1 Introduction

Sasakian geometry has become an important and active subject, especially after the appearance of the fundamental treatise of Boyer and Galicki [3]. Chapter 7 of this book contains an extended discussion of the topological problems in the theory of Sasakian, and, more generally, K-contact manifolds. These are odd-dimensional analogues to Kähler and symplectic manifolds, respectively.

The precise definition is as follows. Let \((M,\eta )\) be a co-oriented contact manifold with a contact form \(\eta \in \Omega ^1(M)\), that is \(\eta \wedge (d\eta )^n>0\) everywhere, with \(\dim M=2n+1\). We say that \((M,\eta )\) is K-contact if there is an endomorphism \(\Phi \) of TM such that:

  • \(\Phi ^2=-{{\mathrm{Id}}}+ \xi \otimes \eta \), where \(\xi \) is the Reeb vector field of \(\eta \) (that is \(i_\xi \eta =1\), \(i_\xi (d\eta )=0\)),

  • the contact form \(\eta \) is compatible with \(\Phi \) in the sense that \(d\eta (\Phi X,\Phi Y)\,=\,d\eta (X,Y)\), for all vector fields XY,

  • \(d\eta (\Phi X,X)>0\) for all nonzero \(X\in \ker \eta \), and

  • the Reeb field \(\xi \) is Killing with respect to the Riemannian metric defined by the formula \(g(X,Y)\,=\,d\eta (\Phi X,Y)+\eta (X)\eta (Y)\).

In other words, the endomorphism \(\Phi \) defines a complex structure on \({\mathcal {D}}=\ker \eta \) compatible with \(d\eta \), hence \(\Phi \) is orthogonal with respect to the metric \(g|_{\mathcal {D}}\). By definition, the Reeb vector field \(\xi \) is orthogonal to \(\ker \eta \), and it is a Killing vector field.

Let \((M,\eta ,g,\Phi )\) be a K-contact manifold. Consider the contact cone as the Riemannian manifold \(C(M)=(M\times {\mathbb {R}}^{>0},t^2g+dt^2)\). One defines the almost complex structure I on C(M) by:

  • \(I(X)=\Phi (X)\) on \(\ker \eta \),

  • \(I(\xi )=t{\partial \over \partial t},\,I\left( t{\partial \over \partial t}\right) =-\xi \), for the Killing vector field \(\xi \) of \(\eta \).

We say that \((M,\eta ,\Phi ,g,I)\) is Sasakian if I is integrable. Thus, by definition, any Sasakian manifold is K-contact.

There are several topological obstructions to the existence of the aforementioned structures on a compact manifold M of dimension \(2n+1\), for example:

  1. (1)

    the evenness of the pth Betti number for p odd with \(1\, \le \, p \, \le \, n\), of a Sasakian manifold,

  2. (2)

    some torsion obstructions in dimension 5 discovered by Kollár [17],

  3. (3)

    the fundamental group of Sasakian manifolds are special,

  4. (4)

    the cohomology algebra of a Sasakian manifold satisfies the hard Lefschetz property,

  5. (5)

    formality properties of the rational homotopy type.

An early result [13] establishes that the odd Betti numbers up to the middle dimension of Sasakian manifolds must be even. The parity of \(b_1\) was used to produce the first examples of K-contact manifolds with no Sasakian structure [3, example 7.4.16]. More refined tools are needed in the case of even Betti numbers. The cohomology algebra of a Sasakian manifold satisfies a hard Lefschetz property [4]. Using it examples of K-contact non-Sasakian manifolds are produced in [5] in dimensions 5 and 7. These examples are nilmanifolds with even Betti numbers, so in particular they are not simply connected.

The fundamental group can also be used to construct K-contact non-Sasakian manifolds. Fundamental groups of Sasakian manifolds are called Sasaki groups, and satisfy strong restrictions. Using this it is possible to construct (non-simply connected) compact manifolds which are K-contact but not Sasakian [8].

When one moves to the case of simply connected manifolds, K-contact non-Sasakian examples of any dimension \(\ge \)9 were constructed in [16] using the evenness of the third Betti number of a compact Sasakian manifold. Alternatively, using the hard Lefschetz property for Sasakian manifolds there are examples [19] of simply connected K-contact non-Sasakian manifolds of any dimension \(\ge \)9.

In [24] and in [2] the rational homotopy type of Sasakian manifolds is studied. In [2] it is proved that all higher order Massey products for simply connected Sasakian manifolds vanish, although there are Sasakian manifolds with non-vanishing triple Massey products. This yields examples of simply connected K-contact non-Sasakian manifolds in dimensions \(\ge \)17. However, Massey products are not suitable for the analysis of lower dimensional manifolds.

Hence, the problem of the existence of simply connected K-contact non-Sasakian compact manifolds (open problem 7.4.1 in [3]) is still open in dimensions 5 and 7. Dimension 5 is the most difficult one, and it is treated in [3] separately. Here one has to use the obstructions of [17] which are very subtle torsion obstructions associated to the classification of Kähler surfaces. By definition, a simply connected compact oriented 5-manifold is called a Smale–Barden manifold. These manifolds are classified topologically by \(H_2(M,{\mathbb {Z}})\) and the second Stiefel–Whitney class. Chapter 10 of the book by Boyer and Galicki is devoted to a description of some Smale–Barden manifolds which carry Sasakian structures. The following problem is still open (open problem 10.2.1 in [3]).

Do there exist Smale–Barden manifolds which carry K-contact but do not carry Sasakian structures?

In this note we solve the described problem in the easier case of dimension 7 (the solution is still possible by means of homotopy theory combined with symplectic surgery).

Theorem 1

There exist 7-dimensional compact simply connected K-contact manifolds which do not admit a Sasakian structure.

We then turn around to the study of the rational homotopy type of K-contact and Sasakian simply connected manifolds of dimension 7. In particular, we prove:

Corollary 2

Let M be a simply connected compact K-contact 7-dimensional manifold. Suppose that the cup product map \(H^2(M)\times H^2(M)\longrightarrow H^4(M)\) is non-zero. Then M does not admit a Sasakian structure.

