well known fact that Sp(n) ×U(1) as a subset of Spin(4n) has no invariant spinors (which is equivalent to the fact that quaternion–Kähler manifolds have no parallel spinors), it can be deduced that the equal rank coset Sp(n)/Sp(n − 1) ×U(1) has no invariant spinors for n > 2. Let us then consider SU(2)n/SU(2)diag, where we take n odd for simplicity. $\mathfrak m$ consists of n − 1 copies of the adjoint representation, with weights n − 1 times (2),(0),(−2), in the Dynkin label notation. From this one can determine the weights of the spin representation, and deduce that the trivial representation occurs with multiplicity

Equation (6.64)

This case thus generalizes, but with a lot of invariant spinors. We leave it open whether or not the dilatino equation can be solved in general, but see the discussion at the end of the following paragraph.

SU(n + 1)/U(1)n

is quite interesting in itself, as the problem of finding invariant spinors becomes a simple combinatorial problem on the roots, so we treat it in detail. Furthermore for odd n, there are no invariant spinors, contrary to all the other non-symmetric examples considered so far. The subalgebra is a Cartan algebra here, and the decomposition of $\mathfrak m\otimes \mathbb C$ into irreducibles is given by the root space decomposition. The positive roots are

Equation (6.65)

with 1 ⩽ a < bn + 1, and terms that run out of the n entries of the root vector are dropped. They satisfy the relations

Equation (6.66)

Similarly, the weights of the spin representation determine the irreducible representations in $S(\mathfrak m)$, and there are invariant spinors if and only if the zero weight occurs. As usual, the weights of $S(\mathfrak m)$ are of the form $\frac{1}{2}\sum _{\alpha \in R^+} \pm \alpha$, with arbitrary combinations of signs.

Proposition 6.1. For even n invariant spinors exist, but not for odd n.

Proof. n even: we have to show that there exists a certain combination of signs such that ∑ab ± αab = 0. Group the roots into pairs of three or four elements according to the following triangle.

Inside each box there is a relation of the type

Equation (6.67)

Fix the signs of the roots αab accordingly and denote them by sign(a, b); then, we certainly have

and the corresponding spinor is invariant. On the other hand, this shows clearly that there will be many more invariant spinors for large n, but we will not try to determine their number.

n odd: the reason that we could sum all the roots to zero by an appropriate choice of signs in the even n case was that in every box the indices that appeared, appeared exactly twice, as in relations (6.66). Adding another row to the diagram of the form

Equation (6.68)

for odd n, we get every index appearing with odd multiplicity. It follows that the roots can no longer add up to zero.

In the case of even n > 2, there are several invariant Majorana spinors and γ(H) acts non-trivially on them, so that it is far from obvious whether this action can be compensated for by a linear dilaton on an extra $\mathbb R$ factor, and for a well-chosen spinor ε. Analyzing our explicit construction of the spinors, this seems rather unlikely, although a proof that it is not possible probably requires more elaborate techniques. The models with n ⩾ 3 are not relevant for heterotic supergravity anyway, but they would very much break the pattern of the other solutions: they are of even dimension greater than 6, so that their cone cannot support parallel spinors [35], contrary to all the solutions we found so far (cf the discussion in the conclusion). We therefore conjecture that it is not possible to solve the dilatino equation on them. The same applies to SU(2)n/SU(2) for odd n, and every other naturally reductive coset of even dimension different from 6.

SO(5)/SO(3).

