Hostname: page-component-857557d7f7-9f75d Total loading time: 0 Render date: 2025-11-21T23:32:18.307Z Has data issue: false hasContentIssue false

Pin±-structures on non-oriented 4-manifolds via Lefschetz fibrations

Published online by Cambridge University Press:  11 November 2025

Valentina Bais*
Affiliation:
Department of Mathematics, SISSA, via Bonomea 265, 34136 Trieste, Italy (vbais@sissa.it)
Rights & Permissions [Opens in a new window]

Abstract

We study necessary and sufficient conditions for a 4-dimensional Lefschetz fibration over the 2-disk to admit a ${\text{Pin}}^{\pm}$-structure, extending the work of A. Stipsicz in the orientable setting. As a corollary, we get existence results of ${\text{Pin}}^{+}$ and ${\text{Pin}}^-$-structures on closed non-orientable 4-manifolds and on Lefschetz fibrations over the 2-sphere. In particular, we show via three explicit examples how to read-off ${\text{Pin}}^{\pm}$-structures from the Kirby diagram of a 4-manifold. We also provide a proof of the well-known fact that any closed 3-manifold M admits a ${\text{Pin}}^-$-structure and we find a criterion to check whether or not it admits a ${\text{Pin}}^+$-structure in terms of a handlebody decomposition. We conclude the paper with a characterization of ${\text{Pin}}^+$-structures on vector bundles.

Information

Type
Research Article
Creative Commons
Creative Common License - CCCreative Common License - BY
This is an Open Access article, distributed under the terms of the Creative Commons Attribution licence (http://creativecommons.org/licenses/by/4.0), which permits unrestricted re-use, distribution and reproduction, provided the original article is properly cited.
Copyright
© The Author(s), 2025. Published by Cambridge University Press on behalf of The Royal Society of Edinburgh.

1. Introduction

${\text{Pin}}^{\pm}$-structures can be thought of as the non-oriented analogue of Spin-structures. We refer the reader to [Reference Kirby and Taylor17] and [Reference Stolz24, Section 2] for a precise definition of such Lie groups and for the background on ${\text{Pin}}^{\pm}(n)$-structures on low dimensional manifolds. In this note, we find necessary and sufficient conditions for a (possibly non-orientable) Lefschetz fibration over the 2-disk to support a ${\text{Pin}}^{-}$ or a ${\text{Pin}}^+$-structure. In particular, such conditions are expressed in terms of the homology classes of the vanishing cycles of the Lefschetz fibration, with coefficients taken in ${\mathbb{Z}}_2$ and ${\mathbb{Z}}_4$ respectively. As an application of this criterion, we provide explicit necessary and sufficient conditions for a given Lefschetz fibration over the 2-sphere to support a ${\text{Pin}}^{\pm}$-structure. We refer the reader to Section 3 for a quick recap on basic facts about non-orientable Lefschetz fibrations. Our main reference for this subject is [Reference Miller and Özbağcı19]. We also remark that in this note we consider ${\text{Pin}}^{\pm}$-structures on the tangent bundles of the manifolds under consideration, while in [Reference Hambleton, Kreck and Teichner13] they are defined on the normal bundles of embedded submanifolds.

In Section 4, we prove the following statement, extending the work of A. Stipsicz on Spin-structures over Lefschetz fibrations in [Reference Stipsicz23, Theorem 1.1].

Theorem 1. Let X be a smooth 4-manifold and $f: X \to D^2$ a Lefschetz fibration with regular fibre Σ. There is no ${\text{Pin}}^-$-structure on X if and only if there are k + 1 vanishing cycles $c_0, c_1, \dots, c_k$ such that $[c_0]=\sum_{i=1}^k [c_i] \in H_1(\Sigma; {\mathbb{Z}}_2)$ and $k+\sum_{1 \leq i \lt j \leq k} c_i \cdot c_j \equiv 0 \mod{2}$.

As we will see in Section 4, Theorem 1 is a consequence of the fact that ${\text{Pin}}^-$-structures on a Lefschetz fibration over the 2-disk with regular fibre Σ naturally correspond to maps

\begin{equation*} q^-: H_1(\Sigma;{\mathbb{Z}}_2) \rightarrow {\mathbb{Z}}_4 \end{equation*}

such that $q^-(x+y)=q^-(x)+q^-(y)+2x \cdot y$ for every $x, y \in H_1(\Sigma;{\mathbb{Z}}_2)$ and $q^-([c])=2$ on every vanishing cycle $c \subset \Sigma$, see Lemma 1. Moreover, such correspondence is equivariant with respect to the action of $H^1(X;{\mathbb{Z}}_2)$.

Section 5 is devoted to the study of ${\text{Pin}}^+$-structures on Lefschetz fibrations. By [Reference Degtyarev and Finashin7], a ${\text{Pin}}^+$-structure on the regular fibre Σ corresponds to a map

\begin{equation*}q_0^+: H_1(\Sigma;{\mathbb{Z}}_4) \rightarrow {\mathbb{Z}}_2\end{equation*}

with the property that $q_0^+(x+y)=q_0^+(x)+ q_0^+(y)+x \cdot y \in {\mathbb{Z}}_2$ for all $x,y \in H_1(\Sigma;{\mathbb{Z}}_4)$. We show that we can choose $q_0^+$ to be the restriction of a ${\text{Pin}}^+$-structure defined on the whole fibration if and only if the following holds.

Theorem 2. The total space of the Lefschetz fibration $f:X \to D^2$ with vanishing cycles $c_1, \dots, c_n \subset \Sigma$ supports a ${\text{Pin}}^+$-structure if and only if Σ supports a ${\text{Pin}}^+$-structure and

\begin{equation*}\text{rank} (C)= \text{rank}(C \mid A)\end{equation*}

where C is the ${\mathbb{Z}}_2$-reduction of the n × r matrix $(c_{ij})$ whose rows are given by the components of $[c_1], \dots, [c_n] \in H_1(\Sigma;{\mathbb{Z}}_4) \cong \mathbb{Z}^r$ with respect to a fixed basis $e_1, \dots, e_r$ of the free ${\mathbb{Z}}_4$-module $H_1(\Sigma;{\mathbb{Z}}_4)$ and A is the column vector with entries $A_i=1+q_0^+([c_i]) \in {\mathbb{Z}}_2$ for $i=1, \dots, n$.

A key tool in the proof of the above result is Lemma 2, in which we show that ${\text{Pin}}^+$-structures on X correspond to maps

\begin{equation*}q^+: H_1(\Sigma;{\mathbb{Z}}_4) \rightarrow {\mathbb{Z}}_2\end{equation*}

such that $q^+(x+y)=q^+(x)+q^+(y)+x\cdot y$ for every $x,y \in H_1(\Sigma;{\mathbb{Z}}_4)$ and $q^+([c])=1$ on every vanishing cycle $c \subset \Sigma$. The correspondence is also in this case equivariant with respect to the $H^1(X;{\mathbb{Z}}_2)$-action.

We remark that the study of ${\text{Pin}}^{+}$ and ${\text{Pin}}^-$-structures on non-orientable 4-manifolds is essentially reduced to the case of non-orientable Lefschetz fibrations over the 2-disk due to the following result, which is a straightforward consequence of [Reference Miller and Özbağcı19, Theorem 1.1].

Theorem 3 (Miller-Özbağcı [Reference Miller and Özbağcı19])

Let X be a closed non-orientable smooth 4-manifold. There is a decomposition

\begin{equation*}X=L \cup_{\partial} H\end{equation*}

where L is a non-orientable Lefschetz fibration over the 2-disk and H is a non-orientable 1-handlebody.

In particular, the proof of [Reference Miller and Özbağcı19, Theorem 1.1] shows an explicit way of endowing a (possibly non-orientable) 2-handlebody with the structure of a Lefschetz fibration over the 2-disk, see also [Reference Harer14] and [Reference Etnyre and Fuller8]. Moreover, a 4-manifold $X=L \cup_{\partial} H$ as in Theorem 3 admits a ${\text{Pin}}^{\pm}$-structure if and only if L does. This latter conditions can be checked using Theorem 1 and Theorem 2.

In Section 6 we provide explicit examples in which, starting from a handlebody decomposition of a non-orientable 4-manifold, we endow its 2-handlebody with the structure of a Lefschetz fibration over the 2-disk. We then apply our results in order to check whether such 4-manifolds support a ${\text{Pin}}^-$ or a ${\text{Pin}}^+$-structure. In the case they do, we find all the possible structures by looking at the $H^1(X,{\mathbb{Z}}_2)$-action on the associated quadratic enhancements on the regular fibres. The non-orientable handlebodies under consideration describe the three 4-manifolds $\mathbb{R} \mathbb{P}^4$, $S^2 \mathbin{\widetilde{\smash{\times}}} \mathbb{R} \mathbb{P}^2$ and $S^2 \times \mathbb{R} \mathbb{P}^2$ and can be found in [Reference Akbulut, Gordon and Kirby2], [Reference Bais and Torres4], [Reference Miller and Naylor18] and [Reference Torres25]. In particular, that there are two non-orientable S 2-bundles over $\mathbb{R} \mathbb{P}^2$ and we denote by $S^2 \mathbin{\widetilde{\smash{\times}}} \mathbb{R} \mathbb{P}^2$ the non-trivial one, see [Reference Bais and Torres4, Section 2.3].