Formality is a very useful rational homotopy property that has been widely used to distinguish between symplectic and Kähler manifolds [21] (see Sect. 6 for definitions and details). Simply connected compact manifolds of dimension \(\le \)6 are always formal, so formality becomes interesting in dimension 7. We study this property in detail giving a precise characterisation for Sasakian manifolds (see Theorem 15). In particular, we have the following:

Corollary 3

Let M be a simply connected compact Sasakian 7-dimensional manifold. Then M is formal if and only if all triple Massey products are zero.

2 Gompf–Cavalcanti manifold

Let \((M,\omega )\) be a symplectic manifold of dimension 2n. For every \(0\le k \le n\), we define the Lefschetz map as \(L_\omega :H^{n-k}(M) \rightarrow H^{n+k}(M)\), \(L_\omega ([\beta ])=[\beta \wedge \omega ^{n-k}]\). We say that M satisfies the hard Lefschetz property if \(L_\omega \) is an isomorphism for every \(0 \le k \le n\).

Proposition 4

There exists a simply connected 6-dimensional symplectic manifold \((M,\omega )\) such that \(\dim \ker \,(L_\omega : H^2(M)\rightarrow H^4(M))\) is odd.

Proof

Gompf constructs in [14, Theorem 7.1] an example of a simply connected 6-dimensional symplectic manifold \((M,\omega )\) which does not satisfy the hard Lefschetz property, that is, the Lefschetz map \(L_\omega : H^2(M)\rightarrow H^4(M)\) is not an isomorphism. If \(\dim \ker L_\omega \) is already odd then we have finished.

So let us suppose that \(\dim \ker L_\omega \) is even. Take a cohomology class \(a\in H^2(M)\) which belongs to the kernel of \(L_\omega \). In [7, Lemma 2.4] Cavalcanti proves that given a symplectic manifold \((M,\omega )\) as above satisfying that there exists a symplectic surface \(S\hookrightarrow M\) with \(\langle a,[S]\rangle \ne 0\), then there is another 6-dimensional symplectic manifold \((M',\omega ')\) (the symplectic blow-up of M along S) satisfying

$$\begin{aligned} \dim \ker \left( L_{\omega '}: H^2\left( M'\right) \rightarrow H^2\left( M'\right) \right) = \dim \ker \left( L_\omega : H^2\left( M\right) \rightarrow H^2\left( M\right) \right) -1. \end{aligned}$$

The symplectic blow-up of M along S is constructed in [20], where it is proved that the fundamental groups \(\pi _1(M')\cong \pi _1(M)\), hence \(M'\) is simply connected. This means that the simply connected 6-dimensional symplectic manifold \(M'\) satisfies that \(\dim \ker (L_{\omega '} : H^2(M')\rightarrow H^4(M'))\) is odd, as required.

It remains to find \(S\hookrightarrow M\) as required. The cohomology class a is non-zero, so there is some \(b\in H^4(M,{\mathbb {Z}})\) such that \(a\cup b\ne 0\). It is easy to see that there is a rank 2 complex vector bundle \(E\rightarrow M\) with \(c_1(E)=0,\,c_2(E)=2b\). This corresponds to the fact that the map \([M,B{\text {SU}}(2)]\rightarrow H^4(M,{\mathbb {Z}})\) given by the second Chern class exhausts \(2\, H^4(M,{\mathbb {Z}})\). A short proof runs as follows: \(B{\text {SU}}(2)\) has trivial 3-skeleton and it has \(\pi _4(B{\text {SU}}(2))={\mathbb {Z}}\) and \(\pi _5(B{\text {SU}}(2))={\mathbb {Z}}_2\). Represent the cohomology class b by a cocycle \(\varphi _b:C_4(M)\rightarrow {\mathbb {Z}}\), where \(C_4(M)\) is the space of cellular chains. Given b, we define \(f:M\rightarrow B{\text {SU}}(2)\) inductively on the skeleta (in what follows we denote by X[k] the k-skeleton of a space X). It is trivial on the 3-skeleton of M. For every 4-cell c, we define \(f:c\rightarrow B{\text {SU}}(2)[4]=S^4\) to have degree \(\varphi _b(c)\in {\mathbb {Z}}\). As M is simply connected there are no 5-cells, so it only remains to attach the 6-cell \(c_6\) to the 4-skeleton M[4]. The attaching map is given by some \(g:S^5\rightarrow M[4]\). When composed with f, we have a map \(f\circ g: S^5 \rightarrow B{\text {SU}}(2)\), which gives an obstruction element \(o_f\in \pi _5(B{\text {SU}}(2))={\mathbb {Z}}_2\). If we multiply b by two, then the map \(\varphi _b\) gets multiplied by 2. The corresponding f is given by composing f with a double cover of \(S^4\), hence the obstruction element is \(2o_f=0\). This means that the map f associated to 2b can be extended to \(M\rightarrow B{\text {SU}}(2)\).

Now take the rank 2 bundle \(E \rightarrow M\) just constructed. Assume that \([\omega ]\) is a an integral cohomology class (which can always be done by perturbing \(\omega \) slightly to make it rational and multiplying it by a large integer). Let \(L\rightarrow M\) be the line bundle with first Chern class \(c_1(L)=[\omega ]\). We now use the asymptotically holomorphic techniques introduced by Donaldson [10]. Specifically, the result of [1] guarantees the existence of a suitable large \(k\gg 0\) and a section of \(E\otimes L^{\otimes k}\) whose zero locus is a symplectic manifold (an asymptotically holomorphic manifold in fact). This zero locus \(S\subset M\) is a symplectic surface, and the cohomology class defined by S is \(c_2(E\otimes L^{\otimes k})= c_2(E)+2k c_1(L)=2b+2k[\omega ]\). Therefore \(\langle a,[S]\rangle =\langle a, 2b+2k[\omega ] \rangle =2\langle a,b\rangle \ne 0\), as required. \(\square \)

We will call the manifold produced in Proposition 4 the Gompf–Cavalcanti manifold, because it is constructed by the surgery technique of Gompf [14] together with the symplectic blow-up of Cavalcanti [7]. Note however that this is not a unique one but a family of manifolds.

3 Simply-connected K-contact non-Sasakian manifolds in dimension 7

We show the existence of simply connected compact K-contact non-Sasakian manifolds in dimension 7 by proving that the Boothby–Wang fibration over the Gompf–Cavalcanti manifold is K-contact but non-Sasakian. The existence of a K-contact structure on such fibration is shown in [2] and [16]. For the convenience of the reader we briefly recall these constructions.

