1 Introduction

In this paper we have considered only smooth manifolds. However using the results of Milnor [14], Kister [8] and Kirby–Siebenmann [7] the arguments given for smooth manifolds can be carried over to topological manifolds.

Let \(f:M \rightarrow N\) be a continuous map of closed connected oriented manifolds of dimension \(n\). We aim to study the rational homotopy of \(M \times N \backslash \Gamma (f)\), where \(\Gamma (f)=\{(x,f(x))|x\in M\}\) is the graph of \(f\).

We refer the reader to [3] and [5] for standard definitions and terminologies of: Free commutative differential graded algebra (CDGA), (minimal) Sullivan CDGA, (minimal) Sullivan model for a CDGA and for a space X, homotopy between morphisms of CDGAs, formality of spaces.

Definition 1.1

([5], [4, 22], [12, 20]) A model for a continuous map \(f:X\rightarrow Y\) between connected topological spaces is a morphism of CDGAs, \(\psi _f:(A_Y,d_{A_Y})\rightarrow (A_X,d_{A_X})\), such that there exists a homotopy commutative diagram, where \((\Lambda W, d_W),(\Lambda V, d_V)\) are minimal models of \(Y\) and \(X\) respectively and the vertical arrows are quasi-isomorphisms:

$$\begin{aligned} \left. \begin{array}{lll} (A_Y,d_{A_Y})&{}\buildrel {\psi _f}\over \rightarrow &{} (A_X,d_{A_X})\\ \cong _q \;\uparrow \; \phi _W &{}&{} \; \cong _q \;\uparrow \; \phi _V \\ (\Lambda W, d_W) &{} \buildrel {\psi _{\Lambda f}}\over \rightarrow &{} (\Lambda V, d_V) \\ \; \cong _q \;\downarrow \; \rho _W &{}&{} \; \cong _q \;\downarrow \; \rho _V \\ (A_{PL}(Y), d_Y) &{} \buildrel {A_{PL}(f)}\over \rightarrow &{} (A_{PL}(X), d_X). \end{array}\right. \end{aligned}$$

A continuous map \(f:X\rightarrow Y\) between formal spaces is said to be formal (also referred to as formalizable in the literature) if \(f^*:(H^*(Y; \mathbb {Q}), 0) \rightarrow (H^*(X; \mathbb {Q}), 0)\) is a model for \(f\).

Many authors have studied rational homotopy type of maps and their formality; see e.g. [4, 13, 1618, 22].

These authors have studied spaces \(C(f)\) closely associated with \(f\), like the mapping cone of \(f\), the cofibre of \(f\), the homotopy fibre of \(f\) and have recorded results related to problems of following types:

Problem 1.2

  1. 1.

    Find out if the rational homotopy of \(C(f)\) is determined by the rational homotopy of \(f\).

  2. 2.

    Given that \(f\) is formal, determine if \(C(f)\) is necessarily formal.

We study the following problems of a similar nature:

Problem 1.3

  1. 1.

    Find out if the rational homotopy of \(M \times N \backslash \Gamma (f)\), where \(\Gamma (f)\) is the graph of \(f\), is determined by the rational homotopy of \(f\).

  2. 2.

    Given that \(f\) is formal, determine if \(M \times N \backslash \Gamma (f)\) is necessarily formal.

The motivation of this paper comes from the work of Lambrechts and Stanley [9] on the rational homotopy type of configuration spaces of two points, and the observation that if \(f\) is the identity map \(1_M:M \rightarrow M\) then, the graph of \(f\), \(\Gamma (f) = \Delta \), the diagonal, and \(M \times M \backslash \Gamma (f) = M \times M \backslash \Delta = F(M, 2)\), the configuration space of two points. We begin with the following immediate observations:

  1. 1.

    If \(f: {\mathbb {R}}^n \rightarrow {\mathbb {R}}^n\) is any continuous map then \( {\mathbb {R}}^n \times {\mathbb {R}}^n \backslash \Gamma (f)\) is homeomorphic to \(F({\mathbb {R}}^n, 2)\) under the homeomorphism \((u, v) \mapsto (u, v-f(u)+u)\).

  2. 2.

    For a topological space \(X\) if \(f:X \rightarrow X\) is a homeomorphism then \(X \times X \backslash \Gamma (f)\) is homeomorphic to \( X \times X \backslash \Delta = F(X, 2)\) under the homeomorphism \((x, y) \mapsto (x, f^{-1}(y))\).

  3. 3.

    Suppose that \(X = S^2\vee S^3\), \(f:X\rightarrow X\) is the constant map \(x\mapsto (0,0,1)\) and \(g:X\rightarrow X\) is the constant map \(x\mapsto (0,0,0,-1)\). Then \(X \times X\setminus \Gamma (f)\simeq X \times (\mathbb {R}^2\vee S^3)\) whereas \(X \times X\setminus \Gamma (g)\simeq X \times (S^2 \vee \mathbb {R}^3)\). Therefore they are not homotopy equivalent. Thus for an arbitrary space \(X\) and a continuous map \(f:X \rightarrow X\), \(X \times X \backslash \Gamma (f)\) need not even be homotopy equivalent to \(F(X, 2)\).

Lambrechts and Stanley [9, 10] proved that if \(M\) is a simply-connected closed manifold such that \(H^2(M;{\mathbb {Q}}) = 0\) then the rational homotopy type of \(F(M,2)\) depends only on the rational homotopy type of \(M\). They also proved in [9] that if \(M\) is a 2-connected closed manifold, then \(F(M,2)\) is formal if and only if \(M\) is formal by constructing explicitly a CDGA model for \(F(M,2)\) out of a differential Poincaré duality algebra model of \(M\) and a model of the diagonal map.

We show that starting from a CDGA model of \(f: M \rightarrow N\) one can construct a CDGA model of \((1_M \times f) \circ \Delta _M: M \rightarrow M \times N\), where \(\Delta _M (x) = (x, x)\) for \(x\in M\). This allows us by Corollary 1.5 of [10], to conclude that if we take \(M, \;N\) to be simply connected and \(H^2(N;\mathbb {Q}) =0\) then the rational homotopy type of \(M \times N \setminus \Gamma (f)\) is determined by the rational homotopy class of \(f\):

Theorem 1.4

Let \(f:M \rightarrow N\) be a continuous map of closed connected oriented manifolds of dimension \(n\) such that \(H^1(M;\mathbb {Q}) = 0 = H^1(N;\mathbb {Q}) = H^2(N;\mathbb {Q}),\) then a CDGA-model of \(M \times N\setminus \Gamma (f)\) can be explicitly determined out of any CDGA-model of \(f\).

Corollary 1.5

If \(f:M \rightarrow N\) is a continuous map of closed simply connected manifolds of dimension \(n\) such that \(H^2(N;\mathbb {Q}) =0,\) then the rational homotopy type of \(M \times N \backslash \Gamma (f)\) depends only on the rational homotopy class of \(f\).