Section 7 is devoted of the study of necessary and sufficient conditions for the existence of a ${\text{Pin}}^{\pm}$-structure on a Lefschetz fibration over the 2-sphere. This is the non-orientable version of [Reference Stipsicz23, Theorem 1.3].

Theorem 4. Let $f: X \rightarrow S^2$ be a (possibly non-orientable) Lefschetz fibration with regular fibre Σ. Then X supports a ${\text{Pin}}^-$-structure if and only if $X \setminus \nu(\Sigma)$ does and there exists a smoothly embedded surface $\sigma \subset X$ which is dual to Σ in $H_2(X;{\mathbb{Z}}_2)$ and such that

\begin{equation*}[\sigma]^2+(w_1(\sigma)\cup w_1(\nu(\sigma))([\sigma])+w_1^2(\nu(\sigma))=0\in {\mathbb{Z}}_2.\end{equation*}

Analogously, X supports a ${\text{Pin}}^+$-structure if and only if $X \setminus \nu(\Sigma)$ does and there exists a smoothly embedded surface $\sigma \subset X$ which is dual to Σ in $H_2(X;{\mathbb{Z}}_2)$ and such that

\begin{equation*}\chi(\sigma)+[\sigma]^2+(w_1(\sigma)\cup w_1(\nu(\sigma)))([\sigma])=0\in {\mathbb{Z}}_2.\end{equation*}

In analogy to the Spin case, one can define the ${\text{Pin}}^{+}$ and ${\text{Pin}}^-$ cobordism groups in each dimension, see [Reference Kirby and Taylor17, Section 1]. As sets, they consist of the equivalence classes of closed n-dimensional smooth manifolds with a fixed ${\text{Pin}}^{+}$ (resp. ${\text{Pin}}^-$)-structure, which are identified whenever they co-bound a ${\text{Pin}}^{+}$ (resp. ${\text{Pin}}^-$) $(n+1)$-manifold. The group operation is the one induced by the disjoint union of manifolds. We remark that ${\text{Pin}}^+$ and ${\text{Pin}}^-$ cobordism groups are drastically different. In particular, the 4-dimensional ${\text{Pin}}^+$-cobordism group is

\begin{equation*} \Omega_4^{{\text{Pin}}^+} \cong {\mathbb{Z}}_{16}\end{equation*}

(see [Reference Giambalvo10, Section 2] and [Reference Kirby and Taylor17, Section 3]) and the η-invariant modulo $2 \mathbb{Z}$ of the twisted Dirac operator associated to a ${\text{Pin}}^+$-structure on a 4-manifold is a complete invariant for ${\text{Pin}}^+$-cobordism classes, see [Reference Stolz24]. In particular, there are instances in which such invariant can detect exotic behaviours in the non-orientable 4-dimensional realm, in the sense that there are examples of non-orientable ${\text{Pin}}^+$ 4-manifolds which share the same homeomorphism type but define two different classes in $\Omega_4^{{\text{Pin}}^+}$ and hence cannot be diffeomorphic. To the best of our knowledge, following a suggestion in [Reference Gilkey11], this was shown for the first time in [Reference Stolz24], where the η-invariant of ${\text{Pin}}^+$-structures is used to detect the exotic $\mathbb{R} \mathbb{P}^4$ constructed in [Reference Cappell and Shaneson5]. We remark that another exotic $\mathbb{R} \mathbb{P}^4$ is constructed in [Reference Fintushel and Stern9] and its η-invariant is computed in [Reference Olędzki21]. An analogous approach has then been used also in [Reference Bais and Torres4] and [Reference Torres25]. On the other hand

\begin{equation*}\Omega_4^{{\text{Pin}}^-} =0\end{equation*}

(see again [Reference Anderson, Brown and Peterson3, Theorem 5.1] and [Reference Kirby and Taylor17, Section 3]) and hence it is not possible to find ${\text{Pin}}^-$ exotic 4-manifolds using this strategy. In particular, the study of ${\text{Pin}}^+$-structures on 4-manifolds could potentially lead to better understanding of exotic behaviours in the non-orientable realm and this fact is one of the main motivations for this note. In particular, in Section 5 we show that if $X =L \cup H$ is a closed 4-manifold with a decomposition as in the statement of Theorem 3, then the datum of a ${\text{Pin}}^+$-structure on X is equivalent to the datum of a map

\begin{equation*}q^+: H_1(\Sigma;{\mathbb{Z}}_4)\rightarrow {\mathbb{Z}}_2\end{equation*}

such that $q^+(x+y)=q^+(x)+q^+(y)+ x \cdot y$ for all $x,y \in H_1(\Sigma;{\mathbb{Z}}_4)$ taking value $1 \in {\mathbb{Z}}_2$ on all the vanishing cycles of L, where Σ denotes a regular fibre. We leave the reader with the following question.

Question: Is there a formula for the η-invariant modulo $2 \mathbb{Z}$ of a ${\text{Pin}}^+$-structure on X in terms of the corresponding map $q^{+}:H_1(\Sigma;{\mathbb{Z}}_4) \rightarrow {\mathbb{Z}}_2$?

Section 2 contains a short discussion on how to check the vanishing of w 2 and $w_1^2$ in terms of the embeddings of homologically essential surfaces inside the ambient 4-manifold. In Section 7 we study Lefschetz fibrations over the 2-sphere, 8 is devoted to a brief discussion on the 3-dimensional case, while Section 9 contains a characterization of ${\text{Pin}}^+$-structures on vector bundle which resembles Milnor’s characterization of Spin-structures, see [Reference Milnor20, Alternative definitions 2].

All manifolds in this note are smooth. We will not assume our 4-manifolds to be orientable, unless otherwise stated.

2. ${\text{Pin}}^{\pm}$-structures and embedded surfaces

For every $n \in \mathbb{N}$, there are two distinct central extensions

(1)\begin{equation} {\text{Pin}}^{\pm}(n) \rightarrow O(n) \end{equation}

which are topologically a disjoint union $\text{Spin}(n) \cup \text{Spin}(n)$. In the orientable setting, it is known that the existence of a Spin-structure on a real vector bundle ξ on a manifold X is equivalent to the vanishing of $w_2(\xi) \in H^2(X,{\mathbb{Z}}_2)$, where wi is the ith Stiefel-Whitney class of ξ. In the case in which X is 4-dimensional, this equivalent to having an even intersection form, provided that $H^2(X;\mathbb{Z})$ has no 2-torsion [Reference Gompf and Stipsicz12, Corollary 5.7.6]. This can be shown via the Wu formula, which states that for any $a \in H_2(X;{\mathbb{Z}}_2)$ one has

(2)\begin{equation}\langle w_2(X),a\rangle = a^2 \text{ modulo} \ 2\end{equation}

where a 2 denotes the algebraic count of self-intersections of a surface in X representing a, see [Reference Gompf and Stipsicz12, Proposition 1.4.18]. In a similar way, there are cohomological obstructions to the existence of ${\text{Pin}}^+$ and ${\text{Pin}}^-$-structures on the tangent bundle of a manifold, namely the vanishing of $w_2(X)$ and of $(w_2+w_1^2)(X)$ respectively, see [Reference Kirby and Taylor17, Lemma 1.3]. Moreover, both ${\text{Pin}}^{+}$ and ${\text{Pin}}^-$-structures (when they do exist) are torsors over $H^1(X,{\mathbb{Z}}_2)$, meaning that there is a free and transitive action of $H^1(X;{\mathbb{Z}}_2)$ over the sets of such geometric structures. However, the Wu formula does not hold when X is non-orientable, but one can still interpret the vanishing of $w_2(X)$ and $(w_2+w_1^2)(X)$ in terms of properties of embedded surfaces. Indeed, for any 4-manifold X and $a \in H_2(X;{\mathbb{Z}}_2)$ there is a (possibly non-orientable) smoothly embedded surface $\Sigma \subset X$ representing a (see [Reference Gompf and Stipsicz12, Remark 1.2.4]) and one can show that

(3)\begin{equation} \langle w_2(X),a\rangle = \chi(\Sigma) +(w_1(\Sigma) \smile w_1(\nu(\Sigma)))([\Sigma])+[\Sigma]^2 \text{ modulo } 2. \end{equation}

Since the tangent bundle of a 4-manifold supports a ${\text{Pin}}^+$-structure if and only if $w_2(X)$ vanishes, this is equivalent to checking that the evaluation (3) is trivial on any homology class $a \in H_2(X;{\mathbb{Z}}_2)$.

