1 Introduction

Let G be a compact Lie group and BG be the classifying space of G. For a G-space X, the G-equivariant cohomology of X with coefficients in a field \(\textbf{k}\) is the singular cohomology of the Borel construction \(X_G\) (see [6])

$$\begin{aligned} H^*_G(X;\textbf{k}):=H^*(X_G;\textbf{k}). \end{aligned}$$

There is a natural fibration \(X\rightarrow X_G \rightarrow BG\) associated with \(X_G\) called the Borel fibration. If the inclusion of the fiber \(\iota _X: X\rightarrow X_G\) induces a surjection on cohomology \(\iota _X^*: H^*_G(X;\textbf{k})\rightarrow H^*(X;\textbf{k})\), X is called (cohomologically) equivariantly formal over \(\textbf{k}\). This term was coined in 1998 in Goresky-Kottwitz-MacPherson [18]. But this condition had already been studied by A. Borel in  [5, § 4] and [6, Ch. XII] where X is called totally non-homologous to zero in \(X_G\) (also, see [7, Ch. VII]).

For some special groups G shown below, the equivariant formality of a G-action can be interpreted in some other ways (see [5, § 4], [1, Ch. 3], and [2, Sec. 4]).

  • When BG is simply connected (e.g., G is a torus \(T^r=(S^1)^r\)), X is equivariantly formal if and only if the Serre spectral sequence of the Borel fibration of X degenerates at the \(E_2\) stage.

  • When G is the p-torus \(\mathbb Z^r_p\) (p is prime), X being equivariantly formal is equivalent to either one of the following conditions.

    1. (i)

      The Serre spectral sequence with \(\mathbb Z_p\)-coefficients of the Borel fibration of X degenerates at the \(E_2\) stage and the induced action of \(\mathbb Z^r_p\) on \(H^*(X;\mathbb Z_p)\) is trivial.

    2. (ii)

      \(H^*_{\mathbb Z^r_p}(X;\mathbb Z_p)\cong H^*(X;\mathbb Z_p)\otimes H^*(B\mathbb Z^r_p;\mathbb Z_p)\) is a free \(H^*(B\mathbb Z^r_p;\mathbb Z_p)\)-module.

Due to the above fact, we call a \(\mathbb Z^r_p\)-action on X weakly equivariantly formal if we only assume that the Serre spectral sequence (with \(\mathbb Z_p\)-coefficients) of the Borel fibration of X degenerates at the \(E_2\) stage. So an equivariantly formal \(\mathbb Z^r_p\)-action is always weakly equivariantly formal.

When \(G=T^r\) or \(\mathbb Z^r_2\) and \(\textbf{k}=\mathbb Q\) or \(\mathbb Z_2\) respectively, there is another equivalent description of equivariantly formal G-actions given by the so called “Atiyah-Bredon sequence” (see Bredon [8] and Franz-Puppe [16] for the \(T^r\) case, and Allday-Franz-Puppe [2] for the \(\mathbb Z^r_2\) case). In addition, there are many sufficient conditions for a \(T^r\)-action to be equivariantly formal (for example: vanishing of odd-degree cohomology, all homology classes being representable by \(T^r\)-invariant cycles, etc.).

Equivariantly formal G-spaces provide many nice examples in geometry and topology. Some of them are as follows:

  • Smooth compact toric varieties.

  • Hamiltonian G-actions on symplectic manifolds which have moment maps (see Atiyah-Bott [3] and Jeffrey-Kirwan [22]).

  • Quasitoric manifolds and small covers defined in Davis-Januszkiewicz [14].

  • Torus manifolds with vanishing odd degree cohomology (see Masuda-Panov [27]).

In addition, when \(G=T^r\) or \((\mathbb Z_p)^r\), the following theorem gives us an easy way to recognize equivariantly formal G-actions.

Theorem 1.1

(see Theorem (3.10.4) in Allday-Puppe [1]) Let \(G=T^r\) or \((\mathbb Z_p)^r\) where p is a prime and \(\textbf{k}=\mathbb Q\) or \(\mathbb Z_p\) respectively. Let X be a paracompact G-space with only finitely many orbit types and \(\dim _{\textbf{k}} H^*(X;\textbf{k})<\infty \). Then, the fixed point set \(X^G\) always satisfies

$$\begin{aligned} \dim _{\textbf{k}} H^*(X^G;\textbf{k}) \le \dim _{\textbf{k}} H^*(X;\textbf{k}) \end{aligned}$$

where the equality holds if and only if X is equivariantly formal over \(\textbf{k}\). Here \(\dim _{\textbf{k}} H^*(X;\textbf{k})\) denotes the sum of the rank of the cohomology groups of X in all dimensions over \(\textbf{k}\).

A very special case is when \(G=\mathbb Z_2\) and \(X^{\mathbb Z_2}\) consists only of isolated points. By Theorem 1.1, we have

$$\begin{aligned} |X^{\mathbb Z_2}| = \dim _{\mathbb Z_2}H^*(X^{\mathbb Z_2};\mathbb Z_2) \le \dim _{\mathbb Z_2} H^*(X;\mathbb Z_2) \end{aligned}$$
(1)

Such a \(\mathbb Z_2\)-action on X is equivariantly formal if and only if the number of the fixed points reaches the maximum, i.e., \(|X^{\mathbb Z_2}| = \dim _{\mathbb Z_2}H^*(X;\mathbb Z_2)\). In this case, the involution determined by the \(\mathbb Z_2\)-action is called an \(\textrm{m}\)-involution on X (this term was named by Puppe [28]).

There is an interesting relation between \(\textrm{m}\)-involutions on closed manifolds and binary codes. It was shown in [28] that one can obtain a self-dual binary code from any \(\textrm{m}\)-involution on an odd-dimensional closed manifold. This motivates the study in Chen-Lü-Yu [12] on the \(\textrm{m}\)-involutions on a special kind of closed manifolds called small covers (see [14]). In this paper, we want to study a more general type of closed manifolds with 2-torus actions defined below.

Definition 1.2

(see Lü-Masuda [25]) A 2-torus manifold is a closed connected smooth n-manifold W with a non-free effective smooth action of \(\mathbb Z^n_2\). For such a manifold W, since \(\dim (W)=n=\textrm{rank}(\mathbb Z^n_2)\) and the \(\mathbb Z^n_2\)-action is effective, the fixed point set \(W^{\mathbb Z^n_2}\) must be discrete. Then, since W is compact, \(W^{\mathbb Z^n_2}\) is a finite set of isolated points (if not empty). Note that we require all 2-torus manifolds to be connected in this paper.

  • For brevity, we call a 2-torus manifold W equivariantly formal or weakly equivariantly formal if the \(\mathbb Z^n_2\)-action on W is so, respectively.

  • We call W locally standard if for every point \(x \in W\), there is a \(\mathbb Z^n_2\)-invariant neighborhood \(V_x\) of x such that \(V_x\) is equivariantly homeomorphic to an invariant open subset of a real n-dimensional faithful linear representation space of \(\mathbb Z^n_2\). An equivalently way to describe such a neighborhood \(V_x\) is: \(V_x\) is weakly equivariantly homeomorphic to an invariant open subset of \(\mathbb R^n\) under the standard \(\mathbb Z^n_2\)-action defined by: for any \((x_1,\cdots , x_n)\in \mathbb R^n\) and \((g_1,\cdots , g_n)\in \mathbb Z^n_2\),

    $$\begin{aligned} \quad \ (g_1,\cdots , g_n) \cdot (x_1,\cdots , x_n)\longmapsto \big ((-1)^{g_1}x_1,\cdots , (-1)^{g_n}x_n \big ). \end{aligned}$$
  • Every non-zero element \(g\in \mathbb Z^n_2\) determines a nontrivial involution \(\tau _g\) on W, called a regular involution on W.

We will prove in Theorem 3.3 that if a 2-torus manifold is equivariantly formal, then it must be locally standard.

For an n-dimensional locally standard 2-torus manifold W, the orbit space \(Q\!=\!W/\mathbb Z^n_2\) naturally becomes a connected smooth nice n-manifold with corners and with non-empty boundary (since the \(\mathbb Z^n_2\)-action is non-free). Moreover,

  • The \(\mathbb Z^n_2\)-action on W determines a characteristic function

    $$\begin{aligned} \lambda _W: \{F_1,\cdots , F_m\}\rightarrow \mathbb Z^n_2 \end{aligned}$$

    where \(F_1,\cdots , F_m\) are all the facets (codimension-one faces) of Q.

  • The free part of the \(\mathbb Z^n_2\)-action on W determines a principal \(\mathbb Z^n_2\)-bundle \(\xi _W\) over Q.

It is shown in Lü-Masuda [25, Lemma 3.1] that W can be recovered from the data \((Q,\lambda _W, \xi _W)\) up to equivariant homeomorphism. In addition, let \(\pi : W\rightarrow Q\) denote the orbit map. If f is a codimension-k face of Q, then \(W_f:=\pi ^{-1}(f)\) is a codimension-k submanifold of W called a facial submanifold of W. Let \(G_f\) denote the isotropy subgroup of \(W_f\). Then, \(W_f\) is also a 2-torus manifold with respect to the induced action of \(\mathbb Z^n_2/G_f\). In the following, when we say \(W_f\) is equivariantly formal, we always consider \(W_f\) being equipped with the induced \(\mathbb Z^n_2/G_f\)-action from W.

The main purpose of this paper is to answer the following two questions:

Question-1: What kind of 2-torus manifolds are equivariantly formal?

Question-2: What kind of locally standard 2-torus manifolds have regular \(\textrm{m}\)-involutions?

Generally speaking, it is very hard to compute the equivariant cohomology of a locally standard 2-torus manifold W directly from its orbit space Q and the data \((\lambda _W, \xi _W)\). So it is difficult to judge whether W is equivariantly formal by directly verifying the condition in the definition. Meanwhile, it was proved by Masuda-Panov [27] that a smooth \(T^n\)-action on a connected smooth 2n-manifold with non-empty fixed points is equivariantly formal if and only if the \(T^n\)-action is locally standard and every face of its orbit space is acyclic (also see Goertsches-Töben [19, Theorem 10.19] for a reformulation of this result). This result is also implied by Franz [15, Theorem 1.3]. The arguments in [27] inspire us to prove the following parallel result for 2-torus manifolds.

Theorem 1.3

Let W be a 2-torus manifold with orbit space Q.

  1. (i)

    W is equivariantly formal if and only if W is locally standard and Q is mod 2 face-acyclic.

  2. (ii)

    W is equivariantly formal and \(H^*(W;\mathbb Z_2)\) is generated by its degree-one part as a ring if and only if W is locally standard and Q is a mod 2 homology polytope.

The definitions of “mod 2 face-acyclic” and “mod 2 homology polytope” are given in Definition 2.1.

The main strategy in our proof of Theorem 1.3 is very similar to the strategy used in [27] for equivariantly formal torus manifolds. Besides, our proof uses the mod 2 GKM theory introduced in Biss-Guillemin-Holm [4] which allows us to observe the equivariant cohomology of an equivariantly formal 2-torus manifold by restricting to its fixed point set (see Sect. 2.3).

Remark 1.4

If a 2-torus manifold W is assumed to be locally standard in the first place, Theorem 1.3 (i) can also be derived from Chaves [11, Theorem 1.1] whose proof uses the theory of syzygies in the mod 2 equivariant cohomology (see Allday-Franz-Puppe [2, Theorem 10.2]) and the mod 2 “Atiyah-Bredon sequence”. But we will use a completely different approach in our proof here.

Using Theorem 1.3, we can easily derive the following theorem which gives an answer to Question-2.

Theorem 1.5

Let W be an n-dimensional locally standard 2-torus manifold with orbit space Q. Then, there exists a regular \(\textrm{m}\)-involution on W if and only if Q is mod 2 face-acyclic (or equivalently W is equivariantly formal) and the values of the characteristic function \(\lambda _W\) on all the facets of Q consist exactly of a linear basis of \(\mathbb Z^n_2\).

A nice manifold with corners Q is called k-colorable if we can assign k different colors to all the facets of Q so that no two adjacent facets are of the same color. Clearly, there exists a 2-torus manifold over Q whose characteristic function takes value in a linear basis of \(\mathbb Z^n_2\) if and only if Q is n-colorable.