We relativize results of [9] by defining a class \(\Delta _{\psi }\) depending on a morphism \(\psi \) of differential Poincaré duality algebras similar to the diagonal class \(\Delta \) in [9]. We know from Theorem 1.1 of [11] that if \(M, \; N\) are closed connected oriented manifolds with \(H^1(M; {\mathbb {Q}}) = 0 = H^1(N; {\mathbb {Q}}),\) then they admit differential Poincaré duality algebra models. We also know from Proposition 1 of [2] that if \(f:M \rightarrow N\) is a continuous map of simply connected manifolds of dimension \(n \ge 7\) (it is a conjecture that this dimensional restriction can be removed) such that \(H^2(f)\) is injective, then \(f\) admits a Sullivan model \(\psi _f:(A_N,d_{A_N})\rightarrow (A_M,d_{A_M})\), where \((A_M,d_{A_M}),\;(A_N,d_{A_N})\) are Poincaré duality algebras. Thus assuming the existence of such a model of \(f\) we can construct a specific CDGA model of \(M \times N\!\setminus \!\Gamma (f)\) from the model of \(f\):

Theorem 1.6

Let \(f:M \rightarrow N\) be a continuous map of closed connected oriented manifolds of dimension \(n\) such that \(H^1(M;\mathbb {Q}) = 0 = H^1(N;\mathbb {Q}) = H^2(N;\mathbb {Q}).\) If \(\psi _f:(A_N,d_{A_N})\rightarrow (A_M,d_{A_M})\) is a model of \(f,\) where \((A_M,d_{A_M})\), \((A_N,d_{A_N})\) are oriented differential Poincaré duality algebras, \(\{a_i\}_{1\le i \le l}\) a homogeneous basis of \(A_N\) and \(\{a^*_i\}_{1\le i \le l}\) its Poincaré dual basis

$$\begin{aligned}\Delta = \sum _{i=1}^{l_N} (-1)^{\deg \; (a_i)} a_i \otimes a_i^* \in (A_N \otimes A_N)^n, \end{aligned}$$

is the diagonal class (defined in [9, p. 1030) and \({\Delta }_{\psi _f} := (\psi _f \otimes 1)(\Delta )\), called the class of \({\psi _f}\), then the ideal \(({\Delta }_{\psi _f}) = {\Delta }_{\psi _f}.(A_M \otimes A_N)\) is a differential ideal of \(A_M \otimes A_N\), and the quotient CDGA

$$\begin{aligned}\left( \frac{A_M \otimes A_N}{({\Delta }_{\psi _f})},\,\, \overline{d_{A_M} \otimes d_{A_N}}\right) \end{aligned}$$

is a CDGA model of \(M \times N \backslash \Gamma (f)\).

This allows us to prove that if \(f\) is formal then \(M \times N\!\setminus \!\Gamma (f)\) is also formal.

Corollary 1.7

Let \(f:M \rightarrow N\) be a formal map of closed connected oriented formal manifolds of dimension \(n\) such that \(H^1(M;\mathbb {Q}) = 0 = H^1(N;\mathbb {Q}) = H^2(N;\mathbb {Q}),\) then \(M \times N\setminus \Gamma (f)\) is a formal space.

As the first major step in this paper we compute the cohomology algebra \(H^*(M \times N \backslash \Gamma (f))\) using the Leray spectral sequence as described in §2 of Totaro [21]. We analyze at depth the cohomology class \(\mu _f\), as described in Chapter 30 of Greenberg and Harper [6].

Let \(\tilde{\mu } \in H^n(M \times N, M \times N\!\setminus \!\Gamma (f); \mathbb {Z})\) be the Thom class as in Lemma 2.1, and \(\tilde{\mu '} = j^*(\tilde{\mu }) \in H^n(M \times N; \mathbb {Z})\), where \(j: M \times N \rightarrow (M \times N, M \times N \setminus \Gamma (f))\) is the natural injection. Let us now take cohomology with rational coefficients (or coefficients in any field). Let \(\{b_i\}_{1 \le i \le {l_N}}\) be a homogeneous basis of \(H^*(N; {\mathbb {Q}})\), and let \(\{b_i^*\}_{1 \le i \le {l_N}}\) be its Poincaré dual basis, that is \(\langle b_i \cup b_j^*, [N]\rangle = \delta _{ij}\), where \([N]\) is the fundamental homology class of \(N\). Let \(\mu '_N := \sum _{i=1}^{l_N} (-1)^{\deg \; (b_i)} b_i \times b_i^* \in (H^*(N \times N; {\mathbb {Q}}))^n\) be the diagonal class of \(N\), and let \(\mu _f := (f \times id_N)^*(\mu '_N)\), called the cohomology class of the graph of \(f\) (see p. 284 of [6]). Then there is a unique class \(\Delta _f \in (H^*(M; {\mathbb {Q}})\otimes H^*(N; {\mathbb {Q}}))^n\) which maps to \(\mu _f\) under the isomorphism of the Künneth theorem \((H^*(M; {\mathbb {Q}})\otimes H^*(N; {\mathbb {Q}}))^n \xrightarrow {\cong } (H^*(M \times N; {\mathbb {Q}}))^n\), we call \(\Delta _f\) the class of \(f\).

Theorem 1.8

Let \(f:M \rightarrow N\) be a continuous map of closed oriented manifolds of dimension \(n\ge 2\) and \(\Gamma (f)\) be the graph of \(f\). Then we have the following isomorphism of rings:

$$\begin{aligned} H^*(M \times N \backslash \Gamma (f); \mathbb {Z}) \cong \frac{H^*(M \times N; \mathbb {Z})}{(\tilde{\mu }')}, \end{aligned}$$

where \((\tilde{\mu }')\) denotes the ideal \(H^*(M \times N; \mathbb {Z}) \cup \tilde{\mu }'\). For coefficients in \(\mathbb {Q}\), and \(\Delta _f =\sum ^{l_N}_{i=1}(-1)^{\deg \;(b_i)}f^*(b_i)\otimes b_i^{*}\), we have the following isomorphism of algebras:

$$\begin{aligned} H^*(M \times N\!\setminus \!\Gamma (f); {\mathbb {Q}}) \cong \frac{H^*(M \times N; {\mathbb {Q}})}{(\mu _f)} \cong \frac{H^*(M; {\mathbb {Q}})\otimes H^*(N; {\mathbb {Q}})}{(\Delta _f)}. \end{aligned}$$

The paper is arranged as follows. In §2 we determine the cohomology algebra of \(M \times N \backslash \Gamma (f)\). In §3 we adapt for a continuous map \(f:M\rightarrow N\) some results of [9] and [10] to construct a CDGA model of \(M \times N \backslash \Gamma (f)\), and prove Theorem 1.4 and Corollary 1.5. In §4 we prove several results about differential Poincaré duality algebra associated to a morphism \(\psi \) by adapting similar results of [9], proved in the absolute case. These results and some results of [10] are used for the proofs of Theorem 1.6 and Corollary 1.7. In the final §5 we give some examples and applications of our results.

2 The cohomology algebra of \(M \times N \backslash \Gamma (f)\)

In this section we prove Theorem 1.8.

We assume that \(f:M \rightarrow N\) is a continuous map of closed connected oriented manifolds of dimension \(n\) with graph \(\Gamma (f)\). Since \(\Gamma (f) \subset M \times N\) is an embedding, one can derive the following analogue of Corollary 11.2 of Milnor-Stasheff [15] (and also of Corollary (30.2) of Greenberg-Harper [6]) in a similar fashion. We follow the notations which preceded the statement of Theorem 1.8 in §1.