On the other hand, the obstruction for ${\text{Pin}}^-$-structures to exist is $(w_2+w_1^2)(X)$ [Reference Kirby and Taylor17, Lemma 1.3] so, in order to understand when such structures exist, we need to compute $w_1^2(X)$ and one can easily show that

(4)\begin{equation} \langle w_1^2(X),[\Sigma] \rangle = w_1^2(\Sigma)+ w_1^2(\nu(\Sigma)). \end{equation}

Example 1 ( $\mathbb{R} \mathbb{P}^4$)

The 4-dimensional real projective space $\mathbb{R} \mathbb{P}^4$ is an example of non-orientable 4-manifold supporting two distinct ${\text{Pin}}^+$-structures but no ${\text{Pin}}^-$-structure. We have that $H_2(\mathbb{R} \mathbb{P}^4;{\mathbb{Z}}_2) \cong {\mathbb{Z}}_2$ is generated by the homology class of $\mathbb{R} \mathbb{P}^2$. Moreover, the tubular neighbourhood $\nu(\mathbb{R} \mathbb{P}^2) \subset \mathbb{R} \mathbb{P}^4$ is diffeomorphic to the twisted 2-disk bundle $D^2 \mathbin{\widetilde{\smash{\times}}} \mathbb{R} \mathbb{P}^2$ defined as the quotient of $D^2\times S^2$ via the involution

\begin{align*} D^2 \times S^2 \rightarrow D^2 \times S^2 \\ (x,y) \mapsto (\rho_{\pi}(x), -y) \end{align*}

where ρπ denotes the rotation of S 2 of π radians about a fixed axis. One can show that such a quotient is diffeomorphic to

(5)\begin{equation} D^2 \mathbin{\widetilde{\smash{\times}}} \mathbb{R} \mathbb{P}^2 \cong (D^2 \times D^2) \cup_{\varphi} (D^2 \times \text{Mb}) \end{equation}

where Mb denotes the Möbius strip and φ is the map

\begin{equation*} \varphi: D^2 \times S^1 \rightarrow D^2 \times S^1, \quad (x, \theta) \mapsto (\rho_{\theta}(x), \theta) \end{equation*}

and $\rho_{\theta}: D^2 \rightarrow D^2$ denotes the rotation of the 2-disk of angle θ with respect to the origin, see [Reference Akbulut1, Section 0] and [Reference Bais and Torres4, Example 7].

By (3) we have that

\begin{equation*} \langle w_2(\mathbb{R} \mathbb{P}^4), [\mathbb{R} \mathbb{P}^2] \rangle = \chi(\mathbb{R} \mathbb{P})^2+ (w_1(\mathbb{R} \mathbb{P}^2)\cup w_1(\nu (\mathbb{R} \mathbb{P}^2)))([\mathbb{R} \mathbb{P}^2])+[\mathbb{R} \mathbb{P}^2]^2=0 \in {\mathbb{Z}}_2 \end{equation*}

since $\chi(\mathbb{R} \mathbb{P}^2)=1=[\mathbb{R} \mathbb{P}^2]^2$ and $w_1(\nu (\mathbb{R} \mathbb{P}^2))=0$, while (4) implies that

\begin{equation*} \langle w_1(\mathbb{R} \mathbb{P}^4)^2, [\mathbb{R} \mathbb{P}^2] \rangle = w_1^2(\mathbb{R} \mathbb{P}^2) + w_1^2(\nu (\mathbb{R} \mathbb{P}^2))= 1+0=1\in {\mathbb{Z}}_2 \end{equation*}

where $w_1^2(\mathbb{R} \mathbb{P}^2)$ is computed as the self-intersection of a loop in $\mathbb{R} \mathbb{P}^2$ whose ${\mathbb{Z}}_2$-homology class generates $H_1(\mathbb{R} \mathbb{P}^2;{\mathbb{Z}}_2)$.

3. Non-orientable Lefschetz fibrations

Since the conditions 3 and 4 are not immediate to check in the non-orientable setting and depend heavily on the embeddings of the homologically essential surfaces, in the following sections we develop another way to combinatorially understand when a 4-manifold is ${\text{Pin}}^+$ or ${\text{Pin}}^-$ by means of a specific kind of decomposition. To do this, we will need the notion of non-orientable Lefschetz fibration over the 2-disk.

We start by recalling the definition of Lefschetz fibration. Our main reference for this topic in the non-orientable realm is [Reference Miller and Özbağcı19].

Definition 1. Let X be a compact, connected 4-manifold and let B be a compact, connected surface, both with possibly non-empty boundary. A Lefschetz fibration is a smooth submersion

\begin{equation*}f : X \rightarrow B \end{equation*}

away from finitely many points in the interior of B such that each fibre contains at most one critical point and f is a fibre bundle with surface fibre on the complement of the critical values. Moreover, around each critical point, we require f to conform to the local complex model

\begin{equation*}(z_1, z_2) \rightarrow z_1z_2.\end{equation*}

Remark 1. If X is non-orientable and B is oriented, each regular fibre is a non-orientable surface. In this case it makes no sense to ask for the orientation of the local model around a critical point to be compatible to the one of X.

In the following, we will just consider the case in which $B=D^2$ and Σ is the regular fibre. It is possible to show that every Lefschetz fibration

\begin{equation*}f: X \rightarrow D^2\end{equation*}

is obtained from the product one

\begin{equation*}\Sigma \times D^2 \rightarrow D^2\end{equation*}

by gluing 4-dimensional 2-handles along $\Sigma \times \partial D^2$ with framing ±1 with respect to the fibre framing. The attaching curves of such 2-handles are push-offs in distinct fibres of simple closed curves inside Σ, which are called the vanishing cycles of the fibration and are denoted in the following by $c_1, \dots, c_n \subset \Sigma$. Moreover, $\langle w_1(X),[c_i] \rangle =0$ for all $i=1, \dots, n$ and hence the tubular neighbourhood of these attaching regions is necessarily diffeomorphic to a product of S 1 with the unit interval. We will call two-sided all curves with this property.

4. ${\text{Pin}}^-$-structures on 4-manifolds via Lefschetz fibrations

It is well known that the datum of a Spin-structure over an orientable surface is equivalent to the one of a quadratic enhancement

\begin{equation*} s: H_1(\Sigma;{\mathbb{Z}}_2) \to {\mathbb{Z}}_2\end{equation*}

satisfying the condition

\begin{equation*}s(x+y)=s(x)+s(y)+ x \cdot y\end{equation*}

for all $x,y \in H_1(\Sigma;{\mathbb{Z}}_2)$, where · denotes the ${\mathbb{Z}}_2$-intersection number between cycles, see [Reference Johnson16] and [Reference Stipsicz23]. A geometric interpretation of this algebraic object can be given as follows. The 1-dimensional Spin-cobordism group is $\Omega_1^{\text{Spin}}\cong {\mathbb{Z}}_2$ [Reference Milnor20] and every closed simple curve $\gamma \subset \Sigma$ in a Spin surface inherits a Spin-structure. In particular, if s is the quadratic enhancement associated to the fixed Spin-structure on Σ, then $s([\gamma])=0$ if and only if the induced Spin-structure on γ is the one bounding the unique Spin-structure on the 2-disk.

It is possible to give a similar description for ${\text{Pin}}^-$ structures on (not necessarily orientable) surfaces, as shown by the following result.

Theorem 5 (Kirby-Taylor, [Reference Kirby and Taylor17])

There is a 1:1 correspondence between ${\text{Pin}}^-$-structures on a surface Σ and quadratic enhancements

\begin{equation*} q^{-}: H_1(\Sigma; {\mathbb{Z}}_2) \to {\mathbb{Z}}_4\end{equation*}

with the property that

\begin{equation*}q^{-}(x+y)=q^{-}(x)+q^{-}(y)+2 \ x \cdot y\end{equation*}

for any $x,y \in H_1(\Sigma; {\mathbb{Z}}_2)$.

In particular, if $\gamma \subset \Sigma$ is a simple closed curve, then $q^{-}([\gamma])$ is even if and only if γ is two-sided and this is the case whenever γ is a vanishing cycle for a Lefschetz fibration. Moreover, recall that the 1-dimensional ${\text{Pin}}^-$-cobordism group is $\Omega_1^{{\text{Pin}}^-}\cong {\mathbb{Z}}_2$ (see [Reference Anderson, Brown and Peterson3, Theorem 5.1] and [Reference Kirby and Taylor17, Section 0]) and every closed simple curve $\gamma \subset \Sigma$ in a ${\text{Pin}}^-$ surface inherits a ${\text{Pin}}^-$ structure. We have that $q^{-}([\gamma])=0$ if and only if γ inherits the bounding ${\text{Pin}}^-$-structure.

In order to prove Theorem 1, we will need the following Lemma.