Let \((B,\omega )\) be a symplectic manifold such that the cohomology class \([\omega ]\) is integral. Consider the principal \(S^1\)-bundle \(\pi : M\rightarrow B\) given by the cohomology class \([\omega ]\,\in \, H^2(B,\,{\mathbb {Z}})\). Fibrations of this kind were first considered by Boothby and Wang and are called Boothby–Wang fibrations. By [25], the total space M carries an \(S^1\)-invariant contact form \(\eta \) such that \(\eta \) is a connection form whose curvature is \(d\eta =\pi ^*\omega \). We have the following result (which is known, compare Theorem 6.1.26 and Proposition 7.1.2 in [3]).

Theorem 5

Any Boothby–Wang fibration admits a K-contact structure on the total space.

Proof

To prove this theorem we need to introduce a certain tool, called the universal contact moment map in the sense of Lerman [18]. Recall that by our assumption the given contact distribution \({\mathcal {D}}\) is determined by the contact form \(\eta \), that is \({\mathcal {D}}=\ker \eta \). Consider its annihilator \({\mathcal {D}}^0\subset T^*M\). Clearly, \({\mathcal {D}}^0\) is a line bundle, and, therefore, it has two components after the removal of the zero section,

$$\begin{aligned} {\mathcal {D}}^0{\setminus } M={\mathcal {D}}^0_+\sqcup {\mathcal {D}}^0_{-}. \end{aligned}$$

Single out one of these components, say \({\mathcal {D}}^0_{+}\). Consider the Lie algebra of contact vector fields \(\chi (M,\eta )\) on M. It is known that this Lie algebra can be identified with a space of sections of the vector bundle \(TM/{\mathcal {D}}\), that is \(\chi (M,\eta )\cong \Gamma (M,TM/{\mathcal {D}})\). Because of that there is a natural pairing between points of the line bundle \({\mathcal {D}}^0\) and contact vector fields given by the formula

$$\begin{aligned} {\mathcal {D}}^0\times \chi (M,\eta )\rightarrow {\mathbb {R}},\quad ((p,\beta ),X)\mapsto \langle \beta ,X_p\rangle \end{aligned}$$

where \(\beta \in {\mathcal {D}}^0,X_p\in T_pM, p\in M\). Suppose that a Lie algebra \({\mathfrak {g}}\) acts on M by contact vector fields, that is, there exists a representation \(\rho :{\mathfrak {g}}\rightarrow \chi (M,\eta )\). Define the universal moment map as the map

$$\begin{aligned} \psi :{\mathcal {D}}_{+}^0\rightarrow {\mathfrak {g}}^* \end{aligned}$$

by the formula

$$\begin{aligned} \langle \psi (p,\beta ),X\rangle =\langle (p,\beta ),\rho (X)\rangle =\langle \beta ,\rho (X)_p\rangle , \end{aligned}$$

where \((p,\beta )\in ({\mathcal {D}}_{+}^0)_p\subset T^*_pM,X\in {\mathfrak {g}}\). Now the proof becomes a consequence of the following criterion proved by Lerman [18].

Proposition 6

A compact co-orientable contact manifold \((M,\eta )\) admits a K-contact metric g if and only if there exists an action of a torus T on M preserving the contact structure \({\mathcal {D}}\) and a vector \(X\in {\mathfrak {t}}=L(T)\) so that the function \(\langle \psi ,X\rangle :{\mathcal {D}}^0_{+}\rightarrow {\mathbb {R}}\) is strictly positive. \(\square \)

We continue with the proof of Theorem 5. Consider the \(S^1\)-action on M given by the Reeb vector field. Let \({\mathfrak {g}}=L(S^1)\), and \(\rho :{\mathfrak {g}}\rightarrow \chi (M,\eta )\) be the homomorphism of Lie algebras determined by this action (thus, \({\mathfrak {g}}={\mathfrak {t}}=L(S^1)\) in this particular situation). Since the \(S^1\)-action is free, \(\rho (X)_p\not =0\) for any \(p\in M\). Now,

$$\begin{aligned} \langle \psi ,X\rangle (p,\beta )=\langle \psi (p,\beta ),X\rangle =\langle \beta ,\rho (X)_p\rangle . \end{aligned}$$

Note that in the considered case \(\beta \in ({\mathcal {D}}^0_{+})_p\subset T_p^*M\), and, therefore, \(\beta \not =0\). Also \((p,\beta )\) belongs to the annihilator of the distribution \({\mathcal {D}}\), while \(\rho (X)\) is transversal to \({\mathcal {D}}\), since it is given by the Reeb vector field. Thus, for any point p, \(\langle (p,\beta ),\rho (X)_p\rangle \not =0\). Hence, X may be chosen to yield positive sign everywhere, and we complete the proof by applying Proposition 6. \(\square \)

Remark 7

Proposition 7.1.2 from [3] is due to Rukimbira. In this work we give a different proof based on Lerman’s criterion given by Proposition 6.

The following gives a proof of Theorem 1.

Theorem 8

The total space of the Boothby–Wang fibration over the Gompf–Cavalcanti manifold is a simply connected K-contact non-Sasakian manifold of dimension 7.

Proof

Let \((M,\omega )\) be a Gompf–Cavalcanti manifold as given by Proposition 4. We can assume that \([\omega ]\) is an integral cohomology class. Let

$$\begin{aligned} S^1\rightarrow E\rightarrow M \end{aligned}$$
(1)

be the associated Boothby–Wang fibration. By Theorem 5, E has a K-contact structure. Now we need to prove that E cannot carry Sasakian structures.

There is an exact sequence

$$\begin{aligned} H_2(M)\rightarrow H_1\left( S^1\right) ={{\mathbb {Z}}}\rightarrow H_1(E) \rightarrow 0 \end{aligned}$$

from the Serre spectral sequence. The map \(H_2(M)\rightarrow {{\mathbb {Z}}}\) is cupping with \([\omega ]\in H^2(M)\). Taking \([\omega ]\) integral cohomology class and primitive, we have that \(H_2(M)\rightarrow {{\mathbb {Z}}}\) is surjective and hence \(H_1(E)=0\). The long homotopy exact sequence gives \(\pi _1(S^1)={{\mathbb {Z}}} \rightarrow \pi _1(E) \rightarrow \pi _1(M)=0\), hence \(\pi _1(E)\) is abelian. Therefore E is simply connected.