Remark 1.6

By Theorem 1.5 and the construction in Puppe [28], we can obtain a self-dual binary code \(\mathcal {C}_Q\) from an n-colorable mod 2 face-acyclic nice smooth n-manifold with corners Q when n is odd. This generalizes the self-dual binary codes from n-colorable simple convex n-polytopes in Chen-Lü-Yu [12]. Moreover, we can write down \(\mathcal {C}_Q\) explicitly in the same way as the self-dual binary code obtained in [12, Corollary 4.5].

The paper is organized as follows. In Sect. 2, we review the definitions and some basic facts of locally standard 2-torus manifolds and quote some well known results that are useful for our proof. In Sect. 3, we study various properties of equivariantly formal 2-torus manifolds. Since the philosophy of our study is very similar to the study of torus manifolds with vanishing odd degree cohomology in Masuda-Panov [27], many lemmas in this paper are parallel to those in [27]. In Sect. 4, we prove some special properties of equivariantly formal 2-torus manifolds whose mod 2 cohomology rings are generated by their degree-one part. Then finally, in Sect. 5, we prove Theorem 1.3 and Theorem 1.5.

2 Preliminaries

2.1 Manifolds with Corners and Locally Standard 2-Torus Manifolds

Recall a (smooth) n-dimensional manifold with corners Q is a Hausdorff space together with a maximal atlas of local charts onto open subsets of \(\mathbb R_{\ge 0}^n \) such that the transition functions are (diffeomorphisms) homeomorphisms which preserve the codimension of each point. Here, the codimension c(x) of a point \(x=(x_1,\cdots ,x_n)\) in \(\mathbb R_{\ge 0}^n\) is the number of \(x_i\) that is 0. So we have a well defined map \(c: Q\rightarrow \mathbb Z_{\ge 0}\) where c(q) is the codimension of a point \(q\in Q\). An open face of Q of codimension k is a connected component of \(c^{-1}(k)\). A (closed) face is the closure of an open face. A face of codimension one is called a facet of Q. When Q is connected, we also consider Q itself as a face (of codimension zero).

  • For any \(k\in \mathbb Z_{\ge 0}\), the k-skeleton of Q is the union of all the faces of Q with dimension \(\le k\).

  • The face poset of Q, denoted by \(\mathcal {P}_Q\), is the set of faces of Q ordered by reversed inclusion (so Q is the initial element).

A manifold with corners Q is said to be nice if either its boundary \(\partial Q\) is empty or \(\partial Q\) is non-empty and any codimension-k face of Q is a component of the intersection of k different facets in Q. If Q is nice, \(\mathcal {P}_Q\) is a simplicial poset. But in general \(\mathcal {P}_Q\) may not be the face poset of a simplicial complex. Indeed, \(\mathcal {P}_Q\) is the face poset of a simplicial complex if and only if all non-empty multiple intersections of facets of Q are connected (see [27, Sec. 5.2]).

Definition 1.1

Let Q be a nice manifold with corners.

  • We call Q mod 2 face-acyclic if every face of Q (including Q itself) is a mod 2 acyclic space.

  • We call Q a mod 2 homology polytope if Q is mod 2 face-acyclic and \(\mathcal {P}_Q\) is the face poset of a simplicial complex.

A topological space B is called mod 2 acyclic if \(H^*(B;\mathbb Z_2)\cong H^*(pt;\mathbb Z_2)\).

It is not difficult to prove the following lemma (see [27, p.743 Remark] for a short argument).

Lemma 1.2

If Q is mod 2 face-acyclic, then every face of Q has a vertex and the 1-skeleton of Q is connected.

In the following, let W be an n-dimensional locally standard 2-torus manifold with orbit space Q. Then, Q is a smooth nice manifold with corners with \(\partial Q eq \varnothing \). Let \(\pi : W\rightarrow Q\) denote the projection, and let the set of facets of Q be

$$\begin{aligned} \mathcal {F}(Q) =\{F_1,\cdots , F_m \}. \end{aligned}$$

Then, \(\pi ^{-1}(F_1),\cdots ,\pi ^{-1}(F_m)\) are embedded codimension-one closed connected submanifolds of W, called the characteristic submanifolds of W. Moreover, the \(\mathbb Z^n_2\)-action on W determines a characteristic function on Q which is a map

$$\begin{aligned} \lambda _W: \mathcal {F}(Q)\rightarrow \mathbb Z^n_2 \end{aligned}$$
(2)

where \(\lambda _W(F_i)\in \mathbb Z^n_2\) is the generator of the \(\mathbb Z_2\) subgroup that pointwise fixes the submanifold \(\pi ^{-1}(F_i)\), \(1\le i \le m\). Since the \(\mathbb Z^n_2\)-action is locally standard, the function \(\lambda _W\) satisfies the following linear independence condition:

$$\begin{aligned} \quad \begin{array}{l} \text {whenever the intersection of }k\text { different facets }F_{i_1},\cdots , F_{i_k}\text { is non-empty,} \\ \text {the elements }\lambda _W(F_{i_1}),\cdots , \lambda _W(F_{i_k})\text { are linearly independent when viewed} \\ \text {as vectors of }\mathbb Z^n_2\text { over the field }\mathbb Z_2. \end{array} \end{aligned}$$

For a codimension-k face f of Q, let \(F_{i_1},\cdots , F_{i_k}\) be all the facets containing f. Then, the isotropy subgroup of the facial submanifold \(W_f\) is

$$\begin{aligned} G_f = \text {the subgroup generated by }\{\lambda _W(F_{i_1}),\cdots , \lambda _W(F_{i_k})\} \subseteq \mathbb Z^n_2. \end{aligned}$$
(3)

By the linear independence condition of \(\lambda _W\), \(G_f\cong \mathbb Z^k_2\). Hence \(W_f\) is also a 2-torus manifold with respect to the induced action of \(\mathbb Z^n_2/G_f \cong \mathbb Z^{n-k}_2\).

In addition, W determines a principal \(\mathbb Z^n_2\)-bundle over Q as follows. We take a small invariant open tubular neighborhood for each characteristic submanifold of W and remove their union from W. Then, the \(\mathbb Z^n_2\)-action on the resulting space is free, and its orbit space can naturally be identified with Q, which gives a principal \(\mathbb Z^n_2\)-bundle over Q, denoted by \(\xi _W\). It is shown in Lü-Masuda [25] that W can be recovered (up to equivariant homeomorphism) from (\(Q, \xi _W,\lambda _W\)). For example, when \(\xi _W\) is a trivial \(\mathbb Z^n_2\)-bundle, W is equivariantly homeomorphic to the following “canonical model” determined by \((Q,\lambda _W)\).

$$\begin{aligned} M_Q(\lambda _W) := Q\times \mathbb Z^n_2 /\sim \end{aligned}$$
(4)

where \((q,g)\sim (q',g')\) if and only if \(q=q'\) and \(g-g'\in G_{f(q)}\) where f(q) is the unique face of Q that contains q in its relative interior. This canonical model is a generalization of a result of Davis-Januszkiewicz [14, Prop. 1.8]. We will see that the canonical model plays an important role in our proof of Theorem 1.3 in Sect. 5.

2.2 Borel Construction and Equivariant Cohomology

For a topological group G, there exists a contractible free right G-space EG called the universal G-space. The quotient \(BG= EG /G\) is called the classifying space for free G-actions. For example, when \(G=\mathbb Z_2^n\), we can choose

$$\begin{aligned} E\mathbb Z^n_2 = (E\mathbb Z_2)^n = (S^{\infty })^n, \ \ B\mathbb Z_2^n = (B\mathbb Z_2)^n = (\mathbb RP^{\infty })^n. \end{aligned}$$

Let X be a topological space with a left G-action (we call X a G-space for brevity). The Borel construction of X is denoted by

$$ EG\times _G X = EG \times X /\sim $$

where \((e,x)\sim (eg,g^{-1}x)\) for any \(e\in EG\), \(x\in X\) and \(g\in G\).

The equivariant cohomology of X with coefficients in a field \(\textbf{k}\) is defined as

$$\begin{aligned} H^*_G(X;\textbf{k}):= H^*(EG\times _G X;\textbf{k}). \end{aligned}$$

Convention: The term “cohomology” of a space in this paper, always mean singular cohomology if not specified otherwise.

The Borel construction determines a canonical fibration called Borel fibration:

$$\begin{aligned} X \rightarrow EG\times _G X \rightarrow BG. \end{aligned}$$
(5)

The map \(\rho \) collapsing X to a point induces a homomorphism

$$\begin{aligned} \rho ^* : H^*_G(pt ;\textbf{k}) = H^*(BG;\textbf{k}) \rightarrow H^*_G(X;\textbf{k}) \end{aligned}$$
(6)

which defines a canonical \(H^*(BG;\textbf{k})\)-module structure on \(H^*_G(X;\textbf{k})\). A useful fact is: when X is a paracompact space with finite cohomology dimension, and \(G=T^r\) or \((\mathbb Z_p)^r\) where p is a prime and \(\textbf{k}=\mathbb Q\) or \(\mathbb Z_p\) respectively, \(\rho ^*\) is injective if and only if the fixed point set \(X^G\) is non-empty (see [21, Ch. IV]).

In general, \(H^*_G(X;\textbf{k})\) may not be a free \(H^*(BG;\textbf{k})\)-module. The following localization theorem due to A. Borel (see [21, p. 45]) says that we can compute the free \(H^*(BG;\textbf{k})\)-module part of \(H^*_G(X;\textbf{k})\) by restricting to the fixed point set.

Theorem 1.3

(Localization Theorem) Let \(G=T^r\) or \((\mathbb Z_p)^r\) where p is a prime and \(\textbf{k}=\mathbb Q\) or \(\mathbb Z_p\) respectively. For a paracompact G-space X with finite cohomology dimension, the following localized restriction homomorphism is an isomorphism:

$$\begin{aligned} S^{-1} H^*_G(X;\textbf{k}) \rightarrow S^{-1} H^*_G(X^G;\textbf{k}) = H^*(X^G;\textbf{k})\otimes _{\textbf{k}} (S^{-1} H^*(BG;\textbf{k})) \end{aligned}$$

where \(S=R-\{0\}\) where R is the polynomial subring of \(H^*(BG;\textbf{k})\). So the kernel of the restriction \(H^*_G(X;\textbf{k}) \rightarrow H^*_G(X^G;\textbf{k})\) lies in the \(H^*(BG;\textbf{k})\)-torsion of \(H^*_G(X;\textbf{k})\). In particular if X is equivariantly formal, \(H^*_G(X;\textbf{k})\rightarrow H^*_G(X^G;\textbf{k})\) is injective.

The Borel construction can also be applied to a G-vector bundle \(\pi : E\rightarrow X\) (i.e., both E and X are G-spaces and the projection \(\pi \) is G-equivariant). In this case, the Borel construction \(E_G\) of E is a vector bundle over \(X_G\) whose mod 2 Euler class, denoted by \(e^G(E)\), lies in \(H^*_G(X;\mathbb Z_2)\). Note that using \(\mathbb Z_2\)-coefficients allows us to ignore the orientation of a vector bundle.

2.3 Mod 2 GKM-Theory

Let W be an n-dimensional equivariantly formal 2-torus manifold. Then, the fixed point set \(W^{\mathbb Z^n_2}\) is a finite non-empty set (by Theorem 1.1), and \(H^*_{\mathbb Z^n_2}(W;\mathbb Z_2)\) is a free module over \(H^*(B\mathbb Z^n_2;\mathbb Z_2)\). Moreover, \(H^*_{\mathbb Z^n_2}(W;\mathbb Z_2)\) can be computed by the so called Mod 2 GKM-theory (see Biss-Guillemin-Holm [4]) which is an extension of the GKM-theory in [18] to 2-torus actions. In this section, we briefly review some results related to our study. The reader is referred to [4] and [24] for more details.

For each \(1\le i \le n\), let \(\rho _i\in \textrm{Hom}(\mathbb Z^n_2,\mathbb Z_2)\) be the homomorphism defined by

$$ \rho _i((g_1,\cdots , g_n))=g_i,\ \forall (g_1,\cdots , g_n)\in \mathbb Z^n_2. $$

By a canonical isomorphism \(\textrm{Hom}(\mathbb Z^n_2,\mathbb Z_2)\cong H^1(B\mathbb Z^n_2;\mathbb Z_2)\), we can identify \(H^*(B\mathbb Z^n_2;\mathbb Z_2)\) with the graded polynomial ring \(\mathbb Z_2[\rho _1,\cdots ,\rho _n]\) where \(\textrm{deg}(\rho _i)=1\), \(1\le i \le n\).