Lemma 1. Let $f: X \rightarrow D^2$ be a Lefschetz fibration with regular fibre Σ and vanishing cycles $c_1, \dots, c_n \subset \Sigma$. There is a natural one to one correspondence between the set of ${\text{Pin}}^-$-structures on X and the set of quadratic enhancements

\begin{equation*}q^-:H_1(\Sigma;{\mathbb{Z}}_2) \rightarrow {\mathbb{Z}}_4\end{equation*}

such that $q^-(x+y)=q^-(x)+q^-(y)+2x\cdot y$ for all $x, y \in H_1(\Sigma;{\mathbb{Z}}_2)$ and $q^-([c_i])=2$ for all $i=1, \dots, n$. Moreover, such correspondence is equivariant with respect to the free and transitive action of $H^1(X;{\mathbb{Z}}_2)$.

Proof. The existence of a ${\text{Pin}}^-$-structure on X is equivalent to the existence of a ${\text{Pin}}^-$-structure on Σ that extends to the 2-handles. This follows from the fact that any ${\text{Pin}}^-$-structure on X induces by restriction a ${\text{Pin}}^-$ structure on a (trivial) tubular neighbourhood $\nu(\Sigma)\cong \Sigma \times D^2$ and all the ${\text{Pin}}^-$-structures on $\Sigma \times D^2$ are pull-backs of the ones on Σ via the projection map $\Sigma \times D^2 \rightarrow \Sigma$, see also the proof of [Reference Stipsicz23, Theorem 1.1]. Since such 2-handles are attached to $\Sigma \times \partial D^2$ with relative odd framing, this corresponds to a ${\text{Pin}}^-$-structure on Σ restricting to the non-bounding one on every vanishing cycle. The conclusion follows from Theorem 5.

The free and transitive action of $H^1(X;{\mathbb{Z}}_2)$ on the set of ${\text{Pin}}^-$-structures on X can be seen as follows. In [Reference Kirby and Taylor17, Section 3], it is shown that $H^1(\Sigma;{\mathbb{Z}}_2)$ acts on the set of ${\text{Pin}}^-$-structures of the surface Σ by

(6)\begin{equation} q^{-}_{\gamma}(x)=q^{-}(x)+2 \cdot \gamma(x) \end{equation}

for all $\gamma \in H^1(\Sigma;{\mathbb{Z}}_2)$ and $x \in H_1(\Sigma;{\mathbb{Z}}_2)$. In particular, ${\text{Pin}}^-$-structures on Σ equivariantly correspond to quadratic enhancements and the action (6) is well defined on $H^1(X;{\mathbb{Z}}_2)\subset H^1(\Sigma;{\mathbb{Z}}_2)\cong H^1(X;{\mathbb{Z}}_2)\oplus K$, where $K \subset H^1(\Sigma;{\mathbb{Z}}_2)$ is the kernel of the map induced by the inclusion $\Sigma \subset X$.

In light of this fact, it is possible to characterize the existence of ${\text{Pin}}^-$-structures on non-orientable Lefschetz fibrations over the 2-disk in terms of the ${\mathbb{Z}}_2$-homology classes of the vanishing cycles.

Proof of Theorem 1

By Lemma 1, X does not support any ${\text{Pin}}^-$-structure if and only if it is not possible to build a map

\begin{equation*}q^-: H_1(\Sigma; {\mathbb{Z}}_2) \rightarrow {\mathbb{Z}}_4 \end{equation*}

such that $q^-(x+y)=q^-(x)+q^-(y)+2x\cdot y$ for all $x,y \in H_1(\Sigma;{\mathbb{Z}}_2)$ and such that $q^+([c_i])=2$ for all $i=1, \dots, n$. The remaining part of the proof is essentially the same as the proof of [Reference Stipsicz23, Theorem 1.1]. We will now sketch it for the convenience of the reader. Let $V\subset H_1(\Sigma;{\mathbb{Z}}_2)$ be the vector subspace generated by the classes of the vanishing cycles. Notice that any given enhancement $q^-$ on V can be extended to the whole $H_1(\Sigma;{\mathbb{Z}}_2)$ by setting $q^-=0$ on a basis $u_1, \dots, u_k$ of an orthogonal vector subspace and using the relation (6) to define it on the vectors of the form $\sum_j u_{i_j}+v$, where $v \in V$. In particular, in order to define a ${\text{Pin}}^-$-structure on X, one needs to find a quadratic enhancement $q^-: V \rightarrow {\mathbb{Z}}_4$ which takes value 2 on the class of any vanishing cycle. Let $v_1, \dots, v_r$ be a basis of V represented by vanishing cycles and let $V_i=\langle v_1, \dots, v_i \rangle$ for $i=1, \dots, r$. We then want to define $q^-$ inductively on each Vi as in the proof of [Reference Stipsicz23, Theorem 1.1], setting $q^-(v_i)=2$ at each step. One then notices that things go wrong if and only if there are k classes $v_{i_1}, \dots, v_{i_k}$ such that their sum $v_0=\sum_{j=1}^k v_{i_j} \in H_1(\Sigma;{\mathbb{Z}}_2)$ is again represented by a vanishing cycle and $q^{-}(v_0)\neq 2 \in {\mathbb{Z}}_4$. This implies that $q^-(v_0)=0 \in {\mathbb{Z}}_4$, being v 0 is represented by a curve with trivial tubular neighbourhood. Since by construction we had that $q^-(v_{i_j})=2$ for $j=1, \dots, k$, using (6) the condition $q^-(q_0)=0 \in Z_4$ translates into

\begin{equation*}q^-(v_0)=\sum_{j=1}^kq^-(v_{i_j})+ 2\cdot \Big( \sum _{1 \leq i_n \leq i_m \leq k} v_{i_n} \cdot v_{i_m} \Big)=2\cdot \Big(k+\sum _{1 \leq i_n \leq i_m \leq k} v_{i_n} \cdot v_{i_m}\Big)=0 \in {\mathbb{Z}}_4.\end{equation*}

The conclusion then follows by noticing that $2 \cdot x =0 \in {\mathbb{Z}}_4$ if and only if $x=0 \in {\mathbb{Z}}_2$.

Note that, if we restrict to orientable surfaces, there is a natural map

\begin{equation*}\Omega_1^{\text{Spin} }\to \Omega_1^{{\text{Pin}}^- }\end{equation*}

giving a group isomorphism, see [Reference Kirby and Taylor17, Theorem 2.1]. At the level of the associated quadratic forms, the enhancement

\begin{equation*}s: H_1(\Sigma;{\mathbb{Z}}_2)\to {\mathbb{Z}}_2\end{equation*}

corresponds to

\begin{equation*}q^{-}: H_1(\Sigma;{\mathbb{Z}}_2)\to {\mathbb{Z}}_4\end{equation*}

given by

\begin{equation*}q^{-}(x) = 2 \cdot s(x) \end{equation*}

for every $x \in H_1(\Sigma;{\mathbb{Z}}_2)$, where $2 \cdot$ denotes the inclusion ${\mathbb{Z}}_2 \subset {\mathbb{Z}}_4$. In particular, the condition on the vanishing cycles we found in Theorem 1 coincides with the one in [Reference Stipsicz23, Theorem 1.1] and this is due to the fact that, when restricting to orientable Lesfschetz fibrations over D 2, being Spin coincides with being ${\text{Pin}}^-$.

5. ${\text{Pin}}^+$-structures on Lefschetz fibrations over the 2-disk

As a consequence of [Reference Degtyarev and Finashin7, Theorem A], there is a canonical affine bijective correspondence between ${\text{Pin}}^+$-structures on a surface Σ and the set of maps

(7)\begin{equation} q^{+}: H_1(\Sigma;{\mathbb{Z}}_4) \to {\mathbb{Z}}_2 \end{equation}

with the property that

(8)\begin{equation} q^{+}(x+y)=q^{+}(x)+q^{+}(y)+x \cdot y \end{equation}

for every $x, y \in H_1(\Sigma;{\mathbb{Z}}_4)$, where $x \cdot y\in {\mathbb{Z}}_2$ denotes the algebraic intersection modulo 2 between representatives of the ${\mathbb{Z}}_2$-reductions of the classes $x,y \in H_1(\Sigma;{\mathbb{Z}}_4)$.