The Gysin exact sequence associated to (1) is

$$\begin{aligned} H^1(M)=0 \mathop {\longrightarrow }\limits ^{\wedge \omega } H^3(M) \longrightarrow H^3(E) \longrightarrow H^2(M) \mathop {\longrightarrow }\limits ^{\wedge \omega } H^4(M). \end{aligned}$$

Thus

$$\begin{aligned} b^3(E)=b^3(M)+ \dim \left( \ker L_\omega :H^2(M) \rightarrow H^4(M)\right) . \end{aligned}$$

As M is a 6-manifold, we have that \(b^3(M)\) is even (by Poincaré duality, the intersection pairing on \(H^3(M)\) is an antisymmetric non-degenerate bilinear form, hence the dimension of \(H^3(M)\) is even). By construction, \(\dim (\ker L_\omega :H^2(M) \rightarrow H^4(M))\) is odd, so \(b^3(E)\) is odd. As the third Betti number of a 7-dimensional Sasakian manifold has to be even [13], we have that E cannot admit a Sasakian structure. \(\square \)

4 Regularity and quasi-regularity

A Sasakian or a K-contact structure on a compact manifold M is called quasi-regular if there is a positive integer \(\delta \) satisfying the condition that each point of M has a foliated coordinate chart (Ut) with respect to \(\xi \) (the coordinate t is in the direction of \(\xi \)) such that each leaf for \(\xi \) passes through U at most \(\delta \) times. If \(\delta \,=\, 1\), then the Sasakian or K-contact structure is called regular (see [3, p. 188]).

If N is a Kähler manifold whose Kähler form \(\omega \) defines an integral cohomology class, then the total space of the circle bundle \(S^1 \hookrightarrow M \mathop {\longrightarrow }\limits ^{\pi } N\) with Euler class \([\omega ]\in H^2(M,{\mathbb {Z}})\) is a regular Sasakian manifold with contact form \(\eta \) such that \(d \eta \,=\, \pi ^*(\omega )\). The converse also holds: if M is a regular Sasakian structure then the space of leaves N is a Kähler manifold, and we have a circle bundle \(S^1\rightarrow M \rightarrow N\) as above. If M has a quasi-regular Sasakian structure, then the space of leaves N is a Kähler orbifold with cyclic quotient singularities, and there is an orbifold circle bundle \(S^1 \rightarrow M\rightarrow N\) such that the contact form \(\eta \) satisfies \(d\eta =\pi ^*(\omega )\), where \(\omega \) is the orbifold Kähler form.

Similar properties hold in the K-contact case, substituting Kähler by symplectic (actually almost Kähler). If M has a regular K-contact structure, then it is the total space of a circle bundle \(S^1 \hookrightarrow M \mathop {\longrightarrow }\limits ^{\pi } N\), where \((N,\omega )\) is a symplectic manifold, with Euler class \([\omega ]\in H^2(M,{\mathbb {Z}})\) and \(d \eta \,=\, \pi ^*(\omega )\). If M has a quasi-regular K-contact structure, then it is the total space of an orbifold circle bundle \(S^1 \rightarrow M\rightarrow N\) over a symplectic orbifold N with cyclic quotient singularities and Euler class \([\omega ]\in H^2(M,{\mathbb {Z}})\), where \(\omega \) is the orbifold symplectic form.

A result of [22] says that if M admits a Sasakian structure, then it admits also a quasi-regular Sasakian structure. This also extends to the case of K-contact structures (Rukimbira [23], see also Theorem 7.1.10 in [3]).

Proposition 9

If a compact manifold M admits a K-contact structure, it admits a quasi-regular contact structure.

Proof

Assume that there is a K-contact structure on M. By Proposition 6, there exists a torus action \(T\times M\rightarrow M\) preserving the contact distribution and a vector \(X\in {\mathfrak {t}}\) such that \(\langle \psi ,X\rangle >0\). Choose a vector \(Y\in {\mathfrak {t}}\) with the property that it is tangent to an embedding \(T'=S^1\hookrightarrow T\). Clearly, the corresponding fundamental vector field \(Y_M\) has the property that the leaves of the corresponding foliations are compact. The set of such Y is dense in \({\mathfrak {t}}\). Therefore, for vectors Y which are sufficiently close to X, the condition \(\langle \psi , Y\rangle >0\) is still satisfied.

So it remains to see that there is K-contact structure whose Reeb vector field is \(Y_M\), since this will be quasi-regular because the leaves of the characteristic foliation are all compact. We follow the notations of the proof of Theorem 5. The action of the circle \(T'\) on M preserves \({\mathcal {D}}\), hence the lifted action of \(T'\) on \(T^*M\) preserves \({\mathcal {D}}^0\). Since \(T'\) is connected, the lifted action preserves the connected component \({\mathcal {D}}^0_+\) as well. It follows that for any 1-form \(\beta \) on M with \(\ker \beta ={\mathcal {D}}\), the average \(\bar{\beta }\) of \(\beta \) over \(T'\) still satisfies \(\ker \,\bar{\beta }={\mathcal {D}}\). So \(\bar{\beta }\in {\mathcal {D}}^0\). Now use the formula [derived in [18], formulae (3.4) and (3.5)],

$$\begin{aligned} i_{Y_M}\bar{\beta }=\left\langle \psi \circ \bar{\beta },Y\right\rangle >0. \end{aligned}$$

Now let

$$\begin{aligned} \eta =\left( \langle \psi \circ \bar{\beta },Y\rangle \right) ^{-1}\bar{\beta }, \end{aligned}$$

which satisfies \(i_{Y_M}\eta =1\). Hence \(\eta \) defines the contact structure and \(Y_M\) is its Killing vector field. Then \(TM={\mathcal {D}}\oplus \langle Y_M\rangle \), and the splitting is \(T'\)-invariant. We use the splitting to define the desired Riemannian metric g. Declare \({\mathcal {D}}\) and \(\langle Y_M\rangle \) to be orthogonal and define \(g(Y_M,Y_M)=1\), thus \(Y_M\) becomes a unit normal to \({\mathcal {D}}\). On \({\mathcal {D}}\) we choose a \(T'\)-invariant complex structure compatible with \(d\eta |_{{\mathcal {D}}}\) and define \(g|_{{\mathcal {D}}}(\cdot ,\cdot )=d\eta |_{{\mathcal {D}}}(\cdot ,\Phi \cdot )\). Then g is \(T'\)-invariant and hence \(L_{Y_M}g=0\). Thus we have obtained a K-contact structure on M. \(\square \)

5 Minimal models and formality

Now we want to analyse the rational homotopy type of K-contact and Sasakian simply connected 7-manifolds, in particular the property of formality. Simply connected compact manifolds of dimension \(\le \)6 are always formal [12], so dimension 7 is the first instance in which formality is an issue.