Lemma 2.1

There is a Thom class \(\tilde{\mu } \in H^n(M \times N, M \times N \backslash \Gamma (f); {\mathbb {Z}})\) associated to the oriented normal bundle of the embedding \(\Gamma (f) \subset M \times N\).

We next prove the following analogue of Corollary (30.3) of [6]:

Theorem 2.2

Let \([N] \in H_n(N; {\mathbb {Z}})\) be the fundamental homology class of \(N\), \(j^*:H^n(M \times N, M \times N \backslash \Gamma (f); {\mathbb {Z}}) \rightarrow H^n(M \times N; {\mathbb {Z}})\) be the homomorphism in cohomology induced by the map \(j:M \times N \hookrightarrow (M \times N, M \times N \backslash \Gamma (f))\) and \(\tilde{\mu }' = j^* (\tilde{\mu })\), then \(\tilde{\mu }' / [N] = 1\).

Proof

For \(x \in M\), consider the commutative diagram:

where \(\tilde{i_x}:(N, N \backslash f(x)) \rightarrow (M \times N, M \times N \backslash \Gamma (f))\) is defined by \(\tilde{i_x} (x') = (x, x'), \forall x' \in N\), \(\tilde{i_x}: N \rightarrow M \times N\) its restriction to \((N, \emptyset )\) and \(j_{f(x)}\!:\!N\rightarrow (N,N\!\setminus \!f(x))\) is the natural injection. If \(s: N \rightarrow N^0\) is a section of the orientation sheaf \(N^0\) over \(N\), then

$$\begin{aligned}&1 = [s(f(x)), \tilde{i_x}^* (\tilde{\mu })] = [{j_{f(x)}}_*[N], \tilde{i_x}^* (\tilde{\mu })] = [[N], j_{f(x)}^* \tilde{i_x}^* (\tilde{\mu })] \\&\quad = [[N], (\tilde{i_x} \circ j_{f(x)})^* (\tilde{\mu })] = [[N], (j \circ \tilde{i_x} )^* (\tilde{\mu })] = [ [N], \tilde{i_x}^* \circ j^*(\tilde{\mu })] \\&\quad = [\tilde{i_x}_*[N],j^*(\tilde{\mu })] = [\tilde{i_x}_*[N], \tilde{\mu }']. \end{aligned}$$

Since \(N \cong \{x\} \times N\), if \(\bar{x}\) is the homology class of the zero cycle \(x\), \(\tilde{i_x}_*([N]) = \bar{x} \times [N]\). So from the above expression we get

\(1 = [\tilde{i_x}_*[N], \tilde{\mu }'] = [\bar{x} \times [N], \tilde{\mu }'] = [\bar{x}, \tilde{\mu }' / [N]] \) for all \(x \in M\), which proves the result. \(\square \)

We now prove the following analogue of Lemma 11.8 of [15] (and also of Lemma (30.5) of [6]):

Theorem 2.3

If \(a \in H^*(N; {\mathbb {Z}})\), then \((f^*(a) \times 1) \cup \tilde{\mu }' = (1 \times a) \cup \tilde{\mu }'. \)

Proof

Let \(V_{\Gamma (f)}\) be a tubular neighborhood of the graph \(\Gamma (f)\) in \(M \times N\). So \(\Gamma (f)\) is a deformation retract of \(V_{\Gamma (f)}\). Let \(i_{\Gamma (f)}:\Gamma (f) \rightarrow V_{\Gamma (f)}\) be the inclusion map and \(r_{\Gamma (f)}:V_{\Gamma (f)}\rightarrow \Gamma (f)\) be the retraction map. Then \(r_{\Gamma (f)} \circ i_{\Gamma (f)} = 1_{\Gamma (f)}\) and \(i_{\Gamma (f)} \circ r_{\Gamma (f)} \simeq 1_{V_{\Gamma (f)}}\). Consider the projections \(p_1:M \times N \rightarrow M\) and \(p_2:M \times N \rightarrow N\). Since \(f \circ p_1\) and \(p_2\) coincide on \(\Gamma (f)\), we have \((f \circ p_1)\mid _{V_{\Gamma (f)}} \circ \; i_{\Gamma (f)} \circ r_{\Gamma (f)}= p_2 \mid _{V_{\Gamma (f)}}\circ \;i_{\Gamma (f)} \circ r_{\Gamma (f)} \). So the \((f \circ p_1) \mid _{V_{\Gamma (f)}} \simeq p_2 \mid _{V_{\Gamma (f)}}\). Therefore the cohomology classes \((f \circ p_1)^*(a) = p_1^* \circ f^*(a) = f^*(a) \times 1\) and \(p_2^*(a) = 1 \times a\) have the same image under the restriction homomorphism \(H^i(M \times N; {\mathbb {Z}}) \rightarrow H^i(V_{\Gamma (f)}; {\mathbb {Z}})\). Now using the commutative diagram

it follows that \((f^*(a) \times 1) \cup \tilde{\mu } = (1 \times a) \cup \tilde{\mu }.\) So \(j^*((f^*(a) \times 1) \cup \tilde{\mu }) = j^*((1 \times a) \cup \tilde{\mu }),\; j \) as defined earlier. By the properties of mixed cup products \((f^*(a) \times 1) \cup j^*(\tilde{\mu }) = (1 \times a) \cup j^*(\tilde{\mu }). \) So \((f^*(a) \times 1) \cup \tilde{\mu }' = (1 \times a) \cup \tilde{\mu }'. \) \(\square \)

The next theorem is an analogue of Theorem 11.11 of [15] (and also of Proposition (30.18) of [6]):

Theorem 2.4

Let \(\{b_i\}^{l_N}_{i=1}\) be a homogeneous basis of \(H^*(N; {\mathbb {Q}})\), and let \(\{b_i^{*}\}^{l_N}_{i=1}\) be its Poincaré dual basis, then the cohomology class \(\tilde{\mu }'\) is given by

$$\begin{aligned} \tilde{\mu }' = \sum _{i=1}^{l_N} (-1)^{\deg \, (b_i)}f^*(b_i) \times b_i^{*}, \end{aligned}$$

where deg \((b_i)=k\) if \(b_i\in H^k(N; {\mathbb {Q}}).\)

Proof

Using the Künneth formula we can write \(\tilde{\mu }' \in H^n(M \times N ; {\mathbb {Q}})\) as

$$\begin{aligned} \tilde{\mu }' = c_1 \times b_1^{*} +...+c_{l_N} \times b^{*}_{l_N}, \end{aligned}$$

where \(c_1,\ldots , c_{l_N}\) are certain well defined cohomology classes in \(H^*(M; {\mathbb {Q}})\) with \(\deg \, (b_i) + \deg \, (c_i)= n\). We apply the homomorphism \(/[N]\) on both sides of the identity \((f^*(a) \times 1) \cup \tilde{\mu }' = (1 \times a) \cup \tilde{\mu }',\) where \(a \in H^*(N; {\mathbb {Q}})\). By Property 4 of slant product on p. 288 of Spanier [19] and by our Theorem 2.2 we get

$$\begin{aligned} ((f^*(a) \times 1) \cup \tilde{\mu }') / [N] = f^*(a) \cup (\tilde{\mu }'/ [N]) = f^*(a). \end{aligned}$$
(1)