Recall that the 1-dimensional ${\text{Pin}}^+$-cobordism group is

\begin{equation*}\Omega_1^{{\text{Pin}}^+}\cong \{0\}\end{equation*}

see [Reference Kirby and Taylor17, Section 3]. However, S 1 has two distinct ${\text{Pin}}^+$-structures. These correspond to two different trivializations of the direct sum $TS^1 \oplus \varepsilon$ of the tangent bundle of S 1 with a trivial line bundle, where the first one is given by the restriction to S 1 of a trivialization of the tangent bundle $T D^2$ of the 2-disk and the second one comes from the Lie group framing of S 1. In particular, one can show that such structures have the property that they respectively bound the 2-disk and the Möbius strip, see [Reference Kirby and Taylor17, Section 1]. A close look at the construction of $q^{+}$ starting from a ${\text{Pin}}^+$-structure on Σ shows that $q^{+}([\gamma])=0$ if the induced ${\text{Pin}}^+$-structure on γ is the one bounding a 2-disk, while $q^{+}([\gamma])=1$ if it is the one bounding the Möbius strip, see [Reference Kirby and Taylor17]. Note that, given any two maps $q^{+}_1, q^{+}_2$ corresponding to ${\text{Pin}}^+$-structures on Σ, their difference can be regarded as an element

\begin{equation*}q^{+}_1 - q^{+}_2 \in\text{Hom}(H_1(\Sigma;{\mathbb{Z}}_4),{\mathbb{Z}}_2).\end{equation*}

The proof of Theorem 2 is based on the following Lemma.

Lemma 2. Let $f: X \rightarrow D^2$ be a Lefschetz fibration with regular fibre Σ and vanishing cycles $c_1, \dots, c_n \subset \Sigma$. There is a natural bijection between the set of ${\text{Pin}}^+$-structures on X and the set of quadratic enhancements

\begin{equation*}q^+:H_1(\Sigma;{\mathbb{Z}}_4) \rightarrow {\mathbb{Z}}_2\end{equation*}

such that $q^+(x+y)=q^+(x)+q^+(y)+x \cdot y$ for all $x, y \in H_1(\Sigma; {\mathbb{Z}}_4)$ and $q^+([c_i])=1$ for all $i=1, \dots, n$. Moreover, such correspondence is equivariant with respect to the free and transitive action of $H^1(X;{\mathbb{Z}}_2)$.

Proof. As is the ${\text{Pin}}^-$ case, the existence of a ${\text{Pin}}^+$-structure on X is equivalent to the existence of a ${\text{Pin}}^+$-structure on Σ that extends to the 2-handles. Since such 2-handles are attached to $\Sigma \times \partial D^2$ with relative odd framing, this corresponds to a ${\text{Pin}}^+$-structure on Σ restricting to the one bounding the Möbius strip on every vanishing cycle. The conclusion follows from [Reference Degtyarev and Finashin7].

In the ${\text{Pin}}^+$-case, $H^1(\Sigma;{\mathbb{Z}}_2)$ acts on the set of ${\text{Pin}}^+$-structures of the surface Σ by

(9)\begin{equation} q^{+}_{\gamma}(x)=q^{+}(x)+ \gamma(x) \end{equation}

for all $\gamma \in H^1(\Sigma;{\mathbb{Z}}_2)$ and $x \in H_1(\Sigma;{\mathbb{Z}}_4)$, see [Reference Degtyarev and Finashin7]. In particular, ${\text{Pin}}^+$-structures on Σ equivariantly correspond to quadratic enhancements and the action (9) is well defined on $H^1(X;{\mathbb{Z}}_2)\subset H^1(\Sigma;{\mathbb{Z}}_2)\cong H^1(X;{\mathbb{Z}}_2)\oplus K$, where $K \subset H^1(\Sigma;{\mathbb{Z}}_2)$ is the kernel of the map induced by the inclusion $\Sigma \subset X$.

Let $f: X \to D^2$ be a Lefschetz fibration over the 2-disk with surface fibre Σ. Let $c_1, \dots, c_n \subset \Sigma$ be its vanishing cycles. Note that one can find simple closed curves $e_1, \dots, e_g$ inducing a basis of $H_1(\Sigma;{\mathbb{Z}}_2)\cong {\mathbb{Z}}_2^g$. From now on, we will assume that Σ has a ${\text{Pin}}^+$-structure, since this is a necessary condition for X to be a ${\text{Pin}}^+$-manifold. In particular, this means that Σ is not a closed non-orientable surface of odd Euler characteristic. Let

(10)\begin{equation} q^{+}_0: H_1(\Sigma;{\mathbb{Z}}_4) \to {\mathbb{Z}}_2 \end{equation}

be the quadratic enhancement associated to a fixed ${\text{Pin}}^+$-structure on Σ and let $c_j=\sum_{i=1}^r c_{ji}[e_i].$ This gives the setting of Theorem 2, of which we now provide a proof.

Proof of Theorem 2

Lemma 2 implies that X has a ${\text{Pin}}^+$-structure if and only if there is a map

\begin{equation*}q^+: H_1(\Sigma,{\mathbb{Z}}_4) \rightarrow {\mathbb{Z}}_2\end{equation*}

such that $q^+(x+y)=q^+(x)+q^+(y)+x \cdot y$ for all $x, y \in H_1(\Sigma; {\mathbb{Z}}_4)$ and $q^+([c_i])=1$ for all $i=1, \dots, n$. In particular, every such $q^+$ is necessarily of the form $q^{+}=q^{+}_0+l$ for some linear map $l=\sum_{i=1}^r \lambda_i [e^i]$, where $[e^i]\in H^1(\Sigma;{\mathbb{Z}}_2)$ denotes the dual element to $[e_i]\in H_1(\Sigma;{\mathbb{Z}}_2)$. In particular, this reduces the problem to finding $l= \sum_{i=1}^r \lambda_i e^i$ such that for every $j=1, \dots, n$ we have

\begin{equation*} q^{+}([c_j])=q^{+}_0([c_j])+l([c_j])=q_0^+([c_j])+\sum_{i=1}^r c_{ji} \lambda_i=1.\end{equation*}

The conclusion follows from Rouché-Capelli’s Theorem [Reference Shafarevich and Remizov22, Theorem 2.38].

Remark 2. The proof of Theorem 1 as well as the one of [Reference Stipsicz23, Theorem 1.1] rely on the fact that $H_1(\Sigma; {\mathbb{Z}}_2)$ is a vector space. This is why we cannot adopt the same strategy for Theorem 2, since $H_1(\Sigma;{\mathbb{Z}}_4)$ is just a ${\mathbb{Z}}_4$-module.

6. Some examples

In this section, we show how to apply the results proven so far. In particular, starting from the handlebody decomposition of some small non-orientable 4-manifold M, we endow the union of handles up to index 2 with the structure of a Lefschetz fibration over the 2-disk, following the procedure explained in the proof of [Reference Miller and Özbağcı19, Theorem 1.1]. At this point, we apply Theorem 1 and Theorem 2 to understand whether or not M admits a ${\text{Pin}}^+$ or a ${\text{Pin}}^-$-structure, describing the action of $H^1(M;{\mathbb{Z}}_2)$ in terms of the action of a subgroup of $H^1(\Sigma;{\mathbb{Z}}_2)$ on the quadratic enhancements on the non-singular fibre Σ of the Lefschetz fibration. We refer the reader to [Reference Akbulut1, Section 1.5] for the conventions and backgound on Kirby diagrams of non-orientable 4-manifolds, see also [Reference Akbulut, Gordon and Kirby2], [Reference Bais and Torres4, Section 2.1], [Reference César de Sá6] and [Reference Miller and Naylor18]. The procedure we describe in this section applies to any closed non-orientable 4-manifold represented by means of a Kirby diagram.

Example 2 ( $\mathbb{R} \mathbb{P}^4$)

Figure 1 represents a Kirby diagram for the standard handlebody decomposition of $\mathbb{R} \mathbb{P}^4$ with a single k-handle for each $k=0, \dots, 4$. In particular, the union of all handles up to index 2 corresponds to a tubular neighbourhood $\nu(\mathbb{R} \mathbb{P}^2) \cong D^2 \mathbin{\widetilde{\smash{\times}}} \mathbb{R} \mathbb{P}^2$ of $\mathbb{R} \mathbb{P}^2$ (see [Reference Akbulut, Gordon and Kirby2, Section 0] and [Reference Bais and Torres4, Section 2.5]) and $\mathbb{R} \mathbb{P}^4$ is the result of attaching a 3-handle and a 4-handle to this handlebody. Recall that 3-handles and 4-handles need not be drawn by [Reference César de Sá6] and [Reference Miller and Naylor18], see [Reference Bais and Torres4, Example 7].

Figure 1. A Kirby diagram of $\mathbb{R} \mathbb{P}^4$.

The union of handles up to index 2 can be seen as a Lefschetz fibration over the 2-disk with the Möbius band Mb as fibre and a single vanishing cycle c 1, given by the projection of the $(1,0)$-framed 2-handle, see Figure 2.

Figure 2. The Lefschetz fibration associated to the 2-handlebody of $\mathbb{R} \mathbb{P}^4$.