We start by reviewing concepts about minimal models and formality from [11, 12, 15]. A differential graded algebra (or DGA) over the real numbers \({\mathbb {R}}\), is a pair (Ad) consisting of a graded commutative algebra \(A=\oplus _{k\ge 0} A^k\) over \({\mathbb {R}}\), and a differential d satisfying the Leibnitz rule \(d(a\cdot b) = (da)\cdot b +(-1)^{|a|} a\cdot (db)\), where |a| is the degree of a. Given a differential graded commutative algebra \(({A},\,d)\), we denote its cohomology by \(H^*({A})\). The cohomology of a differential graded algebra \(H^*({A})\) is naturally a DGA with the product inherited from that on A and with the differential being identically zero. The DGA \(({A},\,d)\) is connected if \(H^0({A})\,=\,{\mathbb {R}}\), and A is 1-connected if, in addition, \(H^1({A})\,=\,0\). Henceforth we shall assume that all our DGAs are connected. In our context, the main example of DGA is the de Rham complex \((\Omega ^*(M),\,d)\) of a connected differentiable manifold M, where d is the exterior differential.

Morphisms between DGAs are required to preserve the degree and to commute with the differential. A morphism \(f:(A,d)\rightarrow (B,d)\) is a quasi-isomorphism if the map induced in cohomology \(f^*:H^*(A,d)\rightarrow H^*(B,d)\) is an isomorphism. Quasi-isomorphism produces an equivalence relation in the category of DGAs.

A DGA \((\mathcal {M},\,d)\) is minimal if

  1. (1)

    \(\mathcal {M}\) is free as an algebra, that is, \(\mathcal {M}\) is the free algebra \(\bigwedge V\) over a graded vector space \(V\,=\,\bigoplus _i V^i\), and

  2. (2)

    there is a collection of generators \(\{x_\tau \}_{\tau \in I}\) indexed by some well ordered set I, such that \(|x_\mu |\,\le \, |x_\tau |\) if \(\mu \,< \,\tau \) and each \(d x_\tau \) is expressed in terms of preceding \(x_\mu \), \(\mu \,<\,\tau \).

We say that \((\bigwedge V,\,d)\) is a minimal model of the differential graded commutative algebra \(({A},\,d)\) if \((\bigwedge V,\,d)\) is minimal and there exists a quasi-isomorphism \(\rho :{(\bigwedge V,\,d)}\longrightarrow {({A},\,d)}\). A connected DGA \(({A},\,d)\) has a minimal model unique up to isomorphism. For 1-connected DGAs, this is proved in [9]. In this case, the minimal model satisfies that \(V^1=0\) and the condition (2) above is equivalent to \(dx_\tau \) not having a linear part.

A minimal model of a connected differentiable manifold M is a minimal model \((\bigwedge V,\,d)\) for the de Rham complex \((\Omega ^*(M),\,d)\) of differential forms on M. If M is a simply connected manifold, then the dual of the real homotopy vector space \(\pi _i(M)\otimes {\mathbb {R}}\) is isomorphic to \(V^i\) for any i (see [9]).

A model of a DGA (Ad) is any DGA (Bd) with the same minimal model (that is, they are equivalent with respect to the equivalence relation determined by the quasi-isomorphisms).

A minimal algebra \((\bigwedge V,\,d)\) is called formal if there exists a morphism of differential algebras \(\psi :{(\bigwedge V,\,d)}\,\longrightarrow \, (H^*(\bigwedge V),0)\) inducing the identity map on cohomology. Also a differentiable manifold M is called formal if its minimal model is formal. The formality of a minimal algebra is characterized as follows.

Proposition 10

[9] A minimal algebra \((\bigwedge V,\,d)\) is formal if and only if the space V can be decomposed into a direct sum \(V\,=\, C\oplus N\) with \(d(C) \,=\, 0\) and d injective on N, such that every closed element in the ideal I(N) in \(\bigwedge V\) generated by N is exact.

This characterization of formality can be weakened using the concept of s-formality introduced in [12].

Definition 11

A minimal algebra \(\left( \bigwedge V,\,d\right) \) is s-formal (\(s> 0\)) if for each \(i\le s\) the space \(V^i\) of generators of degree i decomposes as a direct sum \(V^i=C^i\oplus N^i\), where the spaces \(C^i\) and \(N^i\) satisfy the three following conditions:

  1. (1)

    \(d(C^i) = 0\),

  2. (2)

    the differential map \(d:N^i\longrightarrow \bigwedge V\) is injective, and

  3. (3)

    any closed element in the ideal \(I_s=I\left( \bigoplus \nolimits _{i\le s} N^i\right) \), generated by the space \(\bigoplus \nolimits _{i\le s} N^i\) in the free algebra \(\bigwedge \left( \bigoplus \nolimits _{i\le s} V^i\right) \), is exact in \(\bigwedge V\).

A differentiable manifold M is s-formal if its minimal model is s-formal. Clearly, if M is formal then M is s-formal, for any \(s>0\). The main result of [12] shows that sometimes the weaker condition of s-formality implies formality.

Theorem 12

[12] Let M be a connected and orientable compact manifold of dimension 2n or \((2n-1)\). Then M is formal if and only if it is \((n-1)\)-formal.

By Corollary 3.3 in [12] a simply connected compact manifold is always 2-formal. Therefore, Theorem 12 implies that any simply connected compact manifold of dimension not more than six is formal. For simply connected 7-dimensional compact manifolds, we have that M is formal if and only if M is 3-formal.

Theorem 12 also holds for compact connected orientable orbifolds, since the proof of [12] only uses that the cohomology \(H^*(M)\) is a Poincaré duality algebra.