Let \(Q=W/\mathbb Z^n_2\) be the orbit space of W. By our Theorem 3.3 proved later, a 2-torus manifold W being equivariantly formal implies that it is locally standard. Hence Q is a nice manifold with corners. Then, the 1-skeleton of Q, consisting of vertices (0-faces) and edges (1-faces) of Q, is an n-valent graph denoted by \(\Gamma (Q)\). Let V(Q) and E(Q) denote the set of vertices and edges of Q, respectively.

Convention: We will not distinguish a vertex of Q and the corresponding fixed point in \(W^{\mathbb Z^n_2}\) in the rest of the paper.

  • Let \(\pi : W\rightarrow Q\) be the quotient map.

  • For each edge \(e\in E(Q)\), \(\pi ^{-1}(e)\) is a circle whose isotropy subgroup \(G_e\) is a rank \(n-1\) subgroup of \(\mathbb Z^n_2\). Then, we obtain a map

    $$\begin{aligned} \alpha : E(Q) \rightarrow \textrm{Hom}(\mathbb Z^n_2,\mathbb Z_2)\cong H^1(B\mathbb Z^n_2;\mathbb Z_2) \end{aligned}$$

    where for each edge \(e\in E(Q)\), \(\ker (\alpha (e))=G_e\).

  • For each vertex \(p\in V(Q)\), let \(\alpha _p = \{ \alpha (e)\,|\, p\in e \}\subset \textrm{Hom}(\mathbb Z^n_2,\mathbb Z_2) \).

Such a map \(\alpha \) is called an axial function which has the following properties:

  1. (i)

    For every vertex \(p\in V(Q)\), \(\alpha _p\) is a linear basis of \(\textrm{Hom}(\mathbb Z^n_2,\mathbb Z_2)\).

  2. (ii)

    For every edge \(e\in E(Q)\), \(\alpha _p \equiv \alpha _{p'}\) mod \(\alpha (e)\) where \(p,p'\) are the two vertices of e.

By [4, Theorem C] and [4, Remark 5.9], we have the following theorem which is a consequence of the \(\mathbb Z_2\)-version Chang-Skjelbred theorem (see [4, Theorem 4.1] and [10]).

Theorem 1.4

(see [4]) Let W be an n-dimensional equivariantly formal 2-torus manifold. If we choose an element \(\eta _p \in H^*_{\mathbb Z^n_2}(W^{\mathbb Z^n_2};\mathbb Z_2)\) for each \(p\in W^{\mathbb Z^n_2}\), then

$$(\eta _p)\in \bigoplus _{p\in W^{\mathbb Z^n_2}} H^*(B\mathbb Z^n_2;\mathbb Z_2) \cong H^*_{\mathbb Z^n_2}(W^{\mathbb Z^n_2};\mathbb Z_2)$$

is in the image of the restriction homomorphism \(r: H^*_{\mathbb Z^n_2}(W;\mathbb Z_2) \rightarrow H^*_{\mathbb Z^n_2}(W^{\mathbb Z^n_2};\mathbb Z_2)\) if and only if for every edge \(e\in E(Q)\) with vertices p and \(p'\), \(\eta _p - \eta _{p'}\) is divisible by \(\alpha (e)\).

Moreover, we can understand the above axial function \(\alpha \) in the following way. For brevity, we use the following notations for an n-dimensional locally standard 2-torus manifold W in the rest of this section.

  • Let \(G=\mathbb Z^n_2\).

  • Let \(W_i:=W_{F_i}=\pi ^{-1}(F_i)\), \(1\le i \le m\), be all the characteristic submanifolds of W where \(F_1,\cdots , F_m\) are all the facets of Q.

  • Let \(G_i:=\langle \lambda _W(F_i)\rangle \cong \mathbb Z_2\) be subgroup of G that fixes \(W_i\) pointwise.

  • Let \(\nu _i\) be the (equivariant) normal bundle of \(W_i\) in W. So we have the equivariant Euler class of \(\nu _i\), denoted by \(e^G(\nu _i)\in H^1_G(W_i;\mathbb Z_2)\).

  • For any fixed point \(p\in W^{\mathbb Z^n_2}\), let \( I(p):=\{ i \,|\, p\in W_i \}\). We have the decomposition of tangent space \(T_p W\) as

    $$\begin{aligned} T_p W = \bigoplus _{i\in I(p)} \nu _i|_p. \end{aligned}$$

    where \(\nu _i|_p\) denotes the restriction of \(\nu _i\) to p. So \(\nu _i|_p\) is a 1-dimensional linear representation of G whose equivariant Euler class

    $$\begin{aligned} e^G(\nu _i|_p) = e^G(\nu _i)|_p\in H^1(B\mathbb Z^n_2;\mathbb Z_2). \end{aligned}$$

The inclusion map \(\psi _i: W_i\hookrightarrow W\) defines an equivariant Gysin homomorphism \(\psi _{i_{!}}: H^*_{G}(W_i;\mathbb Z_2)\rightarrow H^{*+1}_{G}(W;\mathbb Z_2)\) (see [1, §5.3] for example). For brevity, let

$$\begin{aligned} \tau _i =\tau _{F_i}= \psi _{i_{!}}(1)\in H^1_{G}(W;\mathbb Z_2) \end{aligned}$$

be the image of the identity \(1 \in H^0_{G}(W_i;\mathbb Z_2)\). The element \(\tau _i\) can be thought of as the Poincaré dual of the Borel construction of \(W_i\) in \(H^*_{G}(W;\mathbb Z_2)\) and is called the equivariant Thom class of \(\nu _i\). A standard fact is

$$\begin{aligned} \tau _i|_p\text { agrees with the equivariant Euler class of }\nu _i|_p. \end{aligned}$$

Note that the elements of \(\textrm{Hom}(\mathbb Z^n_2,\mathbb Z_2)\) are in one-to-one correspondence with all the 1-dimensional linear representations of \(\mathbb Z^n_2\). So the canonical isomorphism between \(\textrm{Hom}(\mathbb Z^n_2,\mathbb Z_2)\) and \(H^1(B\mathbb Z^n_2;\mathbb Z_2)\) is given by the equivariant Euler class of a 1-dimensional representations of \(\mathbb Z^n_2\). Then, we have the following identification:

$$\begin{aligned} \alpha _p = \{ \alpha (e)\,|\, p\in e \} \longleftrightarrow \{ e^G(\nu _i)|_p=\tau _i|_p\,;\, i\in I(p) \}. \end{aligned}$$
(7)

where an edge e containing p corresponds to the unique index \(i\in I(p)\) so that the facet \(F_i\) intersects e transversely (or equivalently \(e \nsubseteq F_i\)).

  • For a codimension-k face f of Q, let \(\nu _f\) denote the (equivariant) normal bundle of \(W_f\) in W. Denote by \(\tau _f \in H^k_G(W;\mathbb Z_2)\) the equivariant Thom class of \(\nu _f\). Then, the restriction of \(\tau _f\) to \(H^k_G(W_f;\mathbb Z_2)\) is the equivariant Euler class of \(\nu _f\), denoted by \(e^G(\nu _f)\). In particular, if \(f=Q\), \(W_f=W\) and so \(\tau _f\) is the identity element of \(H^0_G(W_f;\mathbb Z_2)\).

Let \(r_p: H^*_G(W;\mathbb Z_2)\rightarrow H_G^*(p;\mathbb Z_2)\cong H^*(BG;\mathbb Z_2)\) denote the restriction map at a fixed point \(p\in W^G\). Then,

$$\begin{aligned} r=\bigoplus _{p\in W^G} r_p : H^*_G(W;\mathbb Z_2)\rightarrow H^*_G(W^G;\mathbb Z_2) = \bigoplus _{p\in W^G} H^*(BG;\mathbb Z_2). \end{aligned}$$
(8)

By Theorem 2.3, the kernel of r is the \(H^*(BG;\mathbb Z_2)\)-torsion subgroup of \(H^*_G(W;\mathbb Z_2)\).

Clearly, \(r_p(\tau _f) =0\) unless \(p\in (W_f)^G\) (i.e., p is a vertex of f). It follows from (7) that for any \(p\in W^G\),

$$\begin{aligned} r_p(\tau _f)= {\left\{ \begin{array}{ll} \underset{p\in e,\, e\nsubseteq f}{\prod }\ \alpha (e), &{} \text {if }p \in f; \\ \ \ 0, &{} \text {otherwise}. \end{array}\right. } \end{aligned}$$
(9)

In addition, define

$$\begin{aligned} \widehat{H}^*_G(W;\mathbb Z_2):= H^*_G(W;\mathbb Z_2) \big /H^*(BG;\mathbb Z_2)\text {-torsion}. \end{aligned}$$
(10)

By the localization theorem (Theorem 2.3), the restriction homomorphism r induces a monomorphism \( \widehat{H}^*_G(W;\mathbb Z_2) \rightarrow H^*_G(W^G;\mathbb Z_2)\), still denoted by r.

The following proposition is parallel to [27, Proposition 3.3].

Proposition 1.5

Let W be an n-dimensional locally standard 2-torus manifold.

  1. (i)

    For each characteristic submanifold \(W_i\) with \((W_i)^{G}\ne \varnothing \) where \(G=\mathbb Z^n_2\), there is a unique element \(a_i\in H_1(BG;\mathbb Z_2)\) such that

    $$\begin{aligned} \rho ^*(t) = \sum _{i}\langle t,a_i\rangle \tau _i \ \, \text {modulo}\ \, H^*(BG;\mathbb Z_2)\text {-torsion} \end{aligned}$$

    for any element \(t\in H^1(BG;\mathbb Z_2)\). Here the sum is taken over all the characteristic submanifolds \(W_i\) with \((W_i)^G\ne \varnothing \) and \(\rho ^*\) is defined in (6).

  2. (ii)

    For each \(W_i\) with \((W_i)^G\ne \varnothing \), the subgroup \(G_i\) fixing \(W_i\) coincides with the subgroup determined by \(a_i\in H_1(BG;\mathbb Z_2)\) through the identification \(H_1(BG;\mathbb Z_2)\cong \textrm{Hom}(\mathbb Z_2,G)\).

  3. (iii)

    If n different characteristic submanifolds \(W_{i_1},\cdots , W_{i_n}\) have a G-fixed point in their intersection, then the elements \(a_{i_1},\cdots ,a_{i_n}\) form a linear basis of \(H_1(BG;\mathbb Z_2)\) over \(\mathbb Z_2\).

Proof

The argument is completely parallel to the arguments for torus manifolds in the proof of [26, Lemma 1,3, Lemma. 1.5, Lemma 1.7]. Indeed, we can just replace the torus manifold M in [26] by our 2-torus manifold W and replace \(T^n\) by \(\mathbb Z^n_2\) and \(H^2(M;\mathbb Z)\) by \(H^1(W;\mathbb Z_2)\) to obtain our proof here. The details of the proof are left to the reader.\(\square \)

In addition, the following lemma is completely parallel to the torus manifold case [27, Lemma 6.2]. Its proof is also parallel to [27], hence omitted.

Lemma 1.6

Let W be a locally standard 2-torus manifold with orbit space Q. For any \(\eta \in H^*_G(W;\mathbb Z_2)\) and any edge \(e\in E(Q)\), \(r_p(\eta )-r_{p'}(\eta )\) is divisible by \(\alpha (e)\) where p and \(p'\) are the endpoints of e.

2.4 Face Ring

A poset (partially ordered set) \(\mathcal {P}\) is called simplicial if it has an initial element \(\hat{0}\) and for each \(x\in \mathcal {P}\) the lower segment \([\hat{0},x]\) is a boolean lattice (the face lattice of a simplex).

Let \(\mathcal {P}\) be a simplicial poset. For each \(x\in \overline{\mathcal {P}}:=\mathcal {P}-\{\hat{0}\}\), we assign a geometrical simplex whose face poset is \([\hat{0},x]\) and glue these geometrical simplices together according to the order relation in \(\mathcal {P}\). The cell complex we obtained is called the geometrical realization of \(\mathcal {P}\), denoted by \(|\mathcal {P}|\). We may also say that \(|\mathcal {P}|\) is a simplicial cell complex.

For any two elements \(x, x'\in \mathcal {P}\), denote by \(x\vee x'\) the set of their least common upper bounds, and by \(x\wedge x'\) their greatest common lower bounds. Since \(\mathcal {P}\) is simplicial, \(x\wedge x'\) consists of a single element if \(x\vee x'\) is non-empty.