A ${\text{Pin}}^+$-structure on Mb extending to $\mathbb{R} \mathbb{P}^4$ is given by the quadratic enhancement

(11)\begin{equation} q^{+}:H_1(\text{Mb};{\mathbb{Z}}_4) \rightarrow {\mathbb{Z}}_2 \end{equation}

defined by the condition $q^{+}([e_1])=1$, where $[e_1]$ is the generator of $H_1(\text{Mb};{\mathbb{Z}}_4) \cong {\mathbb{Z}}_4$ represented by the core of the band. Indeed, $[c_1]=2[e_1]$ and hence

\begin{equation*}q^+([c_1])=q^{+}(2[e_1])= q^{+}([e_1])+q^{+}([e_1])+e_1^2=1 \in {\mathbb{Z}}_2.\end{equation*}

The other ${\text{Pin}}^+$-structure

\begin{equation*}q^{+}_x(y)=q^{+}(y)+x \cdot y \quad \text{for all } y \in H_1(\text{Mb}; {\mathbb{Z}}_4)\end{equation*}

is obtained by acting on $q^{+}$ by the cohomology class dual to $[e_1]$, where $x \cdot y$ indicates the intersection number between x and the ${\mathbb{Z}}_2$-reduction of y. Note that the condition $q^{+}_x([c_1])=1$ is still satisfied.

On the other hand, if $\mathbb{R} \mathbb{P}^4$ supported a ${\text{Pin}}^-$-structure, then there would be a quadratic enhancement

\begin{equation*}q^{-}:H_1(\text{Mb};{\mathbb{Z}}_2) \rightarrow {\mathbb{Z}}_4\end{equation*}

satisfying $q^{-}([c_1])=2$. But this is impossible, since it would imply that

\begin{equation*}2=q^{-}([c_1])=q^{-}(2[e_1])=q^{-}(0)=0 \in {\mathbb{Z}}_4.\end{equation*}

Example 3 ( $S^2 \mathbin{\widetilde{\smash{\times}}} \mathbb{R} \mathbb{P}^2$)

The non-orientable 4-manifold $S^2 \mathbin{\widetilde{\smash{\times}}} \mathbb{R} \mathbb{P}^2$ is obtained as a quotient of $S^2\times S^2$ under the orientation-reversing involution

\begin{align*} \iota : S^2 \times S^2 \rightarrow S^2 \times S^2 \\ (x,y) \mapsto (-x, \rho_{\pi}(y)) \end{align*}

where $\rho_{\theta}: S^2 \rightarrow S^2$ is the rotation of S 2 of angle θ about a fixed axis for any $\theta \in S^1$, see [Reference Hillman15, Chapter 12]. Moreover, it can also be seen as the result of a Gluck twist on the product $S^2 \times \mathbb{R} \mathbb{P}^2$ along a 2-sphere fibre, see [Reference Bais and Torres4, Lemma 1]. In particular, there is a decomposition

\begin{equation*} S^2 \mathbin{\widetilde{\smash{\times}}} \mathbb{R} \mathbb{P}^2= (S^2 \times D^2) \cup_{\varphi} (S^2 \times \text{Mb})\end{equation*}

where the gluing map is

\begin{align*} \varphi: S^1 \times S^1 \rightarrow S^2 \times S^2 \\ (x, \theta ) \mapsto (\rho_{\theta}(x), \theta). \end{align*}

A Kirby diagram of this manifold is given in Figure 3, see [Reference Akbulut, Gordon and Kirby2, Section 0], [Reference Bais and Torres4, Figure 1], [Reference Miller and Naylor18] and [Reference Torres25, Section 4.1]. It is obtained by considering $S^2 \mathbin{\widetilde{\smash{\times}}} \mathbb{R} \mathbb{P}^2$ as the double of $D^2 \mathbin{\widetilde{\smash{\times}}} \mathbb{R} \mathbb{P}^2$. Figure 4 represents the projection of the attaching circles of the 2-handles onto the page of the trivial Lefschetz fibration given by the union of handles up to index one. In order to endow the 2-handlebody of $S^2 \mathbin{\widetilde{\smash{\times}}} \mathbb{R} \mathbb{P}^2$ with the structure of a Lefschetz fibration over D 2, we need to stabilize the page and add vanishing cycles as in Figure 5, so that the attaching spheres of the 2-handles can be isotoped into different pages and have the right framing. This is Harer’s trick and is explained in the non-orientable setting in [Reference Miller and Özbağcı19, Proof of Theorem 1.1], see also [Reference Harer14] and [Reference Etnyre and Fuller8, Theorem 2.1]. At the level of Kirby diagrams, this corresponds to adding canceling pairs of 1- and 2-handles.

Figure 3. A Kirby diagram of $S^2 \mathbin{\widetilde{\smash{\times}}} \mathbb{R} \mathbb{P}^2$.

Figure 4. Lefschetz fibration associated to the 2-handlebody of $S^2 \mathbin{\widetilde{\smash{\times}}} \mathbb{R} \mathbb{P}^2$.

Figure 5. The 2-handlebody of $S^2 \mathbin{\widetilde{\smash{\times}}} \mathbb{R} \mathbb{P}^2$ as a Lefschetz fibration over D 2.

In Figure 6 we fix a set of simple closed loops $e_1, \dots, e_6$ inducing a homology basis of the page Σ of this new fibration. The quadratic enhancement

\begin{equation*}q^{+}:H_1(\Sigma;{\mathbb{Z}}_4) \rightarrow {\mathbb{Z}}_2\end{equation*}

Figure 6. Curves representing a basis of $H_1(\Sigma).$

defined by the condition $q^{+}([e_i])=1$ for every $i=1, \dots, 6$ has the property that $q^{+}([c_i])=1$ for all the vanishing cycles $c_1, \dots, c_7$ and hence it determines a ${\text{Pin}}^+$-structure on $S^2 \mathbin{\widetilde{\smash{\times}}} \mathbb{R} \mathbb{P}^2$. The other ${\text{Pin}}^+$-structure can be found by acting on $q^{+}$ with $[e_1]\in H_1(\Sigma;{\mathbb{Z}}_2)$.

On the other hand, $S^2 \mathbin{\widetilde{\smash{\times}}} \mathbb{R} \mathbb{P}^2$ does not support a ${\text{Pin}}^-$-structure. Indeed, if this was the case, we could find a quadratic enhancement

\begin{equation*}q^{-}: H_1(\Sigma;{\mathbb{Z}}_2) \rightarrow {\mathbb{Z}}_4\end{equation*}

such that $q^{-}([e_i])=2$ for all $i=1, \dots, 7$ and this would imply that

\begin{align*}2&=q^{-}([c_1])=q^{-}([e_2+e_6])=q^{-}([e_2])+q^{-}([e_6])=q^{-}([c_2])+q^{-}([c_6])=2+2\\ &=0\in {\mathbb{Z}}_4\end{align*}

a contradiction.

Example 4 ( $S^2 \times \mathbb{R} \mathbb{P}^2$)

A Kirby diagram of the product 4-manifold $S^2 \times \mathbb{R} \mathbb{P}^2$ is given in Figure 7, see [Reference Akbulut, Gordon and Kirby2, Section 0], [Reference Bais and Torres4, Figure 1], [Reference Miller and Naylor18] and [Reference Torres25, Section 4.1]. It is obtained by considering $S^2 \times \mathbb{R} \mathbb{P}^2$ as the double of $D^2 \times \mathbb{R} \mathbb{P}^2$. Figure 8 shows the projection of the attaching circles of the 2-handles onto the page of the trivial Lefschetz fibration given by the union of handles up to index one.

Figure 7. A Kirby diagram of $S^2 \times \mathbb{R} \mathbb{P}^2$.

Figure 8. Lefschetz fibration associated to the 2-handlebody of $S^2 \times \mathbb{R} \mathbb{P}^2$.

We use Harer’s trick again (see [Reference Miller and Özbağcı19, Proof of Theorem 1.1], [Reference Harer14] and [Reference Etnyre and Fuller8, Theorem 2.1]) and show that the union of the handles up to index 2 in the fixed decomposition of $S^2 \times \mathbb{R} \mathbb{P}^2$ is given by the surface Σ and by the vanishing cycles $c_1, \dots, c_8$ in Figure 9. Note that, with respect to the case of $S^2 \widetilde \times \mathbb{R} \mathbb{P}^2$, we need to stabilize the page one additional time in order to adjust the framing of the red coloured 2-handle.

Figure 9. The 2-handlebody of $S^2 \times \mathbb{R} \mathbb{P}^2$ as a Lefschetz fibration over D 2.

In Figure 10 we again fix a set of loops $e_1, \dots, e_7 \subset \Sigma$ inducing a first homology basis of the page.

Figure 10. Curves representing a basis of $H_1(\Sigma)$.

The quadratic enhancement

\begin{equation*}q^{-}:H_1(\Sigma;{\mathbb{Z}}_2) \rightarrow {\mathbb{Z}}_4\end{equation*}

defined by the condition $q^{-}([e_i])=2$ for every $i=1, \dots, 7$ has the property that $q^{-}([c_i])=2$ for all the vanishing cycles $c_1, \dots, c_8$ and hence it determines a ${\text{Pin}}^-$-structure on $S^2 \times \mathbb{R} \mathbb{P}^2$. The other ${\text{Pin}}^-$-structure can be found by acting on $q^{-}$ with $[e_1]\in H_1(\Sigma;{\mathbb{Z}}_2)$.