As an example where the common root space decomposition does not exist, and the subalgebra is not in $\mathfrak u(\mathfrak m)$, we consider the following embedding of $\mathfrak {so}(3)$ into $\mathfrak {so}(5)$, whose coset manifold is known to possess a nearly parallel G2-structure [14, 20]

Equation (6.69)

In terms of our basis vectors

Equation (6.70)

of $\mathfrak {so}(5)$ defined above, the subalgebra and its complement are spanned by

Equation (6.71)

This is not a common root space decomposition, as [H2, D] ≠ 0 and $[H_2, E]\notin \mathfrak m$. It is a naturally reductive coset though. There is no complex structure on $\mathfrak m$ such that $\mathfrak {so}(3)\subset \mathfrak u(\mathfrak m)$, and it will be important to know precisely the action of $\mathfrak k$ on $\mathfrak m$. After a rescaling

it is given by

Equation (6.72)

In particular, the weights of $\mathfrak {so}(3)$ in the Dynkin label notation are (6),(4),(2),(0),(−2),(−4),(−6), which belong to the irreducible seven-dimensional representation with highest weight (6) (in general the highest-weight module to highest weight (n) has dimension n + 1). We write $\mathfrak m = (6)$. For the linear dilaton, we will need to extend this to $\mathfrak m^{\prime }= \mathfrak m \oplus \mathbb R$, which is the $\mathfrak k$-module $\mathfrak m^{\prime }=(6)\oplus (0)$, and the weights of the corresponding spinor module are

Equation (6.73)

showing that $S(\mathfrak m^{\prime }) = 2\times \big ((6)\oplus (0)\big )$. We conclude that there are invariant spinors again, and in order to identify them we need to introduce a complex structure. Take its +i eigenspace to be spanned by

Equation (6.74)

where ∂x is the standard basis vector field of $\mathbb R$. Then, the commutation relations above tell us how the basis elements of $\mathfrak k$ are quantized:

Equation (6.75)

Invariant under all of $\mathfrak {so}(3)$ are

Equation (6.76)

and these multiplied by i, or the Majorana–Weyl spinors

Equation (6.77)

More effort is required to determine the action of γ(H) on these spinors than in the $\mathfrak {su}(n)$ case. We need the commutation relations between elements of $\mathfrak m$. They are as follows:

Equation (6.78)

from which we read off that

Equation (6.79)

leading to the result

Equation (6.80)

Defining $\phi :\mathbb R\rightarrow \mathbb R$ by $\phi (x)=-\frac{7}{2} x$, we obtain $\gamma ({\rm d}\phi ) = -\frac{7}{2}{\rm i}(\overline{\zeta }^H -\zeta ^H)$, and

Equation (6.81)

There is thus exactly one Majorana–Weyl spinor on $\mathbb R\times$SO(5)/SO(3) solving the heterotic supersymmetry equations, leading to $\mathcal N=1$ SUSY in the orthogonal two-dimensional space. More abstractly, we could have argued as in the case Spin(7)/G2, using that SO(5)/SO(3) carries a nearly parallel G2-structure with one Killing spinor [20].

7. Conclusion

We have shown that there is a large class of solutions to the heterotic supergravity BPS-equations (1.1) and the Bianchi identity (1.2) on spaces of the form $\mathbb R^{p,1}\times G/K$, with a non-symmetric coset G/K and homogeneous, nonzero fields H, dϕ, A. The dilaton ϕ has to be taken as a linear function on $\mathbb R$, whereas the other fields live on G/K. In the special case of trivial K, one obtains a WZW model coupled to a linear dilaton, and the quantization condition on H fixes the volume modulus to a discrete set of possible values. For nontrivial K, the scale is completely fixed in terms of α', but this time by the Bianchi identity, which was trivially satisfied in the WZW case. It is not clear however how this relation behaves under inclusion of higher order α' corrections.

Since the string coupling depends on the dilaton as gs∝ eϕ, the two limits ϕ → ± are the weak and strong coupling limit, and for ϕ → the worldsheet loop expansion breaks down. This is discussed in more detail for the near-horizon limit of the NS5-brane, corresponding to G/K = SU(2) = S3 in [3]. For type II strings, the authors propose to use string dualities to gain some information about the strong coupling regime, and it would be interesting to apply a similar analysis to the cases at hand. The weak coupling limit ϕ → − on the other hand describes the dynamics of decoupled NS5 branes.