Definition 1.7

(see Stanley [29]) The face ring of a simplicial poset \(\mathcal {P}\) over a field \(\textbf{k}\) is the quotient

$$\begin{aligned} \textbf{k}[\mathcal {P}]:= \textbf{k}[v_x:x\in \mathcal {P}] \big /\mathcal {I}_{\mathcal {P}} \end{aligned}$$

where \(\mathcal {I}_{\mathcal {P}}\) is the ideal generated by all the elements of the form

$$\begin{aligned} v_x v_{x'} - v_{x\wedge x'}\cdot \sum _{x''\in x\vee x'} v_{x''}. \end{aligned}$$

Let Q be a nice manifold with corners. It is easy to see that the face poset of Q is a simplicial poset, denoted by \(\mathcal {P}_Q\). We call \(|\mathcal {P}_Q|\) the simplicial cell complex dual to Q.

We define the face ring of Q to be the face ring of \(\mathcal {P}_Q\). Equivalently, we can write the face ring of Q as

$$\begin{aligned} \textbf{k}[Q]:=\textbf{k}[v_f: f \ \text {a face of }Q] \big /\mathcal {I}_Q. \end{aligned}$$

where \(\mathcal {I}_Q\) is the ideal generated by all the elements of the form

$$\begin{aligned} v_f v_{f'} - v_{f\vee f'}\cdot \sum _{f''\in f\cap f'} v_{f''}. \end{aligned}$$

where \(f\vee f'\) denotes the unique minimal face of Q containing both f and \(f'\).

Convention: For any face f of Q, define the degree of \(v_f\) to be the codimension of f. Then, \(\textbf{k}[Q]=\textbf{k}[\mathcal {P}_Q]\) becomes a graded ring. Note that in the discussion of torus manifolds in [27], the degree of \(v_f\) is defined to be twice the codimension of f to fit the study there.

The f-vector of Q is defined as \(\textbf{f}(Q) =(f_0,\cdots , f_{n-1})\) where \(n=\dim (Q)\) and \(f_i\) is the number of faces of codimension \(i+1\). The equivalent information is contained in the h-vector \(\textbf{h}(Q)=(h_0,\cdots ,h_n)\) determined by the equation:

$$\begin{aligned} h_0 t^n +\cdots + h_{n-1}t+ h_n = (t-1)^n + f_0 (t-1)^{n-1}+\cdots + f_{n-1}. \end{aligned}$$
(11)

The Hilbert series of \(\textbf{k}[Q]\) is \(F(\textbf{k}[Q]):=\sum _i \dim _{\textbf{k}}\textbf{k}[Q]_{i} \cdot t^i\) where \(\textbf{k}[Q]_{i}\) denotes the homogeneous degree i part of \(\textbf{k}[Q]\). By [29, Proposition 3.8],

$$\begin{aligned} F(\textbf{k}[Q];t) = \frac{h_0 + h_1 t +\cdots + h_nt^n}{(1-t)^n}. \end{aligned}$$
(12)

The following construction is taken from [27, Sect. 5]. For any vertex (0-face) \(p\in Q\), we define a map

$$\begin{aligned} s_p: \textbf{k}[Q] \rightarrow \textbf{k}[Q]\big /(v_f: p\notin f). \end{aligned}$$
(13)

If p is the intersection of n different facets \(F_1,\cdots , F_n\), then \(\textbf{k}[Q]\big /(v_f: p\notin f)\) can be identified with the polynomial ring \(\textbf{k}[v_{F_1},\cdots ,v_{F_n}]\).

Lemma 1.8

(Lemma 5.6 in [27]) If every face of Q has a vertex, then the direct sum \(s=\bigoplus _{p} s_p\) over all vertices \(p\in Q\) is a monomorphism from \(\textbf{k}[Q]\) to the sum of polynomial rings \(\textbf{k}[Q]\big /(v_f: p\notin f)\).

A finitely generated graded commutative ring R over \(\textbf{k}\) is called Cohen-Macaulay if there exists an h.s.o.p (homogeneous system of parameters) \(\theta _1,\cdots ,\theta _n\) such that R is a free \(\textbf{k}[\theta _1,\cdots ,\theta _n]\)-module. Clearly, if \(\textbf{k}[Q]=\textbf{k}[\mathcal {P}_Q]\) is Cohen-Macaulay, then it has a l.s.o.p (linear system of parameters).

A simplicial complex K is called a Gorenstein* complex over \(\textbf{k}\) if its face ring \(\textbf{k}[K]\) is Cohen-Macaulay and \(H^*(K;\textbf{k})\cong H^*(S^{d};\textbf{k})\) where \(d=\dim (K)\). The reader is referred to Bruns-Herzog [9] and Stanley [30] for more information of Cohen-Macaulay rings and Gorenstein* complexes.

The following proposition is parallel to [27, Lemma 8.2 (1)].

Proposition 1.9

If Q is an n-dimensional mod 2 homology polytope, then the geometrical realization \(|\mathcal {P}_Q|\) of \(\mathcal {P}_Q\) is a Gorenstein* simplicial complex over \(\mathbb Z_2\). In particular, \(\mathbb Z_2[\mathcal {P}_Q]\) is Cohen-Macaulay and \(H^*(|\mathcal {P}_Q|;\mathbb Z_2)\cong H^*(S^{n-1};\mathbb Z_2)\).

Proof

The proof is almost identical to the proof in [27, Lemma 8.2] except that we use \(\mathbb Z_2\)-coefficients instead of \(\mathbb Z\)-coefficients when applying [30, II 5.1] in the argument.\(\square \)

3 Equivariantly Formal 2-Torus Manifolds

In this section, we study various properties of equivariantly formal 2-torus manifolds. One may find that many discussions on 2-torus manifolds here are parallel to the discussions in [27] on torus manifolds. The condition “vanishing of odd degree cohomology” on a torus manifold in [27] is now replaced by the equivariant formality condition on a 2-torus manifold and, the coefficients \(\mathbb Z\) is replaced by \(\mathbb Z_2\). Many arguments in [27] are transplanted into our proof here while some of them actually become simpler.

In Sect. 3.1, we prove some general results of equivariantly formal \(\mathbb Z^r_2\)-actions on compact manifolds. In particular, we prove that any equivariantly formal 2-torus manifold is locally standard, and the equivariant formality of a 2-torus manifold is inherited by all its facial submanifolds.

In Sect. 3.2, we explore the relations between the equivariant cohomology of a locally standard 2-torus manifold and the face ring of its orbit space.

In Sect. 3.3, we prove that the equivariant formality of a 2-torus manifold is preserved under real blow-ups along its facial submanifolds. Our proof uses a result from Gitler [17].

3.1 Equivariantly Formal \(\Rightarrow \) Locally Standard

Lemma 1.10

Suppose M is a compact manifold whose connected components are \(M_1,\cdots , M_k\). A \(\mathbb Z_2^r\)-action on M is equivariantly formal if and only if each \(M_i\) is \(\mathbb Z_2^r\)-invariant and the restricted \(\mathbb Z_2^r\)-action on \(M_i\) is equivariantly formal.

Proof

The “if” part is obvious. For the “only if” part, assume that \(M_{1},\cdots , M_{s}\), \( s\le k\), are all the components each of which is preserved under the \(\mathbb Z_2^r\)-action. Since the \(\mathbb Z_2^r\)-action on M is equivariantly formal, by Theorem 1.1 we have

$$\dim _{\mathbb Z_2}H^*(M^{\mathbb Z^r_2};\mathbb Z_2)= \dim _{\mathbb Z_2}H^*(M;\mathbb Z_2).$$

So in particular, \(M^{\mathbb Z^r_2}\) is not empty. Clearly, \(M^{\mathbb Z^r_2}\) must lie in \(M_{1}\cup \cdots \cup M_{s}\), so \(s>0\) and \(M^{\mathbb Z^r_2}\) is the disjoint union of \(M_1^{\mathbb Z^r_2},\cdots , M_s^{\mathbb Z^r_2}\). Then, by Theorem 1.1,

$$\begin{aligned} \dim _{\mathbb Z_2}H^*(M^{\mathbb Z^r_2};\mathbb Z_2)= & {} \sum ^s_{i=1} \dim _{\mathbb Z_2}H^*(M_i^{\mathbb Z^r_2};\mathbb Z_2) \le \sum ^s_{i=1} \dim _{\mathbb Z_2}H^*(M_i;\mathbb Z_2)\\\le & {} \dim _{\mathbb Z_2}H^*(M;\mathbb Z_2). \end{aligned}$$

By comparing this inequality with the previous equation, we can deduce that \(s=k\) and on every component \(M_i\), \(\dim _{\mathbb Z_2}H^*(M_i^{\mathbb Z^r_2};\mathbb Z_2) = \dim _{\mathbb Z_2}H^*(M_i;\mathbb Z_2)\). So by Theorem 1.1 again, the \(\mathbb Z_2^r\)-action on \(M_i\) is equivariantly formal.\(\square \)

Lemma 1.11

If a \(\mathbb Z_2^r\)-action on a compact manifold M is equivariantly formal, then for every subgroup H of \(\mathbb Z_2^r\),

  1. (i)

    The action of H on M is equivariantly formal.

  2. (ii)

    The induced action of \(\mathbb Z^r_2\) on \(M^H\) and \(\mathbb Z^r_2/H\) on \(M^H\) are both equivariantly formal.

  3. (iii)

    The induced action of \(\mathbb Z^r_2\) (or \(\mathbb Z^r_2/H\)) on every connected component N of \(M^H\) is equivariantly formal, hence N has a \(\mathbb Z^r_2\)-fixed point.

Proof

  1. (i)

    By Theorem 1.1, it is equivalent to prove

    $$\begin{aligned} \dim _{\mathbb Z_2} H^*(M^H;\mathbb Z_2)=\dim _{\mathbb Z_2} H^*(M;\mathbb Z_2). \end{aligned}$$
    (14)

    Otherwise, assume \(\dim _{\mathbb Z_2} H^*(M^H;\mathbb Z_2)<\dim _{\mathbb Z_2} H^*(M;\mathbb Z_2)\). Observe that the \(\mathbb Z^r_2\)-action on M induces an action of \(\mathbb Z_2^r/H\) on \(M^H\) and we have

    $$\begin{aligned} M^{\mathbb Z^r_2} = (M^H)^{\mathbb Z^r_2 /H}. \end{aligned}$$
    (15)

    So by Theorem 1.1, \(\dim _{\mathbb Z_2} H^*(M^{\mathbb Z^r_2};\mathbb Z_2)\le \dim _{\mathbb Z_2} H^*(M^{H};\mathbb Z_2) <\dim _{\mathbb Z_2} H^*\)\((M;\mathbb Z_2)\), which contradicts the assumption that the \(\mathbb Z_2^r\)-action on M is equivariantly formal. This proves (i).

  2. (ii)

    By (15) and the assumption that the \(\mathbb Z^r_2\)-action is equivariantly formal,

    $$\begin{aligned} \dim _{\mathbb Z_2} H^*\big ((M^H)^{\mathbb Z^r_2 /H};\mathbb Z_2\big )&= \dim _{\mathbb Z_2} H^*\big ( M^{\mathbb Z^r_2};\mathbb Z_2\big ) \\&= \dim _{\mathbb Z_2} H^*\big ( M;\mathbb Z_2\big )\overset{14}{=}\ \dim _{\mathbb Z_2} H^*(M^H;\mathbb Z_2). \end{aligned}$$

    Then, by Theorem 1.1, the action of \(\mathbb Z^r_2/H\) on \(M^H\) is equivariantly formal, so is the action of \(\mathbb Z^r_2\) on \(M^H\).

  3. (iii)

    By the conclusion in (ii) and Lemma 3.1, the induced action of \(\mathbb Z^r_2\) (or \(\mathbb Z^r_2/H\)) on every connected component N of \(M^H\) is equivariantly formal. So by Theorem 1.1, N must have a \(\mathbb Z^r_2\)-fixed point.

\(\square \)

Next, we prove a theorem that is parallel to [27, Theorem 4.1].

Theorem 1.12

If a 2-torus manifold W is equivariantly formal, then W must be locally standard.

Proof

Suppose \(\dim (W)=n\). For a point \(x\in W\), denote by \(G_x\) the isotropy group of x.