On the other hand, $S^2 \times \mathbb{R} \mathbb{P}^2$ can not support a ${\text{Pin}}^+$-structure. Indeed, if this was the case we could find a quadratic enhancement

\begin{equation*}q^{+}: H_1(\Sigma;{\mathbb{Z}}_4) \rightarrow {\mathbb{Z}}_2\end{equation*}

such that $q^{+}([c_i])=1$ for all $i=1, \dots, 8$ and this would imply that

\begin{align*} 1=q^{+}([c_1])=q^{+}([2e_1+e_2+e_6+e_7])=\\ 2q^{+}([e_1])+e_1^2+q^{+}([e_2])+q^{+}([e_6])+q^{+}([e_7])=\\ 1+q^{+}([c_2])+q^{+}([c_6])+q^{+}([c_7])=1+1+1+1=0 \in {\mathbb{Z}}_2 \end{align*}

a contradiction.

7. ${\text{Pin}}^{\pm}$-structures on Lefschetz fibrations over S 2

In this section, we provide a proof of Theorem 4.

Proof. Proof of Theorem 4 The proof is a ${\mathbb{Z}}_2$-coefficients version of the one of [Reference Stipsicz23, Theorem 1.3] using the two formulas (4) and (3) in Section 2. We hence leave it to the interested reader as an exercise.

Remark 3. One can check whether $X \setminus \nu(\Sigma)$ supports a ${\text{Pin}}^{\pm}$-structure by applying Theorem 1 and Theorem 2 respectively.

8. ${\text{Pin}}^+$ and ${\text{Pin}}^-$-structures on 3-manifolds

The following well-known fact can be easily proven by adapting to the ${\text{Pin}}^-$ case the one of [Reference Stipsicz23, Theorem 1.4], see [Reference Kirby and Taylor17].

Theorem 6. Any closed 3-manifold M admits a $Pin^-$-structure.

Sketch of Proof

We reduce to the case in which M is non-orientable, since if M is orientable then it is automatically Spin and hence also ${\text{Pin}}^+$ and ${\text{Pin}}^-$. The idea is to write M as the union of a non-orientable 3-dimensional handlebody H of genus g with g 2-handles and a single 3-handle. Then [Reference Kirby and Taylor17, Corollary 1.12] implies that M admits a ${\text{Pin}}^-$-structure if and only if the closed surface $\partial H$ supports a ${\text{Pin}}^-$-structure which restricts to the one bounding the 2-disk on all the attaching spheres of the 2-handles and on all the belt spheres of the 1-handles. Such a ${\text{Pin}}^-$-structure can always be constructed by means of a quadratic enhancement $q^{-}: H_1(\partial H;{\mathbb{Z}}_2) \rightarrow {\mathbb{Z}}_4$ which vanishes on all such curves. The proof follows the same lines of the one of [Reference Stipsicz23, Theorem 1.4] and we leave the details to the interested reader.

On the other hand, it is not true that any closed 3-manifold admits a ${\text{Pin}}^+$-structure, since the condition $w_2(M)=0$ is not always satisfied in the non-orientable setting. $\mathbb{R} \mathbb{P}^2 \times S^1 \operatorname{\#} N$ for any 3-manifold N is such an example. However, we now show that it is still possible to check this condition by considering a handlebody decomposition of M.

Theorem 7. Let $M=H \cup H'$ be a closed non-orientable 3-manifold obtained by adding g 2-handles and a 3-handle to a genus-g non-orientable handlebody H and let $q_0^+: H_1(\partial H;{\mathbb{Z}}_4) \rightarrow {\mathbb{Z}}_2$ be a quadratic enhancement corresponding to a ${\text{Pin}}^+$ structure on $\partial H$. Let $a_1, \dots, a_g \in H_1(\partial H;{\mathbb{Z}}_4)$ and $b_1, \dots, b_g \in H_1(\partial H;{\mathbb{Z}}_4)$ be the homology classes of the attaching circles of the 2-handles and of the belt circles of the 1-handles of H respectively and set

\begin{equation*}q^{+}_0([a_j])=\alpha_j \quad \text{and} \quad q^{+}_0([b_j])=\beta_j\end{equation*}

for $j=1, \dots, g$. If for a fixed basis $e_1, \dots, e_{2g}$ of $H_1(\partial H;{\mathbb{Z}}_2)$ we have

\begin{equation*}\tilde a_j=\sum_{i=1}^{2g} a_{j,i} e_i \quad \text{and} \quad \tilde b_j=\sum_{i=1}^{2g} b_{j,i} e_i\end{equation*}

for $j=1, \dots, g$, where $\tilde a_j$ and $\tilde b_j$ are the ${\mathbb{Z}}_2$-reduction of aj and bj respectively, then M has a ${\text{Pin}}^+$-structure if and only if

\begin{equation*}\text{rank}(C|A)=\text{rank} (C)\end{equation*}

where C and A are respectively the $2g \times 2g$ matrix and the column vector

(12)\begin{equation} C= \begin{pmatrix} a_{1,1} & \dots & a_{1,2g}\\ \vdots & \ddots & \vdots\\ a_{g,1} & \dots & a_{g,2g} \\ b_{1,1} & \dots & b_{1,2g}\\ \vdots & \ddots & \vdots \\ b_{2g,1} & \dots & b_{g,2g} \end{pmatrix} , \qquad A= \begin{pmatrix} \alpha_1\\ \vdots\\ \alpha_g\\ \beta_1\\ \vdots\\ \beta_g \end{pmatrix} . \end{equation}

Proof. There is a ${\text{Pin}}^+$-structure on M if and only if the non-orientable closed surface $\partial H$ has a ${\text{Pin}}^+$-structure which extends to the attaching circles of the 2-handles and to the belt spheres of the 1-handles in H. In particular, this is equivalent to check whether or not there exists a map

(13)\begin{equation} q^{+}: H_1(\partial H; {\mathbb{Z}}_4) \rightarrow {\mathbb{Z}}_2 \end{equation}

satisfying the condition $q^+(x+y)=q^+(x)+q^+(y)+x \cdot y$ for all $x,y \in H_1(\partial H;{\mathbb{Z}}_4)$ and which vanishes on $a_1, \dots, a_g$ and on $b_1, \dots, b_g$.

If we fix a quadratic enhancement

(14)\begin{equation} q^{+}_0: H_1(\partial H; {\mathbb{Z}}_4) \rightarrow {\mathbb{Z}}_2 \end{equation}

corresponding to a ${\text{Pin}}^+$-structure on $\partial H$, then all the other ${\text{Pin}}^+$-structures on $\partial H$ will be of the form $q^{+}=q^{+}_0+l$ for some linear map $l \in \text{Hom}(H_1(\partial H;{\mathbb{Z}}_4), {\mathbb{Z}}_2)$. It follows that M supports a ${\text{Pin}}^+$-structure if and only if there is $l=\sum_{i=1}^{2g} x_i e^i\in \text{Hom}(H_1(\partial H;{\mathbb{Z}}_4),{\mathbb{Z}}_2)$ such that

\begin{equation*}\sum_{i=1}^{2g} a_{j,i}x_i=\alpha_j \quad \text{and} \quad \sum_{i=1}^{2g}b_{j,i}x_i=\beta_j\end{equation*}

in ${\mathbb{Z}}_2$ for $j=1, \dots, g$, where $e^1, \dots, e^{2g}$ is the dual basis to $e_1, \dots, e_{2g} \in H_1(\partial H; {\mathbb{Z}}_2)$. The conclusion follows from Rouché–Capelli’s theorem.

9. Interpretation of ${\text{Pin}}^+$-structures à la Milnor

Kirby and Taylor showed in [Reference Kirby and Taylor17] that the sets of ${\text{Pin}}^+$ and ${\text{Pin}}^-$-structures on a fixed vector bundle ξ are in one to one correspondence with Spin-structures on $\xi \oplus 3 \cdot \text{det}(\xi)$ and $\xi \oplus \text{det}(\xi)$ respectively. In this section, we remark that for ${\text{Pin}}^+$-structures one can get another equivalent characterization, which recalls the following one of Spin-structures by Milnor.

Theorem 8 (Milnor [Reference Milnor20])

There is a canonical bijection between the set of Spin-structures on a real rank-k vector bundle $\xi: E \rightarrow M$ and the set of homotopy classes of trivializations of $\xi|_{M^1}$ extending to trivializations of $\xi|_{M^2}$, where Mi denotes the ith skeleton of M. Moreover, such correspondence is equivariant with respect to the action of $H^1(M;{\mathbb{Z}}_2)$.

Recall that $x\in H^1(M;{\mathbb{Z}}_2)$ acts on a trivialization of $\xi|_{M^1}$ by flipping the framing of any loop $\gamma \subset M^1$ such that $\langle x, [\gamma] \rangle =1 \in {\mathbb{Z}}_2$.