Also,

$$\begin{aligned} ((1 \times a) \cup \tilde{\mu }')/[N]&= (1 \times a) \cup \sum _{i=1}^{l_N} c_i \times b_i^{*} /[N] = \sum _{i=1}^{l_N} (-1)^{\deg \, (a) \deg \, (c_i)} c_i \nonumber \\&\times (a \cup b_i^{*}) /[N] \nonumber \\&= \sum _{i=1}^{l_N} (-1)^{\deg \, (a) \deg \, (c_i)} c_i \langle a \cup b_i^{*}, [N]\rangle \end{aligned}$$
(2)

Using equations (1) and (2) above and substituting \(b_j\) for \(a\), we get

$$\begin{aligned} f^*(b_j) = \sum _{i=1}^{l_N} (-1)^{\deg \, (b_j) \deg \, (c_i)} c_i \langle b_j \cup b_i^{*}, [N]\rangle = (-1)^{\deg \, (b_j) \deg \, (c_j)} c_j, \end{aligned}$$

but \(\deg \, (b_j) = \deg \, (c_j)\), so we get that

$$\begin{aligned} (-1)^{\deg \, (b_j)} f^*(b_j) \!=\! (-1)^{\deg \, (b_j)}(-1)^{\deg \, (b_j) \deg \, (b_j)} c_j \!=\! (-1)^{\deg \, (b_j) (1 + \deg \, (b_j)} c_j = c_j. \end{aligned}$$

This proves the theorem. \(\square \)

Let \([\Gamma (f)] = (1_M \times f)_* \circ {\Delta _M}_* ([M])\). Let \(\mu _N \in H^n(N \times N, N \times N \backslash \Gamma (1_N); {\mathbb {Z}})\) be the Thom class of the oriented normal bundle of the embedding \( \Gamma (1_N) \hookrightarrow N \times N\), let \(\mu _N' \in H^n(N \times N; {\mathbb {Z}})\) be its image under the homomorphism in cohomology induced by \(j_N :N \times N \rightarrow (N \times N, N \times N \backslash \Gamma (1_N))\) and \(\mu _f = (f \times 1_N)^* (\mu _N') \in H^n(M \times N; {\mathbb {Z}})\).

Remark 2.5

  1. 1.

    When coefficients belong to \({\mathbb {Q}}\) (or any field) by Theorem 11.11 of [15] (or by Proposition (30.18) of [6]), if \(\{a_i\}_{i=1}^{l_M}\) is a homogeneous basis of \(H^*(M;\mathbb {Q})\) and \(\{a_i^*\}_{i=1}^{l_M}\) its Poincaré dual basis, we have \(\mu _M' = \sum _{i=1}^{l_M} (-1)^{\deg \, (a_i)} a_i \times a_i^{*}\), and \(\mu _N' = \sum _{i=1}^{l_N} (-1)^{\deg \, (b_i)} b_i \times b_i^{*}\). Therefore, we see using Theorem 2.4 , that in rational cohomology, \(\mu _f = \tilde{\mu }'\).

  2. 2.

    By Exercises (30.15) and (30.17) of [6] we have \([M \times M] \cap \mu _M' = {\Delta _M}_*([M]) = [\Gamma (1_M)]\), and \([N \times N] \cap \mu _N' = {\Delta _N}_*([N]) = [\Gamma (1_N)]\). Using this and Part 1 of this remark one can prove the following statement which is the assertion of the Exercise (30.22) of [6]:

    $$\begin{aligned}{}[M] \times [N] \cap \tilde{\mu }' = [M] \times [N] \cap \mu _f = [\Gamma (f)] \end{aligned}$$
    (A)

We are now ready to prove Theorem 1.8.

Proof of Theorem.1.8

Let \(h:M \times N\!\setminus \!\Gamma (f)\hookrightarrow M \times N\) denote the inclusion map. The Leray spectral sequence for this inclusion has the form

$$\begin{aligned} E_2^{i, j} = H^i(M\times N; R^jh_*\mathbb {Z}) \Rightarrow H^{i+j}(M\times N \setminus \Gamma (f);\mathbb {Z}). \end{aligned}$$

Here \(R^jh_*\mathbb {Z}\) is the Leray sheaf on \(M \times N\) associated with the presheaf \( U \rightarrow H^j(U \cap (M \times N\!\setminus \!\Gamma (f));\mathbb {Z})\) where \(U\) runs over the family of all open subsets of \(M \times N\). The stalk of \(R^jh_*\mathbb {Z}\) at a point \(\bar{x}=(x_1,x_2) \in M \times N\) is described below.

  1. 1.

    If \(x_2 \ne f(x_1)\), we can choose small coordinate neighborhoods \(U_1\) and \(U_2\) of \(x_1\) and \(x_2\) in \(M\) and \(N\) respectively, such that \(f(U_1)\cap U_2=\emptyset \). If \(U = U_1\times U_2\), then \(U \subset M\times N\!\setminus \!\Gamma (f)\). Therefore, for every \(\bar{x}=(x_1,x_2)\in M\times N, x_2 \ne f(x_1)\), \((R^jh_*\mathbb {Z})_{\bar{x}} = H^j(U \cap (M\times N\!\setminus \!\Gamma (f));\mathbb {Z}) = H^j(U;\mathbb {Z}) = H^j(U_1\times U_2;\mathbb {Z}) = 0 \) if \(j\ne 0\), and \((R^0h_*\mathbb {Z})_{\bar{x}} = {\mathbb {Z}}\).

  2. 2.

    If \(x_2 = f(x_1)\), we can choose small coordinate neighborhoods \(U_1\) and \(U_2\) of \(x_1\) and \(x_2\) in \(M\) and \(N\) respectively, such that \(f(U_1)\subset U_2\). For \(U = U_1 \times U_2\), \((R^jg_*\mathbb {Z})_{\bar{x}} = H^j(U \cap (M\times N\!\setminus \!\Gamma (f));\mathbb {Z}) = H^j(U\!\setminus \!\Gamma (f);\mathbb {Z}) = H^j(U_1 \times U_2 \setminus \Gamma (f) ;\mathbb {Z})\). Let \(h_1: U_1 \xrightarrow {\cong } \mathbb {R}^n\; \text{ and } \; h_2: U_2 \xrightarrow {\cong } \mathbb {R}^n\), be the coordinate charts of \(U_1 \;\text{ and }\; U_2\) respectively. Consider the composite \(\hat{f} = h_2\circ f \circ h_1^{-1}:\mathbb {R}^n \rightarrow \mathbb {R}^n\), then \(\Gamma (\hat{f}) = h_1\times h_2 (\Gamma (f)|_{U_1\times U_2})\). Therefore, \((R^jh_*\mathbb {Z})_{\bar{x}}= H^j(U_1 \times U_2\!\setminus \!\Gamma (f);\mathbb {Z}) \cong H^j((\mathbb {R}^n \times \mathbb {R}^n)\!\setminus \!\Gamma (\hat{f});\mathbb {Z})\). But as observed in the introduction \(\mathbb {R}^n\times \mathbb {R}^n\!\setminus \!\Gamma (\hat{f})\) is homeomorphic to \(F(\mathbb {R}^n, 2)\), so for every \(\bar{x}=(x_1,x_2)\in M\times N, x_2 = f(x_1)\)

    $$\begin{aligned} (R^jh_*\mathbb {Z})_{\bar{x}}= H^j(F(\mathbb {R}^n, 2); \mathbb {Z}) = \left\{ \begin{array}{ll} \mathbb {Z}&{} \quad \text{ if } \ j = 0, n-1\\ 0 &{} \quad \text{ otherwise } \end{array} \right. \end{aligned}$$