  • If \(G_x\) is trivial, then x is in a free orbit of the \(\mathbb Z_2^n\)-action. So W is locally standard near x.

  • Otherwise, let N be the connected component of \(W^{G_x}\) containing x. By Lemma 3.2 (iii), the induced \(\mathbb Z_2^n\)-action on N has a fixed point, say \(x_0\). Since \(W^{\mathbb Z^n_2}\) is discrete, the tangential \(\mathbb Z^n_2\)-representation \(T_{x_0}W\) is faithful. Then, since x and \(x_0\) are in the same connected component fixed pointwise by \(G_x\), the \(G_x\)-representation on \(T_x W\) agrees with the restriction of the tangential \(\mathbb Z^n_2\)-representation \(T_{x_0} W\) to \(G_x\). This implies that W is locally standard near x.

The theorem is proved.\(\square \)

Proposition 1.13

Let W be an equivariantly formal 2-torus manifold with orbit space Q. For any face f of Q, the facial submanifold \(W_f\) is also an equivariantly formal 2-torus manifold.

Proof

Suppose \(\dim (W)=n\) and f is a codimension-k face of Q. By Theorem 3.3, W is locally standard. Then, \(W_f\) is a connected \((n-k)\)-dimensional embedded submanifold of W fixed pointwise by \(G_f\cong \mathbb Z^k_2\) (see (3)). By Lemma 3.2 (iii), the induced action of \(\mathbb Z^n_2/G_f \cong \mathbb Z^{n-k}_2\) on \(W_f\) is equivariantly formal.\(\square \)

3.2 Equivariant Cohomology of Locally Standard 2-Torus Manifolds

Let W be an n-dimensional locally standard 2-torus manifold with orbit space Q. We explore the relation between \(H^*_G(W;\mathbb Z_2)\) where \(G=\mathbb Z^n_2\) and the face ring \(\mathbb Z_2[Q]\) under some conditions on Q. In the following, we use the notations from Sect. 2.3.

First of all, we have a lemma that is parallel to [27, Lemma 6.3].

Lemma 1.14

For any faces f and \(f'\) of Q, the relation below holds in \(\widehat{H}^*_G(W;\mathbb Z_2)\):

$$\begin{aligned} \tau _f\tau _{f'} = \tau _{f\vee f'}\cdot \sum _{f''\in f\cap f'} \tau _{f''}. \end{aligned}$$

Here we define \(\tau _{\varnothing }=0\).

Proof

The proof is parallel to the proof of [27, Lemma 6.3]. The idea is to use the monomorphism \(r: \widehat{H}^*_G(W;\mathbb Z_2) \rightarrow H^*_G(W^G;\mathbb Z_2)\) to map both sides of the identity to the fixed points and then use the formula (9) to check that they are equal. \(\square \)

By Lemma 3.5, we obtain a well-defined homomorphism

$$\begin{aligned} \varphi : \mathbb Z_2[Q]&\longrightarrow \widehat{H}^*_G(W;\mathbb Z_2).\\ v_f\ {}&\longmapsto \tau _f \end{aligned}$$

The following lemma and its proof are parallel to [27, Lemma 6.4].

Lemma 1.15

The homomorphism \(\varphi \) is injective if every face Q has a vertex.

Proof

According to the definitions of r and s (see (8) and (13)), we have \(s=r\circ \varphi \) by identifying \(H^*_G(p,\mathbb Z_2)\) with \(\mathbb Z_2[Q]\big /(v_f: p\notin f)\) for every vertex p of Q. Then, by Lemma 2.8, s is injective if every face of Q has a vertex, so is \(\varphi \).\(\square \)

The following lemma is parallel to [27, Proposition 7.4].

Lemma 1.16

If the 1-skeleton of every face of Q (including Q itself) is connected, then \(\widehat{H}^*_G(W;\mathbb Z_2)\) is generated by the elements \(\tau _{F_1},\cdots ,\tau _{F_m} \in H^1_G(W;\mathbb Z_2)\) as an \(H^*(BG;\mathbb Z_2)\)-module, where \(F_1,\cdots , F_m\) are all the facets of Q.

Proof

The argument is a bit technical, but it is completely parallel to the proof of [27, Proposition 7.4]. The main idea of the proof is to consider the restriction of an element \(\eta \in H^*_G(W;\mathbb Z_2)\) to the fixed point set \(W^G\) via \(r: H^*_G(W;\mathbb Z_2) \rightarrow H^*_G(W^G;\mathbb Z_2)\), and then use \(\tau _{F_1},\cdots ,\tau _{F_m}\) and elements in \(H^*(BG;\mathbb Z_2)\) to spell out \(r(\eta )\) at each fixed point \(p\in W^G\) (see Proposition 2.5). The details of the proof are left to the reader.\(\square \)

The following theorem is parallel to [27, Theorem 7.5].

Theorem 1.17

Let W be a locally standard 2-torus manifold with orbit space Q. If every face f of Q has a vertex and the 1-skeleton of f is connected, then the map \(\varphi : \mathbb Z_2[Q] \rightarrow \widehat{H}^*_G(W;\mathbb Z_2)\) is an isomorphism of graded rings.

Proof

By Lemma 3.6, \(\varphi \) is injective and, by Lemma 3.7, \(\varphi \) is surjective.\(\square \)

Lemma 1.18

Let W be an equivariantly formal 2-torus manifold with orbit space Q. Then the 1-skeleton of every face of Q (including Q itself) is connected.

Proof

Since W is equivariantly formal, the localization theorem (Theorem 2.3) implies that the restriction homomorphism \(r: H^*_{G}(W;\mathbb Z_2) \rightarrow H^*_{G}(W^{G};\mathbb Z_2)\) is injective. In addition, since W is connected, the image of \(H^0_{G}(W;\mathbb Z_2)\) under the restriction homomorphism is isomorphic to \(\mathbb Z_2\). So the “if” part of Theorem 2.4 implies that the 1-skeleton of Q must be connected.

For any proper face f of Q, the facial submanifold \(W_f\) is also an equivariantly formal 2-torus manifold by Proposition 3.4. Then, by applying the above argument to \(W_f\), we obtain that the 1-skeleton of f is also connected.\(\square \)

Corollary 1.19

If W is an equivariantly formal 2-torus manifold, then the map \(\varphi : \mathbb Z_2[Q] \rightarrow H^*_G(W;\mathbb Z_2)\) is an isomorphism of graded rings.

Proof

Since W is equivariantly formal, its equivariant cohomology \(H^*_G(W;\mathbb Z_2)\) is a free module over \(H^*(BG;\mathbb Z_2)\). So by definition, \(\widehat{H}^*_G(W;\mathbb Z_2) = H^*_G(W;\mathbb Z_2)\). For any face f of Q, the facial submanifold \(W_f\) is also an equivariantly formal 2-torus manifold by Proposition 3.4. This implies that f has a vertex. Moreover, the 1-skeleton of f is connected by Lemma 3.9. Then, the corollary follows from Theorem 3.8.\(\square \)

When a 2-torus manifold W is equivariantly formal, Corollary 3.10 tells us that the equivariant cohomology ring of W is completely determined by the face poset of its orbit space (so independent on the characteristic function \(\lambda _W\) or the principal bundle \(\xi _W\)). This suggests that the orbit space of W should be rather special.

The following corollary is parallel to [27, Corollary 7.8]. It generalizes the calculation of the mod 2 cohomology ring of a small cover in [14].

Corollary 1.20

If a 2-torus manifold W is equivariantly formal, then

$$\begin{aligned} H^*(W;\mathbb Z_2)\cong \mathbb Z_2[v_f: f\ \text {a face of }Q] \big /I \end{aligned}$$

where I is the ideal generated by the following two types of elements: \(\mathrm {(a)}\) \(v_fv_{f'} - v_{f\vee f'}\sum _{f''\in f\cap f'} v_{f''}\), \(\mathrm {(b)}\) \(\sum ^m_{i=1}\langle t,a_i\rangle v_{F_i}\), \(t\in H^1(BG;\mathbb Z_2)\).

Here, \(F_1,\cdots ,F_m\) are all the facets of Q, and the elements \(a_i\in H_1(BG;\mathbb Z_2)\) are defined in Proposition 2.5.

Proof

Since W is equivariantly formal, \(\iota _{W}^*: H^*_G(W;\mathbb Z_2)\rightarrow H^*(W;\mathbb Z_2)\) is surjective and its kernel is generated by all \(\rho ^*(t)\) with \(t\in H^1(BG;\mathbb Z_2)\) (see (6)). Then, the statement follows from Corollary 3.10 and Proposition 2.5.\(\square \)

3.3 Real Blow-up of a Locally Standard 2-Torus Manifold Along a Facial Submanifold

Let W be a locally standard 2-torus manifold with orbit space Q. For a codimension-k face f of Q, the facial submanifold \(W_f\) is an embedded connected codimension-k submanifold of W. So the equivariant normal bundle \(\nu _f\) of \(W_f\) in W is a real vector bundle of rank k. If we replace \(W_f \subset W\) by the real projective bundle \(P(\nu _f)\), we obtain a new 2-torus manifold denoted by \(\widetilde{W}^f\) called the real blow-up of W along \(W_f\). This is analogous to the blow-up of a torus manifold in [27, Sec. 9] (also see [20, p. 605] and [15, Sec. 4]).

The orbit space of \(\widetilde{W}^f\), denoted by \(Q^f\), is the result of “cutting off” the face f from Q (see Fig. 1). So \(\widetilde{W}^f\) is also locally standard. Correspondingly, the simplicial cell complex \(|\mathcal {P}_{Q^f}|\) is obtained from \(|\mathcal {P}_{Q}|\) by a stellar subdivision of the face dual to f.

Fig. 1
figure 1

Cutting off a face from a nice manifold with corners

Proposition 1.21

Let W be a locally standard 2-torus manifold with orbit space Q and f be a proper face of Q with codimension-k. Then, \(\widetilde{W}^f\) is equivariantly formal if and only if so is W.

Proof

(a) Let \(\widetilde{\nu }_f\) denote the equivariant normal bundle of \(P(\nu _f)\) in \(\widetilde{W}^f\). Besides, let \(\textrm{Th}(\nu _f)\) and \(\textrm{Th}(\widetilde{\nu }_f)\) be the Thom space of \(\nu _f\) and \(\widetilde{\nu }_f\), respectively. Then, we have a natural commutative diagram of continuous maps:

where i and \(\widetilde{i}\) are the inclusions; t and \(\widetilde{t}\) are the Thom-Pontryagin maps; \(p: \widetilde{W}^f\rightarrow W\) is the blow-down map; \(p_0\) is the restriction of p to \(P(\nu _f)\); and q is the induced map by p in the Thom spaces.