6 Homotopy properties of simply connected Sasakian 7-manifolds

Proposition 13

Let M be a simply connected compact K-contact 7-dimensional manifold. Then a model for M is \((H\otimes \bigwedge (x),d)\), where H is the cohomology algebra of a simply connected symplectic 6-dimensional orbifold and \(dx=\omega \in H^2\) is the class of the symplectic form.

If M is Sasakian, then H is the cohomology algebra of a simply connected 6-dimensional Kähler orbifold.

Proof

Suppose M admits a Sasakian structure. Then M admits a quasi-regular Sasakian structure [22]. Therefore, there is an orbifold circle bundle \(S^1 \rightarrow M \rightarrow B\), where B is a compact Kähler orbifold of dimension 6, with Euler class given by the Kähler form \(\omega \in H^2(B)\). We note that B is simply connected because M is so (see [3, Theorem 4.3.18]). In particular, \(S^1 \rightarrow M \rightarrow B\) is a rational fibration, hence if \({\mathcal {M}}\) is a model for B, then \({\mathcal {M}}\otimes \bigwedge (x)\), with \(|x|\,=\,1\), \(dx\,=\,\omega \), is a model for M.

Now B is a simply connected compact orbifold of dimension 6. So it is 2-formal. Theorem 12 also holds for orbifolds, hence B is formal. Therefore \({\mathcal {M}}\sim (H,0)\), where \(H=H^*(B)\) is the cohomology algebra of B. So a model for M is of the form \((H\otimes \bigwedge (x),d)\), \(dx=\omega \in H^2\).

The case where M admits a K-contact structure is similar. By Proposition 9, it admits a quasi-regular K-contact structure. Therefore, M is an orbifold \(S^1\)-bundle over a symplectic orbifold \(S^1 \rightarrow M \rightarrow B\), with Euler class given by the orbifold symplectic form \(\omega \in H^2(B)\). As above, a model for M is \((H\otimes \bigwedge (x),d)\), \(dx=\omega \in H^2\), where \(H=H^*(B)\). \(\square \)

We prove now Corollary 2.

Corollary 14

Let M be a simply connected compact K-contact 7-dimensional manifold. Suppose that the cup product map \(H^2(M)\times H^2(M)\longrightarrow H^4(M)\) is non-zero. Then M does not admit a Sasakian structure.

Proof

Let us compute the cohomology of M from its model \(({\mathcal {M}},d)=(H\otimes \bigwedge (x),d)\), \(dx=\omega \), where \(H=H^*(B)\) is the cohomology algebra of a 6-dimensional simply connected symplectic manifold. Note that \(\omega \in H^2\) is a non-zero element with \(\omega ^3\in H^6\) generating the top cohomology.

Consider the Lefschetz map \(L_\omega :H^*\rightarrow H^{*+2}\), and let \(K^*=\ker L_\omega \), \(Q^*={{\mathrm{coker}}}L_\omega \). We have a (non-canonical) isomorphism \(H^i(M) \cong Q^i \oplus K^{i-1} x\). Note that \(Q^3=K^3=H^3\) and \(H^6={\mathbb {R}}\). Also \(Q^2=H^2/\langle \omega \rangle \), and \(K^4=\ker (L_\omega : H^4\rightarrow {\mathbb {R}})\) are vector spaces of codimension one. We have the following:

$$\begin{aligned} H^0(M)&={\mathbb {R}}, \\ H^1(M)&=0, \\ H^2(M)&= Q^2, \\ H^3(M)&=H^3 \oplus K^2 x, \\ H^4(M)&=Q^4 \oplus H^3 x, \\ H^5(M)&=K^4 x, \\ H^6(M)&=0, \\ H^7(M)&=\langle \omega ^3 x\rangle . \end{aligned}$$

The map \(H^2(M)\times H^2(M)\rightarrow H^4(M)\) factors through \(Q^2 \times Q^2 \rightarrow Q^4\). Hence if it is non-zero then \(Q^4\ne 0\). In particular, the Lefschetz map \(L_\omega :H^2\rightarrow H^{4}\) is not an isomorphism, so B is not hard Lefschetz.

If M admits a Sasakian structure, then there is a quasi-regular fibration \(S^1 \rightarrow M \rightarrow B\) with B satisfying the hard Lefschetz property (it is a Kähler orbifold, so [26] is applicable). This contradicts the above. \(\square \)

Now we shall study the case of Sasakian 7-manifolds in more detail. Let M be a simply connected compact Sasakian 7-dimensional manifold. Then

$$\begin{aligned} {\mathcal {M}}=\left( H\otimes \bigwedge (x),d\right) \end{aligned}$$

is a model for M, by Proposition 13, where \(H=H^*(B)\) is the cohomology algebra of a simply connected compact 6-dimensional Kähler orbifold. This algebra H has a very rich structure:

  1. (1)

    there is a canonical isomorphism \(H^6 \cong {\mathbb {R}}\), which is given by integration \(\int _M:H^6\rightarrow {\mathbb {R}}\);

  2. (2)

    H is a Poincaré duality algebra, hence \(H^3 \otimes H^3 \rightarrow {\mathbb {R}}\) is an antisymmetric bilinear pairing;

  3. (3)

    there is a scalar product on each \(H^j\). This is given by the Hodge star operator \(* :H^j \rightarrow H^{6-j}\) combined with wedge and integration;

  4. (4)

    H has a Hodge structure, that is, \(H\otimes {\mathbb {C}}\) has a bigrading such that \(H^k\otimes {\mathbb {C}}=\oplus _{p+q=k} H^{p,q}\), where \(H^{p,q}=\overline{H^{q,p}}\), and the wedge product respects the bigrading;

  5. (5)

    there is a distinguished element \(\omega \in H^2\) which is in \(H^{1,1}\). This defines the space of primitive forms \(P=\langle \omega \rangle ^\perp \subset H^2\). Hence \(H^2=\langle \omega \rangle \oplus P\). Moreover \(P=P^{1,1}\oplus P^{2,0}\), where \(P^{1,1}=P\cap H^{1,1}\) and \(P^{2,0}=P \cap (H^{2,0}\oplus H^{0,2})\);