This shows that the Leray sheaf \((R^{n-1}h_*\mathbb {Z})\) is supported and is locally constant along the graph \(\Gamma (f)\) with stalks \({\mathbb {Z}}\), and \((R^0h_*\mathbb {Z})\) is locally constant on \(M \times N\) with stalks \({\mathbb {Z}}\). Since \(M\), \(N\) and hence \(\Gamma (f)\) are orientable, we have

$$\begin{aligned} H^{i}(M \times N; R^jh_*\mathbb {Z})= \left\{ \begin{array}{ll} H^{i}(M \times N; \mathbb {Z}) &{} \quad \text{ if } \ j = 0,\\ H^{i}(\Gamma (f); \mathbb {Z}) &{} \quad \text{ if }\ j = n-1,\\ 0 &{} \quad \text{ otherwise } \end{array} \right. \end{aligned}$$

Hence, the \(E_2\) terms of Leray spectral sequence take the form:

$$\begin{aligned} \begin{array}{ccccc} 0&{}0&{}\cdots &{}0&{}\cdots \\ E_2^{0,n-1}=H^0(\Gamma (f);\mathbb {Z}) &{} E_2^{1,n-1}=H^1(\Gamma (f);\mathbb {Z}) &{}\cdots &{} E_2^{n,n-1}=H^n(\Gamma (f);\mathbb {Z}) &{}\cdots \\ 0&{}0&{}\cdots &{}0&{}\cdots \\ 0&{}0&{}\cdots &{}0&{}\cdots \\ 0&{}0&{}\cdots &{}0&{}\cdots \\ 0&{}0&{}\cdots &{}0&{}\cdots \\ E_2^{0,0}=H^0(M\times N;\mathbb {Z}) &{} E_2^{1,0}=H^1(M\times N;\mathbb {Z}) &{}\cdots &{} E_2^{n,0}=H^n(M\times N;\mathbb {Z}) &{} \cdots \\ \end{array} \end{aligned}$$

and \(E_n\) terms coincide with \(E_2\) terms:

figure a

We now determine \(d_n\) on the \(E_n\) page. Note that the differential \(d_n\) is 0 on the bottom row, since it maps each row to a lower row. Let \(i:\Gamma (f)\rightarrow M\times N\) be the inclusion map and let \(\pi :M\times N\rightarrow \Gamma (f)\) be defined by \(\pi (x_1,x_2)=(x_1,f(x_1))\). Then \(\pi \circ i=1_{\Gamma (f)}\). Therefore, by the functoriality of cohomology, the homomorphisms induced in cohomology, \(\pi ^*\) and \(i^*\) are, respectively, injective and surjective. Hence the long cohomology exact sequence of the pair \((M \times N, \Gamma (f))\) gives rise to the following split short exact sequence:

$$\begin{aligned} 0 \rightarrow H^*(M \times N, \Gamma (f); {\mathbb {Z}})\xrightarrow {j_1^*} H^*(M \times N; {\mathbb {Z}}) \underset{\pi ^*}{\buildrel {i^*}\over \rightleftharpoons } (H^*(\Gamma (f); {\mathbb {Z}}))\rightarrow 0, \end{aligned}$$

where \(j_1:M \times N \rightarrow (M\times N,\Gamma (f))\) is the natural injection.

As \(\Gamma (f)\) is a smooth submanifold with an orientable normal bundle in the smooth manifold \(M \times N\), the differentials \(d_n\) in the above Leray spectral sequence originating from the \((n-1)^{th}\) row are Gysin maps. If \(z \in H^i(M \times N; {\mathbb {Z}}), \; i^*(z)\in H^i(\Gamma (f); {\mathbb {Z}})\) and we have \(d_n(i^*(z)) = \pi ^*(i^*(z)) \cup \tilde{\mu }'\). From the above split exact sequence we see that \(z - \pi ^*(i^*(z)) = j_1^*(w)\), for some element \(w \in H^i(M \times N, \Gamma (f); {\mathbb {Z}})\). Recall that \(\tilde{\mu }' = j^*(\tilde{\mu })\), the image of the Thom class \(\tilde{\mu } \in H^*(M \times N, M \times N \setminus \Gamma (f); {\mathbb {Z}})\), where \(j : M \times N \rightarrow (M \times N, M \times N\!\!\setminus \!\!\Gamma (f))\) is the natural injection. So by Property 8 on p. 251 of [19] and the fact that \(w \cup \tilde{\mu } = 0\) (since the mixed cup product \(\cup : H^{i}(M \times N,\Gamma (f);\mathbb {Z}) \otimes H^{n} (M \times N,M \times N \setminus \Gamma (f);\mathbb {Z}) \rightarrow H^{i+n}(M \times N,\Gamma (f)\cup M \times N \setminus \Gamma (f) ;\mathbb {Z}) = H^{i+n}(M \times N, M \times N;\mathbb {Z}) = 0\) has its image in a zero module) we get that \((z - \pi ^*(i^*(z))) \cup \tilde{\mu }' = j_1^*(w) \cup \tilde{\mu }' = j_1^*(w) \cup j^*(\tilde{\mu }) = 0\). This means, in particular, that \(d_n(1) = \tilde{\mu }'\).

If \(k: M \rightarrow \Gamma (f)\) denotes the mapping \(x \mapsto (x,f(x))\) and \(\pi _M:M \times N \rightarrow M\) be the projection onto \(M\), then \(k^*\) is an isomorphism and \(\pi ^*=\pi _M^*\circ k^* \). Thus \(d_n(i^*(z)) = \pi _M^*(k^*(i^*(z))) \cup \tilde{\mu }' = (k^*(i^*(z)) \times 1) \cup \tilde{\mu }'\). But by Property 4 of slant products on p. 288 of [19], \(\{(k^*(i^*(z)) \times 1)\cup \tilde{\mu }'\}/[N] = (k^*(i^*(z))\cup (\tilde{\mu }'/[N]) = k^*(i^*(z))\), as \(\tilde{\mu }'/[N] = 1\) by Theorem 2.2 above. So we have: if \(i^*(z)\ne 0\) then \((k^*(i^*(z))\times 1) \cup \tilde{\mu }'\ne 0\). Therefore \(d_n\) is injective on the \((n-1)^{th}\)-row of \(E_n = E_2\).