According to [17, §5] and [17, Theorem 3.7], there is a short exact sequence:

$$\begin{aligned} 0\! \longrightarrow \! H^*(\textrm{Th}(\nu _f);\mathbb Z_2)\!\overset{\alpha }{\longrightarrow }\! H^*(W;\mathbb Z_2)\oplus H^*(\textrm{Th}(\widetilde{\nu }_f);\mathbb Z_2)\! \overset{\beta }{\longrightarrow }\! H^*(\widetilde{W}^f;\mathbb Z_2)\!\longrightarrow \! 0. \end{aligned}$$
(16)

where \(\alpha =(t^*, q^*)\) and \(\beta =p^*-\widetilde{t}^*\). This implies:

$$\begin{aligned} \dim _{\mathbb Z_2} H^*(\widetilde{W}^f;\mathbb Z_2) = \dim _{\mathbb Z_2} H^*(W;\mathbb Z_2) + \dim _{\mathbb Z_2} H^*(\textrm{Th}(\widetilde{\nu }_f);\mathbb Z_2) - \dim _{\mathbb Z_2} H^*(\textrm{Th}(\nu _f);\mathbb Z_2). \end{aligned}$$
$$\begin{aligned} \dim _{\mathbb Z_2} H^*(\widetilde{W}^f;\mathbb Z_2) = \dim _{\mathbb Z_2} H^*(W;\mathbb Z_2) + \dim _{\mathbb Z_2} H^*(\textrm{Th}(\widetilde{\nu }_f);\mathbb Z_2) - \dim _{\mathbb Z_2} H^*(\textrm{Th}(\nu _f);\mathbb Z_2). \end{aligned}$$

By the Thom isomorphism, we have

$$\begin{aligned} \dim _{\mathbb Z_2} H^*(\textrm{Th}(\nu _f);\mathbb Z_2)&= \dim _{\mathbb Z_2} H^*(W_f;\mathbb Z_2), \\ \dim _{\mathbb Z_2} H^*(\textrm{Th}(\widetilde{\nu }_f);\mathbb Z_2)&= \dim _{\mathbb Z_2} H^*(P(\nu _f);\mathbb Z_2). \end{aligned}$$

By Leray-Hirsch theorem, \(H^*(P(\nu _f);\mathbb Z_2)\cong H^*(W_f;\mathbb Z_2)\otimes H^*(\mathbb RP^{k-1};\mathbb Z_2)\) (as \(\mathbb Z_2\)-vector spaces), which implies \( \dim _{\mathbb Z_2} H^*(P(\nu _f);\mathbb Z_2) = k \cdot \dim _{\mathbb Z_2} H^*(W_f;\mathbb Z_2)\). So

$$\begin{aligned} \dim _{\mathbb Z_2} H^*(\widetilde{W}^f;\mathbb Z_2) = \dim _{\mathbb Z_2} H^*(W;\mathbb Z_2) + (k-1) \cdot \dim _{\mathbb Z_2} H^*(W_f;\mathbb Z_2). \end{aligned}$$
(17)

If W is equivariantly formal, then W is locally standard and so Q is a nice manifold with corners. It is easy to see

$$ \# \text {vertices of }Q^f = \# \text {vertices of }Q + (k-1)\cdot \# \text {vertices of }f. $$

Since the fixed point set \(W^{G}\) (\(G=\mathbb Z^n_2\)) corresponds to the vertex set of Q which is discrete, the number of fixed points of the G-action satisfies

$$\begin{aligned} |(\widetilde{W}^f)^G| = |W^G| + (k-1)\cdot |(W_f)^G|. \end{aligned}$$
(18)

By Proposition 3.4, \(W_f\) is also equivariantly formal. So by Theorem 1.1,

$$\begin{aligned} \dim _{\mathbb Z_2} H^*(W;\mathbb Z_2)=|W^G|, \ \ \dim _{\mathbb Z_2} H^*(W_f;\mathbb Z_2)=|(W_f)^G|. \end{aligned}$$

It follows from (17) and (18) that \(|(\widetilde{W}^f)^G| = \dim _{\mathbb Z_2} H^*(\widetilde{W}^f;\mathbb Z_2)\). So we deduce from Theorem 1.1 that \(\widetilde{W}^f\) is equivariantly formal.

Conversely, if \(\widetilde{W}^f\) is equivariantly formal, we have

$$\begin{aligned} \dim _{\mathbb Z_2} H^*(W;\mathbb Z_2)&\overset{17}{=}\ \dim _{\mathbb Z_2} H^*(\widetilde{W}^f;\mathbb Z_2) - (k-1) \cdot \dim _{\mathbb Z_2} H^*(W_f;\mathbb Z_2)\\ (\text {by Theorem~1.1}) \,&\le |(\widetilde{W}^f)^G| - (k-1)\cdot |(W_f)^G| \overset{17}{=}\ |W^G| = \dim _{\mathbb Z_2} H^*(W^G;\mathbb Z_2). \end{aligned}$$

But by Theorem 1.1, \( \dim _{\mathbb Z_2} H^*(W^G;\mathbb Z_2) \le \dim _{\mathbb Z_2} H^*(W;\mathbb Z_2)\). So we must have \(\dim _{\mathbb Z_2} H^*(W;\mathbb Z_2)= \dim _{\mathbb Z_2} H^*(W^G;\mathbb Z_2)\), which implies that W is equivariantly formal. The proposition is proved.\(\square \)

The following lemma is parallel to [27, Lemma 9.1]. Its proof is almost identical to the proof in [27], hence omitted.

Lemma 1.22

Let Q be a nice manifold with corners and f be a proper face of Q. Then, \(Q^f\) is mod 2 face-acyclic if and only if so is Q.

4 Equivariantly Formal 2-Torus Manifolds with Mod 2 Cohomology Generated by Degree-One Part

In our study of equivariantly formal 2-torus manifolds, those manifolds whose mod 2 cohomology rings are generated by their degree-one part are of special importance. We will see in Sect. 5 that the study of general equivariantly formal 2-torus manifolds can be reduced to the study of these special 2-torus manifolds by a sequence of real blow-ups along facial submanifolds.

The following lemma is parallel to [27, Lemma 2.3].

Lemma 1.1

Suppose there is an equivariantly formal \(\mathbb Z_2^r\)-action on a compact manifold M where the cohomology ring \(H^*(M;\mathbb Z_2)\) is generated by its degree-one part. Then, for any subgroup H of \(\mathbb Z_2^r\) and every connected component N of \(M^H\), the homomorphism \(i^*: H^*(M;\mathbb Z_2)\rightarrow H^*(N;\mathbb Z_2)\) is surjective where \(i: N\hookrightarrow M\) is the inclusion. In particular, \(H^*(N;\mathbb Z_2)\) is also generated by its degree-one part.

Proof

First, we assume \(H\cong \mathbb Z_2\). We have a commutative diagram as follows:

(19)

where \(H^*_H(N;\mathbb Z_2)\cong H^*(N;\mathbb Z_2)\otimes H^*(BH;\mathbb Z_2)\) and \(\widehat{i}^*_H\) is the homomorphism on equivariant cohomology induced by i. By our assumption, both \(\iota ^*_M\) and \(\iota ^*_N\) are surjective. The following argument is parallel to the proof of [27, Lemma 2.3].

By [7, Theorem VII.1.5], the inclusion \(M^H \hookrightarrow M\) induces an isomorphism \(H^k_H(M;\mathbb Z_2)\rightarrow H^k_H(M^H;\mathbb Z_2) \) for sufficiently large k, which implies that

$$\widehat{i}^*_H: H^k_H(M;\mathbb Z_2)\rightarrow H^k_H(N;\mathbb Z_2)$$

is surjective if k is sufficiently large.

Let \(v_1,\cdots , v_d \in H^1(M;\mathbb Z_2)\) be a set of multiplicative generators of \(H^*(M;\mathbb Z_2)\). For each \(1\le l \le d\), let \(\widehat{v}_l\) be a lift of \(v_l\) in \(H^*_H(M;\mathbb Z_2)\) and \(w_l:= i^*(v_l)\in H^1(N;\mathbb Z_2)\). Let t be a generator of \(H^1(BH;\mathbb Z_2)\cong \mathbb Z_2\). By the commutativity of the above diagram (19),

$$\begin{aligned} \widehat{i}^*(\widehat{v}_l) = b_l t+w_l \ \,\text {for some}\ \ b_l\in \mathbb Z_2. \end{aligned}$$

Then, for an arbitrary element \(\zeta \in H^*(N;\mathbb Z_2)\), there exists a large enough integer \(q\in \mathbb Z\) and a polynomial \(P(x_1,\cdots , x_d)\) such that

$$ \widehat{i}^* \big ( P(\widehat{v}_1,\cdots , \widehat{v}_d) \big ) = \zeta \otimes t^{q}. $$

On the other hand, we have

$$\begin{aligned} \widehat{i}^* \big ( P(\widehat{v}_1,\cdots , \widehat{v}_d) \big ) = P(b_1 t+w_1,\cdots , b_d t+w_d)=\sum _{k\ge 0} P_k(w_1,\cdots ,w_d)\otimes t^k \end{aligned}$$

for some polynomials \(P_k\), \(k\ge 0\). Hence \(\zeta = P_q(w_1,\cdots , w_d) = i^*(P(v_1,\cdots , v_d))\). Therefore, \(i^*\) is surjective and \(H^*(N;\mathbb Z_2)\) is generated by \(w_1,\cdots , w_d \in H^1(N;\mathbb Z_2)\).

For the general case, suppose \(H\cong \mathbb Z^s_2\), \(1\le s\le r\). Then, we have a sequence:

$$\begin{aligned} \{0\}= H_0\subset H_1\subset H_2 \subset \cdots \subset H_s = H \end{aligned}$$

where \(H_l\cong \mathbb Z^l_2\) for each \(0\le l \le s\). Moreover, we have

$$\begin{aligned} M^H = \big ((M^{H_1})^{H_2/H_1})\cdots \big )^{H_s/H_{s-1}} ,\; H_l/H_{l-1}\cong \mathbb Z_2, \, l=1,\cdots , s. \end{aligned}$$

Repeating the above argument for each \(H_l/H_{l-1}\) proves the lemma.\(\square \)

The following lemma is parallel to [27, Lemma 3.4].

Lemma 1.2

Let W be an equivariantly formal 2-torus manifold whose cohomology ring \(H^*(W;\mathbb Z_2)\) is generated by its degree-one part. Then, all non-empty multiple intersections of the characteristic submanifolds of W are equivariantly formal 2-torus manifolds whose mod 2 cohomology rings are generated by their degree-one part as well.

Proof

Let \(F_1,\cdots , F_m\) be all the facets of Q and \(G=\mathbb Z^n_2\) where \(n=\dim (W)\). In the following, we use the notations defined in Sect. 2.3. First of all, since the characteristic submanifold \(W_i\) is a connected component of the fixed point set \(X^{G_i}\), Lemma 4.1 implies that the restriction \(H^*(W;\mathbb Z_2)\rightarrow H^*(W_i;\mathbb Z_2)\) is surjective. So the G-action on \(W_i\) is equivariantly formal (by Proposition 3.4). Then, we have

$$\begin{aligned} H^*_G(W;\mathbb Z_2)&\cong H^*(W;\mathbb Z_2)\otimes H^*(BG;\mathbb Z_2), \\ H^*_G(W_i;\mathbb Z_2)&\cong H^*(W_i;\mathbb Z_2)\otimes H^*(BG;\mathbb Z_2). \end{aligned}$$

It follows that the restriction \(H^*_G(W;\mathbb Z_2)\rightarrow H^*_G(W_i;\mathbb Z_2)\) is also surjective. In addition, by using Proposition 2.5 (i) and a completely parallel argument to the proof of [26, Prop. 3.4(2)], we can prove the following claim:

Claim: \(H^*_G(W;\mathbb Z_2)\) is generated as a ring by all the equivariant Thom classes \(\tau _1,\cdots ,\tau _m\) of the normal bundles of the characteristic submanifolds \(W_1,\cdots , W_m\).

When \(W_{j_1}\cap \cdots \cap W_{j_s} = \varnothing \), \(\tau _{j_1}\cdots \tau _{j_s}\) clearly vanishes. So the above claim implies that for any \(k\ge 0\), \(H^k_G(W;\mathbb Z_2)\) is additively generated by the monomials \(\tau ^{k_1}_{j_1}\cdots \tau ^{k_s}_{j_s}\) such that \(W_{j_1}\cap \cdots \cap W_{j_s}\ne \varnothing \) and \(k_1+\cdots +k_s =k\).

Let N be a connected component of \(W_{i_1}\cap \cdots \cap W_{i_k}\), \(1\le k \le n\). Then, N is the facial submanifold \(W_f\) over some codimension-k face f of Q. So by Lemma 4.1, N is an equivariantly formal 2-torus manifold whose cohomology ring \(H^*(N;\mathbb Z_2)\) is generated by its degree-one part. Moreover, by a completely parallel argument to the proof of [27, Lemma 3.4], we can show that N is the only connected component of \(W_{i_1}\cap \cdots \cap W_{i_k}\) from the above discussion of \(H^k_G(W;\mathbb Z_2)\). The lemma is proved.\(\square \)

The following proposition is parallel to [27, Lemma 8.2 (2)].

Proposition 1.3

Suppose W is an n-dimensional equivariantly formal 2-torus manifold with orbit space Q and the cohomology ring \(H^*(W;\mathbb Z_2)\) is generated by its degree-one part. Then, the geometrical realization \(|\mathcal {P}_Q|\) of the face poset \(\mathcal {P}_Q\) of Q is a Gorenstein* simplicial complex over \(\mathbb Z_2\). In particular, \(\mathbb Z_2[\mathcal {P}_Q]=\mathbb Z[Q]\) is Cohen-Macaulay and \(H^*(|\mathcal {P}_Q|;\mathbb Z_2)\cong H^*(S^{n-1};\mathbb Z_2)\).