  6. (6)

    the Lefschetz map \(L_\omega :H^2\rightarrow H^4\) is an isomorphism. Therefore \(H^4=\langle \omega ^2\rangle \oplus \omega P^{1,1}\oplus \omega P^{2,0}\). By Theorem 3.16 of Chapter V of [27], for \(\alpha _1\in P^{1,1}\) we have \(* \alpha _1=-\alpha _1\wedge \omega \), for \(\alpha _2\in P^{2,0}\) we have \(* \alpha _2=\alpha _2\wedge \omega \), and \(* \omega = \frac{1}{2} \omega ^2\). This implies that \(L_\omega : \langle \omega \rangle \oplus P^{1,1}\oplus P^{2,0} \rightarrow \langle \omega ^2\rangle \oplus \omega P^{1,1}\oplus \omega P^{2,0}\) is of the form \(L_\omega (\alpha )=L_\omega (\alpha _0+\alpha _1+\alpha _2)=\frac{1}{2} *\alpha _0-*\alpha _1+* \alpha _2\), where \(\alpha =\alpha _0+\alpha _1+\alpha _2\) is the decomposition according to \(H^2=\langle \omega \rangle \oplus P^{1,1}\oplus P^{2,0}\).

The Lefschetz map \(L_\omega :H^2\rightarrow H^4\) is an isomorphism so there is an inverse \(L_\omega ^{-1}:H^4\rightarrow H^2\). Using it, we can define a map \({\mathcal {F}}: P \times P\times P\times P \rightarrow {\mathbb {R}}\) by

$$\begin{aligned} {\mathcal {F}}(\alpha ,\beta ,\gamma ,\delta )=\int _M L_\omega ^{-1}(\alpha \wedge \beta ) \wedge \gamma \wedge \delta . \end{aligned}$$

This clearly factors through \({{\mathrm{Sym}}}^2P \times {{\mathrm{Sym}}}^2P\). Using (5) above, we have the alternative description

$$\begin{aligned} {\mathcal {F}}(\alpha ,\beta ,\gamma ,\delta )= 2\langle \left( \alpha \wedge \beta \right) _0, \left( \gamma \wedge \delta \right) _0 \rangle - \langle \left( \alpha \wedge \beta \right) _1, \left( \gamma \wedge \delta \right) _1 \rangle +\langle \left( \alpha \wedge \beta \right) _2, \left( \gamma \wedge \delta \right) _2 \rangle , \end{aligned}$$

from where it follows that \({\mathcal {F}}\) factors as a map  \({{\mathrm{Sym}}}^2({{\mathrm{Sym}}}^2P)\rightarrow {\mathbb {R}}\).

Let \({\mathcal {K}}_M\) be the kernel of the map \({{\mathrm{Sym}}}^2({{\mathrm{Sym}}}^2 P)\rightarrow {{\mathrm{Sym}}}^4 P\). Then we define a map

$$\begin{aligned} {\mathcal {F}}_M={\mathcal {F}}|_{{\mathcal {K}}_M}: {\mathcal {K}}_M\rightarrow {\mathbb {R}}. \end{aligned}$$

We have the following result.

Theorem 15

Let M be a simply connected compact Sasakian 7-dimensional manifold. Then M is formal if and only if \({\mathcal {F}}_M=0\).

Proof

Using Theorem 12, we only have to check whether M is 3-formal. For this we have to construct the minimal model \(\rho :(\bigwedge V,d) \rightarrow {\mathcal {M}}= (H\otimes \bigwedge (x),d)\) up to degree 3. This is easy:

$$\begin{aligned} V^1&=0, \\ V^2&= P, \\ V^3&= H^3 \oplus N^3, \qquad \text {where } N^3={{\mathrm{Sym}}}^2P, \end{aligned}$$

where the differential is given by \(d=0\) on P and \(H^3\), and \(d:N^3 \rightarrow \bigwedge V^2\) is the isomorphism \({{\mathrm{Sym}}}^2P \rightarrow \bigwedge ^2 P\). The map \(\rho \) is given as follows. \(\rho : V^2 =P\rightarrow {\mathcal {M}}^2=H^2\) is defined as the obvious (inclusion) map, \(\rho : H^3 \rightarrow {\mathcal {M}}^3=H^3 \oplus H^2 x\) is the inclusion on the first summand, and \(\rho :N^3={{\mathrm{Sym}}}^2P \rightarrow {\mathcal {M}}^3=H^3 \oplus H^2 x\) is defined as \(\rho (\alpha \cdot \beta )=L_{\omega }^{-1}(\alpha \wedge \beta ) \, x\). Note that

$$\begin{aligned}&d( \rho (\alpha \cdot \beta ))=L_{\omega }^{-1}(\alpha \wedge \beta ) \, dx= L_{\omega }^{-1}(\alpha \wedge \beta )\omega \\&\qquad = \alpha \wedge \beta =\rho (\alpha )\wedge \rho (\beta )=\rho (\alpha \wedge \beta )= \rho \left( d(\alpha \cdot \beta )\right) , \end{aligned}$$

so \(\rho \) is a DGA map. Clearly it is a 3-equivalence (it induces an isomorphism on cohomology up to degree 3 and an inclusion on degree 4).

The space of closed elements is \(C^3=H^3\). Now let us check when the elements \(z\in I(N^3)\) with \(dz=0\) satisfy \([\rho (z)]=0\in H^*(M)\). The only cases to check is when z has degree 5 or 7. If z has degree 5, then \([\rho (z)]\ne 0\) if and only if there exists some \(\beta \in P,\,[\rho (\beta )]\in H^2(M)\), such that \([\rho (z)]\wedge [\rho (\beta )]\ne 0\), by Poincaré duality. Hence \([\rho (z\beta )]\ne 0\). This means that we can restrict to elements z of degree 7, that is \(z \in N^3\cdot \bigwedge ^2 P\).

Let \(z \in N^3\cdot \bigwedge ^2 P \cong {{\mathrm{Sym}}}^2P \times {{\mathrm{Sym}}}^2 P\). Then the map \(d: N^3\cdot \bigwedge ^2P \rightarrow \bigwedge ^4P\) coincides the full symmetrization map \( {{\mathrm{Sym}}}^2P \times {{\mathrm{Sym}}}^2 P \rightarrow {{\mathrm{Sym}}}^4P\). So

$$\begin{aligned} Z= \ker d|_{I(N^3)^7}={\mathcal {K}}_M \oplus {{\mathrm{Ant}}}^2\left( {{\mathrm{Sym}}}^2P\right) , \end{aligned}$$

where \({{\mathrm{Ant}}}^2(W)\) denotes the antisymmetric 2-power of a vector space W.