The gaugino equation γ(F)ε = 0 (together with the condition ∇ε = 0) can be seen as a generalized instanton equation, and is quite interesting in its own right. Here we considered only the simplest solution F = R (3.13), which was shown to solve the Bianchi identity as well in the discussion preceding (3.15). More solutions are known to exist on spaces of the form $\mathbb R^p\times G/H$ however [22, 26, 23, 6], and using these one could try to construct more general supergravity solutions.

The Yang–Mills equations for F- and the H-field equation d*e−2ϕH = 0 are satisfied exactly by our solutions, but the Einstein and dilaton equations (first two equations in (2.2)) receive (α')2-corrections, except for the WZW models. Therefore, it is not clear a priori whether a corresponding CFT really exists. Interestingly, the article [17] by Eguchi, Sugawara and Yamaguchi has a construction of CFTs on exactly the kind of cosets we considered here, coupled to a linear dilaton. Furthermore, they come to the same conclusions about preservation of supersymmetry, namely that one does not get spacetime supersymmetry for SO(n + 1)/SO(n), whereas all the non-symmetric cosets are supersymmetric. But a complete disagreement exists on the level; like in WZW models, they find a discrete set of allowed levels, whereas in our case it seems to be fixed. Additionally it has been argued that the gauged WZW-CFTs used in [17] do not describe sigma models with target space the usual geometric cosets, but that the latter require a modified gauging, which has been successfully applied to quotients with respect to Abelian subgroups only [24].

There is an interesting connection to the cone construction, which was emphasized in [17]. Consider a compact manifold M of Sasaki–Einstein, 3-Sasakian, nearly Kähler (in 6D) or nearly parallel G2 type (in 7D). Then, its cone $\mathbb R_{>0} \times M$, equipped with a particular cone metric, has a parallel spinor and holonomy contained in SU(n), Sp(n), G2 or Spin(7), respectively [5, 12]. Thus, the cone can be used to construct BPS-vacua of string theory (any type, and also of M theory) of the form

Equation (7.1)

with H = dϕ = 0, and A equal to the Lévi-Cività connection for the heterotic case. The amount of supersymmetry preserved is the same as in our models with the radial direction replaced by the linear dilaton. Interest in these models is largely due to the conjectured duality between the worldvolume conformal field theory on branes placed at the tip of the cone and certain string theories on anti-de Sitter spaces, which describe the near-horizon geometry of the brane [28, 1]. For a proposal for the CFT side, see [31].

All of the cosets considered in this paper carry indeed one of the non-integrable structures listed above, although in general with a metric different from the one we used. On SU(n + 1)/SU(n) = S2n + 1, Sp(n)/SU(n), SO(n + 1)/SO(n − 1) we find Sasaki–Einstein structures, whereas Sp(n + 1)/Sp(n) = S4n + 3 even carries a 3-Sasakian structure. The exceptional examples we treated are the nearly Kähler space G2/SU(3) and the nearly parallel G2-manifolds Spin(7)/G2 and SO(5)/SO(3). It seems likely that every spin manifold of these geometric types can serve as a supersymmetric heterotic string background, but I leave this to future work.

There are thus two types of solutions involving the cosets, one on the cone with all fields trivial, and the one with a linear dilaton. A similar situation exists for type II strings and M-theory, where solutions on AdSp × X (X having a Ricci-flat cone) give the near-horizon geometry of branes localized at the tip of the cone over X [1]. In our case, only the Callan–Harvey–Strominger model on SU(2) appears to be a brane limit [16], whereas it has been argued that the general linear dilaton solutions we considered can be obtained as a decoupling limit of string theory on the cone, describing the dynamics of states at the singularity [3, 21].

Acknowledgments

I would like to thank O Lechtenfeld, A D Popov and J M Figueroa-O'Farrill for useful comments and discussions, and the School of Mathematics at the University of Edinburgh for hospitality, where this work was completed. I acknowledge support from the DFG-Graduiertenkolleg 1463. The work was done within the framework of the project supported by the DFG under the grant 436 RUS 113/995.

Please wait… references are loading.
10.1088/1751-8113/45/4/045402