This yields \(E_{n+1}^{i,j} = E_{\infty }^{i,j} = 0\) for \(j \ne 0\), and \(E^{i,0}_{\infty } = E^{i,0}_{n+1} = \frac{E^{i,0}_{n}}{{{\mathrm{im}}}d_n} = \frac{E^{i,0}_{2}}{{{\mathrm{im}}}d_n} = \frac{H^i(M \times N; {\mathbb {Z}})}{{{\mathrm{im}}}d_n}\) for all \(i\). But, for \(0 \le i \le n-1\), we have \(E_{\infty }^{i,0} = E_{n+1}^{i,0} = E_{n}^{i,0} = E_{2}^{i,0} = H^i(M \times N; {\mathbb {Z}})\), and for \(n \le i \le 2n\), \(E_{n+1}^{i,0} = E_{\infty }^{i,0} = \frac{H^i(M \times N; {\mathbb {Z}})}{{{\mathrm{im}}}d_n}\).

Thus

$$\begin{aligned} H^*(M \times N\!\!\setminus \!\!\Gamma (f); {\mathbb {Z}}) \cong \frac{H^*(M \times N; {\mathbb {Z}})}{H^*(M \times N; {\mathbb {Z}}) \cup \tilde{\mu }'} \cong \frac{H^*(M \times N; {\mathbb {Z}})}{(\tilde{\mu }')}, \end{aligned}$$

as rings, where \((\tilde{\mu }')\) denotes the ideal of \(H^*(M \times N; {\mathbb {Z}})\) generated by \(\tilde{\mu }'\).

If we take cohomology with coefficients in \(\mathbb {Q}\), then by an application of the Künneth theorem \(H^*(M; {\mathbb {Q}}) \otimes H^*(N; {\mathbb {Q}}) \xrightarrow {\cong } H^*(M \times N; {\mathbb {Q}})\). Hence there is a unique class \(\Delta _f =\sum ^n_{i=1}(-1)^{\deg \,(b_i)}f^*(b_i)\otimes b_i^{*}\) in \(H^*(M; {\mathbb {Q}}) \otimes H^*(N; {\mathbb {Q}})\) which maps to \(\tilde{\mu }' = \sum ^N_{i=1}(-1)^{\deg \,(b_i)} f^*(b_i)\times b_i^{*} = \mu _f\) in \(H^*(M \times N; {\mathbb {Q}})\) under this isomorphism, and therefore

$$\begin{aligned} H^*(M \times N \setminus \Gamma (f); {\mathbb {Q}}) \cong \frac{H^*(M \times N; {\mathbb {Q}})}{(\mu _f)} \cong \frac{H^*(M; {\mathbb {Q}})\otimes H^*(N; {\mathbb {Q}})}{(\Delta _f)}, \end{aligned}$$

as algebras. \(\square \)

A similar argument can be given when \(\mathbb {Q}\) is replaced by an arbitrary field.

3 A CDGA model of \( M \times N \backslash \Gamma (f)\)

In this section we prove Theorem 1.4 and Corollary 1.5 by constructing a CDGA model of \((1_M \times f) \circ \Delta _M: M \rightarrow M\times M \rightarrow M \times N\) from a CDGA model of \(f: M \rightarrow N\). (We refer the reader to [9], [10], [3] and [5] for necessary definitions, notations and results leading to the construction of a CDGA model of the configuration space of two points \(F(M,2)\)).

Proof of Theorem 1.4

We first note that a CDGA-model \((A_N, d_N)\xrightarrow {\psi _f} (A_M, d_M)\) of \(f\) determines a CDGA-model \((A_M \otimes A_N, d_{A_M} \otimes d_{A_N})\) \(\xrightarrow {1_{A_M} \otimes \psi _f}\) \((A_M \otimes A_M, d_{A_M} \otimes d_{A_M})\) of \(1_M \times f\), whose verification is left to the reader. This together with the fact that a CDGA-model of \(M\) determines a CDGA-model of \(\Delta _M \) (see, e.g. Example 2.48 on p.73 of [5]) yields the conclusion that a CDGA-model of \(f\) determines a CDGA-model \((A_M \otimes A_N) \xrightarrow {1_{A_M} \otimes \psi _f} (A_M \otimes A_M)\xrightarrow {\nu _{A_M}} A_M \), defined by \(x\otimes y \mapsto (\nu _{A_M} \circ (1_{A_M} \otimes \psi _f)) (x\otimes y) = x.\psi _f(y),\) of \((1_M \times f) \circ \Delta _M \).

By our hypotheses \(H^1(M;\mathbb {Q}) = 0\) and \(H^1(N;\mathbb {Q}) = 0 = H^2(N;\mathbb {Q}),\) it follows that for the embedding \((1_M \times f) \circ \Delta _M : M \rightarrow M \times N\), which actually embeds \(M\) as the graph \(\Gamma (f)\) of \(f\) in \(M \times N\), the hypotheses of Theorem 1.4 of [10], namely \(H_1((1_M \times f) \circ \Delta _M;\mathbb {Q})\) is an isomorphism and \(H_2((1_M \times f) \circ \Delta _M;\mathbb {Q})\) is an epimorphism are satisfied. Therefore by applying Theorem 1.4 of [10] one can determine explicitly a model of \(M \times N \setminus \Gamma (f)\) from the above model of \(f\), and our result follows. \(\square \)

Proof of Corollary. 1.5

By hypothesis the dimension of \(N\) is at least three, so \(\text{ dim } \;M = \text{ dim }\; N \ge 3\). Therefore the hypothesis of Corollary 1.5 of [10], namely the codimension of the embedding of \(\Gamma (f)\) in \(M \times N\) is \(\ge 3\), and \(H_*((1_M \times f) \circ \Delta _M;\mathbb {Q})\) is 2-connected are satisfied. Hence, the rational homotopy type of \(M \times N \backslash \Gamma (f)\) is determined by the rational homotopy class of \(1_M \times f\). But the homotopy class of \(1_M \times f\) is determined by the homotopy class of \(f\); hence the corollary holds. \(\square \)

4 The CDGA model of \(M \times N \backslash \Gamma (f)\) based on differential-Poincaré-duality-algebra models of \(M\) and \(N\)

In this section we consider a continuous map \(f:M \rightarrow N\) of closed connected oriented manifolds of dimension \(n\) with \(H^1(M; {\mathbb {Q}}) = 0 = H^1(N; {\mathbb {Q}})\) having a CDGA model \((A_N, d_{A_N}) \buildrel {\psi _f}\over \rightarrow (A_M, d_{A_M})\) in which \((A_M, d_{A_M})\) and \((A_N, d_{A_N})\) are differential connected Poincaré duality algebras of formal dimension \(n\) (refer to the para preceding Theorem 1.6 of the introduction).

We note that all the algebras and relevant results of [9], proved by them in the absolute case, can analogously be developed for any given CDGA morphism of Poincaré duality algebras.

Definition 4.1

Let \((A_2, d_{A_2}) \buildrel {\psi }\over \rightarrow (A_1, d_{A_1})\) be a morphism of Poincaré duality algebras of formal dimension \(n\). Define \(\Delta _{\psi } := (\psi \otimes 1)(\Delta ),\) where \(\Delta \) is the diagonal class defined in the statement of Theorem 1.6 and the element \(\Delta _{\psi }\), which belongs to \( (A_1 \otimes A_2)^n\), will be called the class of \(\psi \).