Proof

By Lemma 4.2, all non-empty multiple intersections of the characteristic submanifolds of W are connected. This implies that \(|\mathcal {P}_Q|\) is a simplicial complex. Moreover, by [30, II 5.1(d)], it is enough to verify the following three conditions to prove that \(|\mathcal {P}_Q|\) is Gorenstein* over \(\mathbb Z_2\):

  1. (a)

    \(\mathbb Z_2[\mathcal {P}_Q]\) is Cohen-Macaulay;

  2. (b)

    Every \((n-2)\)-simplex in \(\mathcal {P}_Q\) is contained in exactly two \((n-1)\)-simplices;

  3. (c)

    \(\chi (\mathcal {P}_Q)=\chi (S^{n-1})\).

Since W is equivariantly formal, \(H^*_G(W;\mathbb Z_2)\) is a free \(H^*(BG;\mathbb Z_2)\)-module and \(\mathbb Z_2[\mathcal {P}_Q]=\mathbb Z_2[Q]\) is isomorphic to \(H^*_G(W;\mathbb Z_2)\) (by Corollary 3.10) where \(G=\mathbb Z^n_2\). This implies (a).

Note that each \((n-2)\)-simplex of \(\mathcal {P}_Q\) corresponds to a non-empty intersection of \(n-1\) characteristic submanifolds of W. The latter intersection is an equivariantly formal 1-manifold by Lemma 4.2, so it is a circle with exactly two G-fixed points. This implies (b).

The proof of (c) is completely parallel to [27, Lemma 8.2 (2)], so we leave it to the reader. The proposition is proved.\(\square \)

Using the above proposition and the lemmas from Sect. 3, we obtain the following theorem that is parallel to [27, Theorem 7.7].

Theorem 1.4

Let W be a 2-torus manifold whose orbit space is Q. Then, W is equivariantly formal and the cohomology ring \(H^*(W;\mathbb Z_2)\) is generated by its degree-one part if and only if the following three conditions are satisfied:

  1. (a)

    \(H^*_G(W;\mathbb Z_2)\) is isomorphic to \(\mathbb Z_2[Q]=\mathbb Z_2[\mathcal {P}_Q]\) as a graded ring.

  2. (b)

    \(\mathbb Z_2[Q]\) is Cohen-Macaulay.

  3. (c)

    \(|\mathcal {P}_Q|\) is a simplicial complex.

Proof

The argument is completely parallel to the proof of [27, Theorem 7.7]. We only need to replace \(T^n\) by \(\mathbb Z^2_n\) and \(\mathbb {Q}\)-coefficients by \(\mathbb Z_2\)-coefficients to obtain our proof here.\(\square \)

5 Proof of Theorem 1.3

In this section, we give a proof of Theorem 1.3. Our proof follows the proof of [27, Theorem 8.3, Theorem 9.3] almost step by step, while some arguments for 2-torus manifolds here are simpler than those for torus manifolds in [27].

5.1 Equivariant Cohomology of the Canonical Model

Let Q be a connected compact smooth nice n-manifold with corners. We call any function \( \lambda : \mathcal {F}(Q)\rightarrow \mathbb Z^n_2\) that satisfies the linear independence relation in Sect. 2.1 a characteristic function on Q. By the same gluing rule in (4), we can obtain a space \(M_Q(\lambda )\) from any characteristic function \(\lambda \) on Q, called the canonical model determined by \((Q,\lambda )\). It is easy to see that \(M_Q(\lambda )\) is a 2-torus manifold of dimension n.

Let \(Q^{\vee }\) denote the cone of the geometrical realization of the order complex \(\textrm{ord}(\overline{\mathcal {P}}_Q)\) of \(\overline{\mathcal {P}}_Q=\mathcal {P}_Q-\{\hat{0}\}\). So topologically, \(Q^{\vee }\) is homeomorphic to \(\text {Cone}(|\mathcal {P}_Q|)\). Moreover, \(Q^{\vee }\) is a “space with faces” (see Davis [13, Sec. 6]) where each proper face f of Q determines a unique “face” \(f^{\vee }\) of \(Q^{\vee }\) that is the geometrical realization of the order complex of the poset \(\{ f' \,|\, f' \subseteq f \}\). More precisely, \(f^{\vee }\) consists of all simplices of the form \(f'_k\subsetneq \cdots \subsetneq f'_1\subsetneq f'_0=f\) in \(\textrm{ord}(\overline{\mathcal {P}}_Q)\). The “boundary” of \(Q^{\vee }\), denoted by \(\partial Q^{\vee }\), is \(\textrm{ord}(\overline{\mathcal {P}}_Q)\) which is homeomorphic to \(|\mathcal {P}_Q|\). So we have homeomorphisms:

$$\begin{aligned} \partial Q^{\vee } \cong |\mathcal {P}_Q|, \ \ Q^{\vee }\cong \textrm{Cone}(|\mathcal {P}_Q|). \end{aligned}$$
(20)

Remark 1.1

When \(|\mathcal {P}_Q|\) is a simplicial complex, the space \(Q^{\vee }\) with the face decomposition was called in [14, p. 428] a simple polyhedral complex.

Suppose \(F_1,\cdots , F_m\) are all the facets of Q. Let \(\mathcal {F}(Q^{\vee }) = \{ F^{\vee }_1,\cdots , F^{\vee }_m\}\). Then, any characteristic function \(\lambda : \mathcal {F}(Q)\rightarrow \mathbb Z^n_2\) induces a map \(\lambda ^{\vee }: \mathcal {F}(Q^{\vee })\rightarrow \mathbb Z^n_2\) where \(\lambda ^{\vee }(F^{\vee }_i) = \lambda (F_i)\), \(1\le i \le m\). Then, by the same gluing rule in (4), we obtain a space \(M_{Q^{\vee }}(\lambda ^{\vee })\) with a canonical \(\mathbb Z^n_2\)-action. By the same argument as in the proof of [27, Proposition 5.14], we can prove the following.

Proposition 1.2

There exists a continuous map \(\phi : Q\rightarrow Q^{\vee }\) which preserves the face structure and induces an equivariant continuous map

$$\begin{aligned} \Phi : M_Q(\lambda ) \rightarrow M_{Q^{\vee }}(\lambda ^{\vee }). \end{aligned}$$

Here \(\phi : Q\rightarrow Q^{\vee }\) is constructed inductively, starting from an identification of vertices and extending the map on each higher-dimensional face by a degree-one map. Since every face \(f^{\vee }\) of \(Q^{\vee }\) is a cone, there are no obstructions to such extensions.

In addition, by a similar argument to that in [14, Theorem 4.8], we can obtain the following result.

Proposition 1.3

\(H^*_G(M_{Q^{\vee }}(\lambda ^{\vee });\mathbb Z_2)\) is isomorphic to \(\mathbb Z_2[Q]\) where \(G=\mathbb Z^n_2\).

On the other hand, \(H^*_G(M_Q(\lambda );\mathbb Z_2)\) could be much more complicated. Indeed, it is shown in [31, Theorem 1.7] that \(H^*_G(M_Q(\lambda );\mathbb Z_2)\) isomorphic to the so called topological face ring of Q over \(\mathbb Z_2\) which involves the mod 2 cohomology rings of all the faces of Q.

5.2 Proof of Theorem 1.3 (ii)

Proof

We first prove the “if” part. Let Q be an n-dimensional mod 2 homology polytope and \(G=\mathbb Z^n_2\). Since \(H^1(Q;\mathbb Z_2)=0\) and W is locally standard, the principal G-bundle \(\xi _W\) determined by W is a trivial G-bundle over Q. Then, by [25, Lemma 3.1], W is equivariantly homeomorphic to the canonical model \(M_Q(\lambda _W)\) (see (4)). So by Proposition 5.3, there exists an equivariant continuous map

$$\begin{aligned} \Phi : W =M_Q(\lambda _W) \rightarrow M_{Q^{\vee }}(\lambda ^{\vee }_W):= W^{\vee }. \end{aligned}$$

Let \(\pi : W\rightarrow Q\) and \(\pi ^{\vee }: W^{\vee }\rightarrow Q^{\vee }\) be the projections, respectively. Let \(F_1,\cdots , F_m\) be all the facets of Q. Since Q is a mod 2 homology polytope, so are \(F_1,\cdots , F_m\). For brevity, let

$$\begin{aligned} W_i = \pi ^{-1}(F_i),\ \ W^{\vee }_i = (\pi ^{\vee })^{-1}(F^{\vee }_i), \ 1\le i \le m. \end{aligned}$$

It is easy to see that the \(\mathbb Z^n_2\)-actions on \(W\backslash \bigcup _i W_i\) and \(W^{\vee }\backslash \bigcup _i W^{\vee }_i\) are both free. Then, we have

$$ H^*_G\Big (\!W,\bigcup _i W_i;\mathbb Z_2 \Big ) \!\cong \! H^*(Q,\partial Q;\mathbb Z_2), \! \ H^*_G\Big (\!W^{\vee },\bigcup _i W^{\vee }_i;\mathbb Z_2\!\Big ) \!\cong \! H^*(Q^{\vee },\partial Q^{\vee };\!\mathbb Z_2). $$

So \(\Phi : W \rightarrow W^{\vee }\) induces a map between the following two exact sequences:

(21)

Each \(W_i\) is a 2-torus manifold over the homology polytope \(F_i\). So using induction and a Mayer-Vietoris argument, we may assume that in the diagram (21), \(\Phi ^*: H^*_G(\bigcup _i W^{\vee }_i;\mathbb Z_2)\rightarrow H^*_G(\bigcup _i W_i;\mathbb Z_2)\) is an isomorphism.

By Proposition 2.9, \(H^*(|\mathcal {P}_Q|;\mathbb Z_2)\cong H^*(S^{n-1};\mathbb Z_2)\). Then, by (20), we obtain

$$\begin{aligned} H^*(Q^{\vee },\partial Q^{\vee };\mathbb Z_2)\cong H^*(D^n,S^{n-1};\mathbb Z_2). \end{aligned}$$

We also have \(H^*(Q,\partial Q;\mathbb Z_2)\cong H^*(D^n,S^{n-1};\mathbb Z_2)\) since Q is an n-dimensional mod 2 homology polytope. By the construction of \(\phi \), it is easy to see that the homomorphism \(\phi ^*: H^*(Q^{\vee },\partial Q^{\vee };\mathbb Z_2)\rightarrow H^*(Q,\partial Q;\mathbb Z_2)\) is an isomorphism. Then, by applying the five-lemma to the diagram (21), we can deduce that \(\Phi ^*: H^*_G(W^{\vee };\mathbb Z_2) \rightarrow H^*_G(W;\mathbb Z_2)\) is also an isomorphism. So by Proposition 5.3, \(H^*_G(W;\mathbb Z_2)\cong \mathbb Z_2[Q]\).

Besides, we also know that \(\mathbb Z_2[Q]\) is Cohen-Macaulay by Proposition 2.9. Then, since \(|\mathcal {P}_Q|\) is a simplicial complex, all the three conditions in Theorem 4.4 are satisfied. Hence W is equivariantly formal and \(H^*(W;\mathbb Z_2)\) is generated by its degree-one part as a ring. The “if” part is proved.

Next, we prove the “only if” part. By the assumption on W and Lemma 4.2, all non-empty multiple intersections of characteristic submanifolds of W are connected and their cohomology rings are generated by their degree-one elements. So we may assume by induction that all the proper faces of Q are mod 2 homology polytopes. In particular, the proper faces of Q are all mod 2 acyclic. From these assumptions, we need to prove that Q itself is mod 2 acyclic.

By Proposition 4.3, \(|\mathcal {P}_Q|\) is a simplicial complex. So \(|\mathcal {P}_Q|\) is the nerve simplicial complex of the cover of \(\partial Q\) by the facets of Q. By a Mayer-Vietoris sequence argument, we can deduce that \(H^*(\partial Q;\mathbb Z_2)\cong H^*(|\mathcal {P}_Q|;\mathbb Z_2)\). This together with Proposition 4.3 shows that

$$\begin{aligned} H^*(\partial Q;\mathbb Z_2)\cong H^*(S^{n-1};\mathbb Z_2). \end{aligned}$$
(22)

Claim: \(H^1(Q;\mathbb Z_2)=0\).