In a similar fashion, we prove the following result. We remark that this follows essentially from the discussion in [Reference Kirby and Taylor17].

Theorem 9. Let $\xi: E \rightarrow M$ be a real rank-k vector bundle. There is a canonical bijection between the set of ${\text{Pin}}^+$-structures on ξ and the set of homotopy classes of $(k-1)$-tuples of everywhere linearly independent sections of $\xi|_{M^1}$ extending to $\xi|_{M^2}$, where Mi denotes the ith skeleton of M. Moreover, such correspondence is equivariant with respect to the action of $H^1(M;{\mathbb{Z}}_2)$.

We remark that the definition of the action of $H^1(M;{\mathbb{Z}}_2)$ on the set of homotopy classes of linearly independent sections of $\xi|_{M^1}$ is completely analogous to the one described after the statement of Theorem 8.

Proof. The vanishing of $w_2(\xi)$ is a necessary and sufficient condition for the existence of both a $(k-1)$-tuple of linearly independent sections of $\xi|_{M^2}$ and a ${\text{Pin}}^+$-structure on ξ. We now suppose that $w_2(\xi)$ is trivial and we endow $\text{det}(\xi)$ with its canonical ${\text{Pin}}^+$-structure, see [Reference Kirby and Taylor17, Addendum to 1.2]. Let $v_1, \dots, v_{k-1}$ be a $(k-1)$-tuple of linearly independent sections of $\xi|_{M^1}$ extending to $\xi|_{M^2}$. We are now going to associate to such a $(k-1)$-tuple a ${\text{Pin}}^+$-structure on $\xi|_{M^2}$. This will uniquely determine a ${\text{Pin}}^+$-structure on ξ, as in the case of Spin-structures. Using $v_1, \dots, v_{k-1}$ one can define a vector bundle isomorphism

(15)\begin{equation} \xi|_{M^2} \cong \varepsilon^{k-1} \oplus \text{det}(\xi) \end{equation}

where $\varepsilon^{k-1}$ denotes the trivial rank- $(k-1)$ real vector bundle. The associated ${\text{Pin}}^+$-structure on $\xi|_{M^2}$ is given by the pull-back via (15) of the ${\text{Pin}}^+$-structure on $\varepsilon^{k-1} \oplus \text{det}(\xi)$ given by the direct sum of the trivial Spin-structure on $\varepsilon^{k-1}$ with the fixed ${\text{Pin}}^+$-structure on $\text{det}(\xi)$. It is now easy to verify that this correspondence satisfies all the desired properties.

Acknowledgements

I am sincerely grateful to my advisors Rafael Torres and Daniele Zuddas for the constant support and for their comments on the first versions of this manuscript. The author has been partially supported by GNSAGA—Istituto Nazionale di Alta Matematica ‘Francesco Severi’, Italy.

References

Akbulut, S.. 4-manifolds, Oxford Graduate Texts in Mathematics, 25. vii+263, (Oxford University Press, 2016).Google Scholar
Akbulut, S., A fake 4-manifold, in Four-manifold theory (Gordon, C. and Kirby, R., eds.), Contemp. Math., vol. 35, American Mathematical Society, Providence, RI, 1984, pp. 75141.10.1090/conm/035/780576CrossRefGoogle Scholar
Anderson, D. W., Brown, E. H. and Peterson, F. P.. Pin cobordism and related topics. Comm. Math. Helv. 44 (1969), 462468.10.1007/BF02564545CrossRefGoogle Scholar
Bais, V. and Torres, R.. Smooth structures on non-orientable 4-manifolds via twisting operations a. 2023, rXiv preprint arXiv: 2308.04227.Google Scholar
Cappell, S. E. and Shaneson, J. L.. Some new four-manifolds. Ann. of Math. (2). 104 (1976), 6172.10.2307/1971056CrossRefGoogle Scholar
César de Sá, E.. Automorphisms of 3-manifolds and representations of 4-manifolds, PhD Thesis, University of Warwick (1977).Google Scholar
Degtyarev, A. and Finashin, S.. Pin-structures on surfaces and quadratic forms. Turkish J. Math. 21 (1997), 187193.Google Scholar
Etnyre, J. B. and Fuller, T.. Realizing 4-manifolds as achiral Lefschetz fibrations. Int. Math. Res. Not. 2006 Art. ID 70272, 21 pp.Google Scholar
Fintushel, R. and Stern, R. J.. An exotic free involution on S 4. Ann. of Math. 113 (1981), 357.10.2307/2006987CrossRefGoogle Scholar
Giambalvo, V.. Pin and Pin’ cobordism. Proc. Amer. Math. Soc. 39 (1973), 365.Google Scholar
Gilkey, P. B.. The eta invariant for even-dimensional $\rm{PIN}_{\rm{c}}$ manifolds. Adv. in Math. 58 (1985), 243284.10.1016/0001-8708(85)90119-7CrossRefGoogle Scholar
Gompf, R. E. and Stipsicz, A. I., 4-manifolds and Kirby calculus, Graduate Studies in Mathematics, vol. 20, American Mathematical Society, Providence, RI, 1999. xv+557 pp.Google Scholar
Hambleton, I, Kreck, M. and Teichner, P.. Nonorientable 4-manifolds with fundamental group of order 2. Trans. Amer. Math. Soc. 344 (1994), 649665.Google Scholar
Harer, J. L., Pencils of curves on 4-manifolds, Ph.D. thesis, University of California, California, 1979.Google Scholar
Hillman, J., Four-manifolds, geometries and knots, Geom. Topol. Monogr., vol. 5, Geom. Topol. Publ., Coventry, 2002. xiv+379 pp.Google Scholar
Johnson, D. L.. Spin structures and quadratic forms on surfaces. J. London Math. Soc. (2). 22 (1980), 365373.10.1112/jlms/s2-22.2.365CrossRefGoogle Scholar
Kirby, R. C. and Taylor, L. R., Pin structures on low-dimensional manifolds, in London Mathematical Society Lecture Note Series, vol. 151, Cambridge University Press, Cambridge, 1989, pp. 177242.Google Scholar
Miller, M. H. and Naylor, P.. Trisections of nonorientable 4-manifolds. Michigan Math. J. 74 (2024), 403447.10.1307/mmj/20216127CrossRefGoogle Scholar
Miller, M. and Özbağcı, B.. Lefschetz fibrations on nonorientable 4-manifolds. Pacific Journal of Mathematics. 312 (2021), 177202.CrossRefGoogle Scholar
Milnor, J. W.. Spin structures on manifolds. Enseign. Math. (2). 9 (1963), 198203.Google Scholar
Olędzki, W. J.. On the eta-invariant of ${{\rm} Pin}^+$-operator on some exotic 4-dimensional projective space. Compositio Math. 78 (1991), 127.Google Scholar
Shafarevich, R. I and Remizov, A. O.. Linear Algebra and Geometry. (Springer, Berlin, 2013).CrossRefGoogle Scholar
Stipsicz, A.I. Spin structures on Lefschetz fibrations, Bull. London Math. Soc. 33 (2001), 466472.10.1017/S0024609301008232CrossRefGoogle Scholar
Stolz, S.. Exotic structures on 4-manifolds detected by spectral invariants. Invent. Math. 94 (1988), 147162.10.1007/BF01394348CrossRefGoogle Scholar
Torres, R.. Smooth structures on nonorientable four-manifolds and free involutions. J. Knot Theory Ramifications. 26 (2017), 20, 1750085.10.1142/S0218216517500857CrossRefGoogle Scholar
Figure 0

Figure 1. A Kirby diagram of $\mathbb{R} \mathbb{P}^4$.

Figure 1

Figure 2. The Lefschetz fibration associated to the 2-handlebody of $\mathbb{R} \mathbb{P}^4$.

Figure 2

Figure 3. A Kirby diagram of $S^2 \mathbin{\widetilde{\smash{\times}}} \mathbb{R} \mathbb{P}^2$.

Figure 3

Figure 4. Lefschetz fibration associated to the 2-handlebody of $S^2 \mathbin{\widetilde{\smash{\times}}} \mathbb{R} \mathbb{P}^2$.

Figure 4

Figure 5. The 2-handlebody of $S^2 \mathbin{\widetilde{\smash{\times}}} \mathbb{R} \mathbb{P}^2$ as a Lefschetz fibration over D2.

Figure 5

Figure 6. Curves representing a basis of $H_1(\Sigma).$

Figure 6

Figure 7. A Kirby diagram of $S^2 \times \mathbb{R} \mathbb{P}^2$.

Figure 7

Figure 8. Lefschetz fibration associated to the 2-handlebody of $S^2 \times \mathbb{R} \mathbb{P}^2$.

Figure 8

Figure 9. The 2-handlebody of $S^2 \times \mathbb{R} \mathbb{P}^2$ as a Lefschetz fibration over D2.

Figure 9

Figure 10. Curves representing a basis of $H_1(\Sigma)$.