As in Proposition 4.3 of [9] the element \(\Delta _{\psi } \) does not depend on the choice of the basis \(\{a_i\}_{1\le i \le l}\) of \(A_2\).

Remark 4.2

If \(s^{-n}A_1\) is the suspension of \(A_1\) as defined in §2 of [9] then there is an \((A_1 \otimes A_2)\)-module structure on \(s^{-n}A_1\) given by

$$\begin{aligned} (x \otimes y).(s^{-n}a)=(-1)^{n\deg \,(x)+n\deg \,(y)+\deg \,(a)\deg \,(y)}s^{-n}(x.a.\psi (y)) \end{aligned}$$

for homogeneous elements \(a, x \in A_1\), and \(y \in A_2\) and an obvious \((A_1 \otimes A_2)\)-module structure on \((A_1 \otimes A_2)\). In other words the \((A_1 \otimes A_2)\)-module structure on \(s^{-n}A_1\) is obtained from the obvious structure of \(A_1 \otimes A_1\)-module on \(s^{-n} A_1\) by transporting it along the CDGA map \(1\otimes \psi : A_1 \otimes A_2 \rightarrow A_1 \otimes A_1.\)

In view of the above, statements below follow from the corresponding statements of [9], proved there in the absolute case.

  1. 1.

    Let \(\psi : (A_2, \omega _{A_2}) \rightarrow (A_1, \omega _{A_1})\) be a CDGA morphism of oriented Poincaré duality algebras of finite dimension and of formal dimension \(n\). Then the map \(\widehat{\Delta _{\psi }}:s^{-n} A_1 \rightarrow A_1 \otimes A_2,\;\text{ defined } \text{ by }\; s^{-n}a \mapsto \Delta _{\psi }.(a \otimes 1)\) is a morphism of (\(A_1 \otimes A_2\))-modules.

  2. 2.

    Let \(\psi : (A_2, d_{A_2}, \omega _{A_2}) \rightarrow (A_1, d_{A_1}, \omega _{A_1})\) be a morphism of oriented differential Poincaré algebras of formal dimension \(n\). Then \(\Delta _{\psi }\) is a cocycle.

  3. 3.

    If \(\psi : (A_2, d_{A_2}, \omega _{A_2}) \rightarrow (A_1, d_{A_1}, \omega _{A_1})\) is a morphism of connected differential oriented Poincaré duality algebras of formal dimension \(n\) and \(\widehat{\Delta _{\psi }}:s^{-n} A_1 \rightarrow A_1 \otimes A_2\;\text{ is } \text{ defined } \text{ by }\; s^{-n}a \mapsto \Delta _{\psi }.(a\otimes 1),\) then

    1. (a)

      \(\widehat{\Delta _{\psi }}\) is a morphism of (\(A_1 \otimes A_2\))-dgmodules and hence its mapping cone

      $$\begin{aligned} C(\widehat{\Delta _{\psi }}):= A_1 \otimes A_2\oplus _{\widehat{\Delta _{\psi }}}ss^{-n}A_1 \end{aligned}$$

      is an (\(A_1 \otimes A_2\))-dgmodule.

    2. (b)

      \(C(\widehat{\Delta _{\psi }})\) is also a \(CDGA\) under the following multiplication rules: If \(a, a', x, y \in A_1\), and \(b, b' \in A_2\),

      $$\begin{aligned} (a \otimes b).(a'\otimes b')&= (-1)^{\deg \,(b)\deg \,(a')}(a.a'\otimes b.b'),\\ (a \otimes b).(ss^{-n} x)&= (-1)^{(n-1)(\deg \,(a)+\deg \,(b))}ss^{-n}(a.\psi (b).x), \\ (ss^{-n} x).(a \otimes b)&= ss^{-n}(x.a.\psi (b)),\\ (ss^{-n} x).(ss^{-n} y)&= 0. \end{aligned}$$

Definition 4.3

Let \((A,\omega )\) be a connected oriented Poincaré duality algebra of formal dimension \(n\) with orientation class \(\omega \) (refer to Definition 4.1 of [9]). Since \(A^n \cong \mathbb {Q}\), there exists a unique element \(\mu \in A^n\) such that \(\omega (\mu )=1\) which called the fundamental class of \(A\).

Our goal in this section is to prove Theorem 1.6. Throughout we consider morphisms \(\psi : (A_2, d_{A_2}, \omega _{A_2}) \rightarrow (A_1, d_{A_1}, \omega _{A_1})\) of connected differential oriented Poincaré duality algebras of formal dimension \(n\).

Lemma 4.4

The ideal \((\Delta _{\psi }):=\Delta _{\psi }.(A_1 \otimes A_2)\) generated by \(\Delta _{\psi }\) in \(A_1 \otimes A_2\) is a differential ideal and the quotient \((A_1 \otimes A_2)/(\Delta _{\psi })\) is a CDGA.

Proof

By Remark 4.2 (2), \(d_{A_1 \otimes A_2}(\Delta _{\psi })=0\); hence the ideal \((\Delta _{\psi })\) is a differential ideal. This implies immediately that the quotient \((A_1 \otimes A_2)/(\Delta _{\psi })\) inherits a \(CDGA\) structure.

\(\square \)

Lemma 4.5

The map \(\widehat{\Delta _{\psi }}\) induces an isomorphism

$$\begin{aligned} \widehat{\Delta _{\psi }}:s^{-n}A_1 \rightarrow (\Delta _{\psi }). \end{aligned}$$

Proof

We first show that im \((\widehat{\Delta _{\psi }})=(\Delta _{\psi })\). Clearly \(\mathrm{im}(\widehat{\Delta _{\psi }})\subseteq (\Delta _{\psi })\). For the reverse inclusion we have

$$\begin{aligned} (\Delta _{\psi })&= \{\Delta _{\psi }.(x\otimes y)\;|x\in A_1, y\in A_2\}\\&= \{\Delta _{\psi }.(-1)^{\deg \,(x) \deg \,(y)}(1 \otimes y).(x\otimes 1)\;|x\in A_1, y\in A_2\}\\&= \{\widehat{\Delta _{\psi }}((-1)^{\deg \,(x) \deg \,(y)} s^{-n}(\psi (y).x))\;|x\in A_1, y\in A_2\}\\&\subseteq \mathrm{im}(\widehat{\Delta _{\psi }}). \end{aligned}$$