Since W is equivariantly formal, \(H^*_G(W;\mathbb Z_2)\) is a free \(H^*(BG;\mathbb Z_2)\)-module. On the other hand, \(H^*(Q,\partial Q;\mathbb Z_2)\) is finitely generated over \(\mathbb Z_2\) since Q is compact. So \(H^*(Q,\partial Q;\mathbb Z_2)\) is a torsion \(H^*(BG;\mathbb Z_2)\)-module. It follows that the whole bottom row in the diagram (21) splits into short exact sequences:

$$\begin{aligned} 0\rightarrow H^k_G(W;\mathbb Z_2) \rightarrow H^k_G \Big (\bigcup _i W_i;\mathbb Z_2 \Big ) \rightarrow H^{k+1}(Q,\partial Q;\mathbb Z_2) \rightarrow 0, \ k\ge 0. \end{aligned}$$
(23)

Take \(k=0\) above, we clearly have \(H^0_G(W;\mathbb Z_2)\cong H^0_G \Big (\bigcup _i W_i;\mathbb Z_2 \Big )\cong \mathbb Z_2\). This implies \(H^1(Q,\partial Q;\mathbb Z_2)=0\). So in the following exact sequence,

$$\begin{aligned} \cdots \rightarrow H^1(Q,\partial Q;\mathbb Z_2)\rightarrow H^1(Q;\mathbb Z_2) \rightarrow H^1(\partial Q;\mathbb Z_2) \rightarrow \cdots , \end{aligned}$$

\(H^1(Q;\mathbb Z_2)\) is mapped injectively into \(H^1(\partial Q;\mathbb Z_2)\cong H^1(S^{n-1};\mathbb Z_2)\). Note that if \(n=1\), the claim is trivial. When \(n=2\), we have \(\partial Q =S^1\) and \(H^1(Q;\mathbb Z_2) = 0\) or \(\mathbb Z_2\). But by the classification of compact surfaces, the latter case is impossible. When \(n\ge 3\), we have \(H^1(\partial Q;\mathbb Z_2)=0\), so \(H^1(Q;\mathbb Z_2)=0\). The claim is proved.

Now since \(H^1(Q;\mathbb Z_2)=0\), by the above proof of the “if” part, there exists an equivariant homeomorphism \(\Phi \) from W to the canonical model \(M_Q(\lambda _W)\). In addition, by (20) and Proposition 4.3, we have

$$\begin{aligned} H^*(\partial Q^{\vee };\mathbb Z_2) \cong H^*( |\mathcal {P}_Q|;\mathbb Z_2) \cong H^*(S^{n-1};\mathbb Z_2). \end{aligned}$$

So we have an isomorphism

$$\begin{aligned} H^*(Q^{\vee },\partial Q^{\vee };\mathbb Z_2)\cong H^*(D^n,S^{n-1};\mathbb Z_2). \end{aligned}$$
(24)

Then, by the construction of \(\phi \), the map \(\phi ^*: H^*(Q^{\vee },\partial Q^{\vee };\mathbb Z_2) \!\rightarrow \! H^*(Q,\partial Q;\mathbb Z_2)\) is an isomorphism in degree n (since Q is connected) and thus is injective in all degrees. So by an extended version of the 5-lemma, we can deduce that in the diagram (21) the map \(\Phi ^*: H^*_G(W^{\vee };\mathbb Z_2)\rightarrow H^*_G(W;\mathbb Z_2)\) is injective. Moreover,

  • \(H^*_G(W^{\vee };\mathbb Z_2)=H^*_G(M_{Q^{\vee }}(\lambda ^{\vee }_W);\mathbb Z_2)\cong \mathbb Z_2[Q]\) by Proposition 5.3, and

  • \(\mathbb Z_2[Q]\cong H^*_G(W;\mathbb Z_2)\) by Corollary 3.10.

So \(H^*_G(W^{\vee };\mathbb Z_2)\) and \(H^*_G(W;\mathbb Z_2)\) have the same dimension over \(\mathbb Z_2\) in each degree. Therefore, the monomorphism \(\Phi ^*: H^*_G(W^{\vee };\mathbb Z_2)\rightarrow H^*_G(W;\mathbb Z_2)\) is actually an isomorphism. Then, by the 5-lemma again, we can deduce from the diagram (21) that \(\phi ^*: H^*(Q^{\vee },\partial Q^{\vee };\mathbb Z_2)\rightarrow H^*(Q,\partial Q;\mathbb Z_2)\) is an isomorphism. So by (24),

$$\begin{aligned} H^*(Q,\partial Q;\mathbb Z_2)\cong H^*(D^n,S^{n-1};\mathbb Z_2) \end{aligned}$$

which implies that Q is mod 2 acyclic by Poincaré-Lefschetz duality. This finishes the proof.\(\square \)

5.3 Proof of Theorem 1.3 (i)

Proof

We can reduce Theorem 1.3 (i) to Theorem 1.3 (ii) by real blow-ups of W along sufficient many facial submanifolds, which corresponds to doing some barycentric subdivisions of the face poset \(\mathcal {P}_Q\) of Q (see Fig. 2). Indeed, after doing enough barycentric subdivisions to \(\mathcal {P}_Q\), we can turn \(|\mathcal {P}_Q|\) into a simplicial complex. Let \(\widehat{W}\) be the 2-torus manifold obtained after these real blow-ups on W and \(\widehat{Q}\) be its orbit space (with \(|\mathcal {P}_{\widehat{Q}}|\) being a simplicial complex).

  1. Fact-1:

    \(\widehat{W}\) is equivariantly formal if and only if so is W (by Proposition 3.12).

  2. Fact-2:

    \(\widehat{Q}\) is mod 2 face-acyclic if and only if so is Q (by Lemma 3.13).

Fig. 2
figure 2

Cutting a vertex and an edge

We first prove the “if” part. Suppose W is locally standard and Q is mod 2 face-acyclic. Then, \(\widehat{W}\) is also locally standard and \(\widehat{Q}\) is a mod 2 homology polytope by Fact-2. So by Theorem 1.3 (ii), \(\widehat{W}\) is equivariantly formal, then so is W.

Next, we prove the “only if” part. If W is equivariantly formal, then so is \(\widehat{W}\), and W is locally standard by Theorem 3.3. So by Corollary 3.10, we have a graded ring isomorphism \(H^*_G(\widehat{W};\mathbb Z_2)\cong \mathbb Z_2[\widehat{Q}]\). Moreover, since \(|\mathcal {P}_{\widehat{Q}}|\) is a simplicial complex, \(\mathbb Z_2[\widehat{Q}]\) is generated by its degree-one elements, then so is \(H^*_G(\widehat{W};\mathbb Z_2)\). In addition, since \(\iota _{\widehat{W}}^*: H^*_G(\widehat{W};\mathbb Z_2)\rightarrow H^*(\widehat{W};\mathbb Z_2)\) is surjective, \(H^*(\widehat{W};\mathbb Z_2)\) is also generated by its degree-one elements. Then, by Theorem 1.3 (ii), \(\widehat{Q}\) is a mod 2 homology polytope. So by Fact-2, Q is mod 2 face-acyclic.\(\square \)

5.4 Proof of Theorem 1.5

Proof

We first prove the “if” part. Assume that there exists a regular \(\textrm{m}\)-involution \(\tau \) on W. By definition the fixed point set \(W^{\tau }\) of \(\tau \) is discrete, then so is \(W^{\mathbb Z^n_2}\subseteq W^{\tau }\). This implies that Q must have vertices. Let p be a vertex of Q and let \(F_1,\cdots , F_n\) be all the facets containing p. By the property of \(\lambda _W\),

$$\begin{aligned} e_1=\lambda _W(F_1),\cdots , e_n=\lambda _W(F_n) \end{aligned}$$

form a linear basis of \(\mathbb Z_2^n\) over \(\mathbb Z_2\). Then, since the \(\mathbb Z_2^n\)-action on W is locally standard, it is easy to see that only when \(g=e_1+\cdots +e_n\) could the fixed point set \(W^{\tau _g}\) be discrete. So we must have \(\tau =\tau _{e_1+\cdots +e_n}\), and in particular

$$\begin{aligned} W^{\tau }=W^{\tau _{e_1+\cdots +e_n}}=W^{\mathbb Z_2^n}. \end{aligned}$$

Hence

$$\begin{aligned} \dim _{\mathbb Z_2} H^*(W^{\mathbb Z_2^n};\mathbb Z_2) = \dim _{\mathbb Z_2} H^*(W^{\tau };\mathbb Z_2) =\dim _{\mathbb Z_2} H^*(W;\mathbb Z_2) \end{aligned}$$

where the second “\(=\)” is due to the assumption that \(\tau \) is an \(\textrm{m}\)-involution. So by Theorem 1.1, W is equivariantly formal. Then, Q is mod 2 face-acyclic by Theorem 1.3. In particular, every face of Q has a vertex and the 1-skeleton of Q is connected (by Lemma 2.2).

It remains to prove that the image of \(\lambda _W: \mathcal {F}(Q)\rightarrow \mathbb Z_2^n\) is exactly \(\{e_1,\cdots , e_n\}\). Indeed, take an edge e of Q whose vertices are p and \(p'\). So the n facets of Q that contain \(p'\) are \(F_1,\cdots , F_{i-1}, F'_i, F_{i+1},\cdots , F_n\) for some \(1\le i \le n\). Then, since \(\tau _{e_1+\cdots +e_n}\) is an \(\textrm{m}\)-involution, we must have

$$\begin{aligned} \lambda _W(F_1)+\cdots + \lambda _W(F_{i-1}) +\lambda _W( F'_i)+ \lambda _W(F_{i+1}) + \cdots +\lambda _W(F_n) = e_1+\cdots +e_n. \end{aligned}$$

This implies \(\lambda _W( F'_i)=e_i\). Then, since the 1-skeleton of Q is connected and every facet F of Q contains a vertex, we can iterate the above argument to prove that every \(\lambda _W(F)\) must take value in \(\{e_1,\cdots , e_n\}\).

Next, we prove the “only if” part. Suppose Q is mod 2 face-acyclic and the values of the characteristic function \(\lambda _W\) of Q consist exactly of a linear basis \(e_1,\cdots , e_n\) of \(\mathbb Z^n_2\). By Theorem 1.3 (i), W is equivariantly formal. So we have

$$\begin{aligned} \dim _{\mathbb Z_2} H^*(W^{\mathbb Z_2^n};\mathbb Z_2)=\dim _{\mathbb Z_2} H^*(W;\mathbb Z_2) \ \ \text {(by Theorem~1.1)}. \end{aligned}$$

On the other hand, our assumption on \(\lambda _W\) implies that the regular involution \(\tau =\tau _{e_1+\cdots +e_n}\) satisfies \(W^{\tau }=W^{\mathbb Z^n_2}\) which is a discrete set. Then, we have

$$\begin{aligned} \dim _{\mathbb Z_2} H^*(W^{\tau };\mathbb Z_2) = \dim _{\mathbb Z_2} H^*(W^{\mathbb Z_2^n};\mathbb Z_2)=\dim _{\mathbb Z_2} H^*(W;\mathbb Z_2). \end{aligned}$$

So \(\tau \) is a regular \(\textrm{m}\)-involution on W by definition. The theorem is proved.\(\square \)

Remark 1.4

If we do not assume a 2-torus manifold W to be locally standard, even if W admits a regular \(\textrm{m}\)-involution, W may not be equivariantly formal or locally standard. For example: let

$$\begin{aligned} S^2=\{(x_1,x_2,x_3)\in \mathbb R^3\,|\, x_1^2+x^2_2 + x^2_3 =1 \}. \end{aligned}$$

Define two involutions \(\sigma \) and \(\sigma '\) on \(S^2\) by

$$\sigma (x_1,x_2,x_3)=(-x_1,-x_2,x_3), \quad \sigma '(x_1,x_2,x_3) = (x_1,x_2,-x_3). $$

It is easy to see that \(\sigma \) is an \(\textrm{m}\)-involution on \(S^2\) with two isolated fixed points (0, 0, 1) and \((0,0,-1)\). But since the \(\mathbb Z^2_2\)-action on \(S^2\) determined by \(\sigma \) and \(\sigma '\) has no global fixed point, it is not equivariantly formal. We can also directly check that this \(\mathbb Z^2_2\)-action on \(S^2\) is not locally standard.

Finally, we propose some questions on weakly equivariantly formal 2-torus manifolds:

Question-3: Does there exist a weakly equivariantly formal 2-torus manifold which is not equivariantly formal?

Question-4: If a 2-torus manifold is weakly equivariantly formal, are there any restrictions on the topology and combinatorial structure of its orbit space?

Question-5: Whether or not a 2-torus manifold being weakly equivariantly formal is determined only by the topology and combinatorial structure of its orbit space?