Now we have to study the map

$$\begin{aligned} \rho : Z \rightarrow H^7(M) =H^6 x, \end{aligned}$$

and see if this is non-zero. This is given (on the basis elements) by

$$\begin{aligned} \rho \left( (\alpha \cdot \beta )\cdot (\gamma \cdot \delta )\right) = \left( L_{\omega }^{-1}(\alpha \wedge \beta ) \wedge \gamma \wedge \delta \right) x, \end{aligned}$$

so \({\mathcal {F}}_M=\rho |_{{\mathcal {K}}_M}\). Note that \(\rho \) automatically vanishes on \( {{\mathrm{Ant}}}^2({{\mathrm{Sym}}}^2P)\), hence M is formal if and only if \(\rho \) vanishes on \({\mathcal {K}}_M\) if and only if \({\mathcal {F}}_M=0\).

According to Theorem 12, to check non-formality we have to test the relevant property (2) on any splitting \(V^3=C^3+ N'{}^3\). If we take another splitting \(V^3=C^3+ N'{}^3\), then the projection \(\pi :V^3\rightarrow N^3\) gives an isomorphism \(\pi :N'{}^3 \rightarrow N^3\), and so an isomorphism \(N'{}^3\cdot {{\mathrm{Sym}}}^2P \cong N^3 \cdot {{\mathrm{Sym}}}^2P\). Clearly, \(d\circ \pi =d\) on \(N'{}^3\), so the spaces of cycles correspond \({\mathcal {K}}' \cong {\mathcal {K}}\). On the other hand \(H^3 \cdot H^2\cdot H^2 =0\), so the maps \(\rho :{\mathcal {K}}\rightarrow H^6 x\) and \(\rho :{\mathcal {K}}'\rightarrow H^6 x\) also correspond. This means that the corresponding \({\mathcal {F}}\) and \({\mathcal {F}}'\) coincide under the isomorphism \({\mathcal {K}}\cong {\mathcal {K}}'\). This means that the choice of splitting is not relevant. \(\square \)

This result means that the formality or non-formality of M only depends on the cohomology algebra H. Theorem 15 can be applied to the examples in Section 5.3 of [2]. For instance for \(B={\mathbb {C}}P^1\times {\mathbb {C}}P^1\times {\mathbb {C}}P^1\), we have a simply connected Sasakian 7-manifold which is non-formal (Theorem 12 of [2]). For \(B={\mathbb {C}}P^3\), we have obviously \(P=0\) and hence M is formal.

The element \({\mathcal {F}}_M\) of Theorem 15 is the principal Massey product defined by Crowley and Nordström [6] for simply connected compact 7-manifolds in general. The principal Massey product is the full obstruction to formality for simply connected compact 7-manifolds.

Now we deduce Corollary 3.

Corollary 16

Let M be a simply connected compact Sasakian 7-dimensional manifold. Then M is formal if and only if all triple Massey products are zero.

Proof

Suppose that \({\mathcal {F}}_M\ne 0\). We choose an orthonormal basis for \(H^2=\langle e_0,e_1,\ldots , e_m\rangle \), where \(e_0=\frac{1}{\sqrt{3}} \omega ,\,P^{1,1}=\langle e_1,\ldots ,e_s\rangle ,\,P^{2,0}=\langle e_{s+1},\ldots , e_m\rangle \). The vector space \({\mathcal {K}}_M\) is generated by elements of the form

$$\begin{aligned} a_{ijkl}= \left( e_i \cdot e_j\right) \cdot (e_k\cdot e_l) - (e_k \cdot e_j)\cdot \left( e_i\cdot e_l\right) , \end{aligned}$$

for \(0\le i,j,k,l \le m\) (here, as usual, the dot product means symmetric product). Now define the numbers

$$\begin{aligned} \lambda _{ijk}= \int _M e_i\wedge e_j\wedge e_k \in {\mathbb {R}}, \end{aligned}$$

for \(0\le i,j,k \le m\). Note that these numbers are fully symmetric on ijk. Also \(\lambda _{000}= \frac{2}{\sqrt{3}}\) and \(\lambda _{ij0}=\frac{1}{\sqrt{3}}\varepsilon _i\delta _{ij}\), for \((i,j)\ne (0,0)\), where \(\varepsilon _i=-1\) for \(1\le i \le s\) and \(\varepsilon _i=1\) for \(s+1\le i \le m\). Then

$$\begin{aligned} L_\omega ^{-1}(e_i\wedge e_j)=2*(e_i\wedge e_j)_0 - (e_i\wedge e_j)_1+(e_i\wedge e_j)_2 =2\lambda _{ij0}e_0 +\sum _{t>0} \varepsilon _t \lambda _{ijt}e_t. \end{aligned}$$

So

$$\begin{aligned} {\mathcal {F}}_M ( (e_i \cdot e_j)\cdot (e_k\cdot e_l))=2\lambda _{ij0}\lambda _{kl0} +\sum _{t>0}\varepsilon _t\lambda _{ijt}\lambda _{klt}. \end{aligned}$$

Evaluating \({\mathcal {F}}_M\) on \(a_{ijkl}\) gives a set of equations to determine the formality of M. M is non-formal when there exists some \(a_{ijkl}\) with \({\mathcal {F}}_M(a_{ijkl})\ne 0\). By [6], we have that the triple Massey product \(\langle e_i, e_j, e_k\rangle \) is a well-defined element of \(H^5(M)\) and it satisfies

$$\begin{aligned} {\mathcal {F}}_M(a_{ijkl}) =\langle e_i, e_j, e_k\rangle \cup e_l. \end{aligned}$$

So \(\langle e_i, e_j, e_k\rangle \ne 0\), as required. \(\square \)

This result is of relevance since it is not known if for general simply connected compact 7-dimensional manifolds there are obstructions to formality different from triple Bianchi-Massey tensor, as remarked in [6]. It is true that for higher dimensional manifolds, there are obstructions to formality even when all Massey products (triple and higher order) can be zero.