Next we show that \(\widehat{\Delta _{\psi }}\) is injective. Let \(\omega _{A_1}\) and \(\omega _{A_2}\) be the orientation classes of the Poincaré duality algebras \(A_1\) and \(A_2\) respectively. So there exist unique elements \(\mu _{A_1} \in A_1^n\) and \(\mu _{A_2} \in A_2^n\), fundamental classes of \(A_1\) and \(A_2\), such that \(\omega _{A_1}(\mu _{A_1}) = 1\) and \(\omega _{A_2}(\mu _{A_2}) = 1\). Fix a basis \(\{a_i\}_{i=1}^l\) of \(A_2\) and its Poincaré dual basis \(\{a_i^*\}_{i=1}^l\). Since \( A_2^0=\mathbb {Q}\) we can assume that \(1 = a_1\in A_2^0 \) and remaining \(a_i,\)’s are of degree \(>0\) so that \(a_i \mu _{A_2} \in A_2^{> n} = 0\) for \(i > 1\). Also \(\omega _{A_2}(a_1.a_1^*) = 1\) implies that \(1.a_i^* = a_1.a_1^* = \mu _{A_2}\) so that \(a_1^*=a_1^{-1}.\mu _{A_2} = \mu _{A_2}\). This yields from the definition of \(\Delta _{\psi }\) that

$$\begin{aligned}&\Delta _{\psi }.(\mu _{A_1} \otimes 1)=\sum ^l_{i=1}(-1)^{\deg \,(a_i)}(\psi (a_i) \otimes a_i^*).(\mu _{A_1} \otimes 1)\\&\quad = \sum ^l_{i=1}(-1)^{\deg \,(a_i^*)}(-1)^{n\deg \,(a^*_i)}(\psi (a_i). \mu _{A_1} \otimes a_i^*) = \psi (a_1).\mu _{A_1} \otimes a_1^* \\&\quad = \psi (a_1).\mu _{A_1} \otimes a_1^{-1}.\mu _{A_2} = \mu _{A_1} \otimes \mu _{A_2} \ne 0. \end{aligned}$$

Let \(a\) be a non-zero element of \(A_1^i\). By Poincaré duality there exists an element \(b \in A_1^{n-i}\) such that \(a.b=\mu _{A_1}.\) We have

$$\begin{aligned} \widehat{\Delta _{\psi }}(s^{-n} a).(b \otimes 1)&= \Delta _{\psi }.(a\otimes 1).(b \otimes 1) = \Delta _{\psi }.(a.b \otimes 1)= \Delta _{\psi }.(\mu _{A_1} \otimes 1)\\&= \mu _{A_1} \otimes \mu _{A_2} \ne 0. \end{aligned}$$

\(\square \)

Consider the projection \(A_1 \otimes A_2 \xrightarrow {\pi }(A_1 \otimes A_2)/(\Delta _{\psi }).\) We extend \(\pi \) to a map \(\hat{\pi }: A_1 \otimes A_2 \oplus _{\widehat{\Delta _{\psi }}}ss^{-n} A_1 \rightarrow (A_1 \otimes A_2)/(\Delta _{\psi })\) by setting \(\hat{\pi }(ss^{-n} A_1) = 0\).

Lemma 4.6

The map \(\hat{\pi }: A_1 \otimes A_2 \oplus _{{\widehat{\Delta _{\psi }}}}ss^{-n} A_1 \rightarrow (A_1 \otimes A_2)/(\Delta _{\psi })\) defined above is a CDGA quasi-isomorphism.

Proof

The map \(\hat{\pi }\) can be seen to be a \(CDGA\) morphism by a straightforward computation. Since \(\widehat{\Delta _{\psi }}\) is injective , we have a short exact sequence

$$\begin{aligned} 0\rightarrow s^{-n}A_1\xrightarrow {{\widehat{\Delta _{\psi }}}} A_1\otimes A_2\xrightarrow {{\pi }}(A_1\otimes A_2)/\mathrm{im}(\widehat{\Delta _{\psi }})\rightarrow 0. \end{aligned}$$

Comparing the long cohomology exact sequence corresponding to this exact sequence and that of the mapping cone \(s^{-n} A_1 \xrightarrow {\widehat{\Delta _{\psi }}} A_1\otimes A_2 \hookrightarrow C(\widehat{\Delta _{\psi }}):= A_1\otimes A_2 \oplus _{\widehat{\Delta _{\psi }}}ss^{-n} A_1\) we get the following commutative diagram:

Applying five lemma and using Lemma 4.5 we get that the map

$$\begin{aligned} A_1\otimes A_2 \oplus _{\widehat{\Delta _{\psi }}}ss^{-n}A_1 \xrightarrow {\hat{\pi }=\pi \oplus 0} (A_1\otimes A_2)/\mathrm{im}(\widehat{\Delta _{\psi }})=(A_1 \otimes A_2)/(\Delta _{\psi }) \end{aligned}$$

is a quasi isomorphism. \(\square \)

We are now ready to prove Theorem 1.6.

Proof of Theorem. 1.6

We have proved in Lemma 4.4 that \((\Delta _{\psi _f})\) is a differential ideal. Since \((A_M, d_{A_M})\) and \((A_N, d_{A_N})\) are connected differential Poincaré duality algebras of formal dimension \(n\) and since \(H^n(A_M, d_{A_M})=H^n(M;\mathbb {Q})\ne 0\) and \(H^n(A_N, d_{A_N})=H^n(N;\mathbb {Q})\ne 0\), Proposition 4.8 of [9] implies that \((A_M, d_{A_M}, \omega _{A_M})\) and \((A_N, d_{A_N}, \omega _{A_N})\) are oriented differential Poincaré duality algebras in the sense of Definition 4.6 of [9] for orientations \(\omega _{A_M} \in \#A_M^n\) and \(\omega _{A_N}\in \#A_N^n\).

We have proved in §3 that a CDGA model \(\psi _f: (A_N, d_{A_N}) \rightarrow (A_M, d_{A_M})\) of \(f\) gives a CDGA model

$$\begin{aligned} \Phi := \nu _{A_M} \circ (1_{A_M} \otimes \psi _f): (A_M \otimes A_N, d_{A_N} \otimes d_{A_M}) \rightarrow (A_M, d_{A_M}) \end{aligned}$$

of \((1_M \times f) \circ \Delta : M \rightarrow M \times N,\) and is given by \(x \otimes y \mapsto x.\psi _f(y)\). Set \(n=2m\). The morphism \(\Phi \) induces an obvious (\(A_M\otimes A_N\))-dgmodule structure on \(A_M\), hence on \(s^{-n}\#A_M\). By Proposition 4.7 of [9] \(s^{-n}\#A_M = s^{-m}(s^{-m}\#A_M)\) is isomorphic to \(s^{-m}A_M\) as an \(A_M\)-dgmodule, therefore also an isomorphism as an (\(A_M\otimes A_N\))-dgmodule.

By Remark 4.2a the map \(\widehat{\Delta _{\psi }}:s^{-n}A_M \rightarrow A_M\otimes A_N\) is a map of (\(A_M\otimes A_N\))-dgmodules. Moreover, it induces an isomorphism in the cohomology in the top degree, because \(\widehat{\Delta _{\psi }}(s^{-n}\mu )= \mu _{A_M} \otimes \mu _{A_N}\) . Thus \(\widehat{\Delta _{\psi }}\) is a shriek map (or, a top-degree map) in the sense of Definition 5.1 of [10].

Let \(I_0\) be a complement of the cocycles in \((A_M \otimes A_N)^{n-3} \) and set \(I = I_0 \oplus (A_M\otimes A_N)^{>n-3}\). Let \(K_0\) be a complement of the cocycles in \((s^{-n}A_M)^{n-2}\) and set \(K = K_0 \oplus (s^{-n}A_M)^{>n-2}\). Consider the quotient

$$\begin{aligned} \frac{A_M\otimes A_N \oplus _{\widehat{\Delta _{\psi }}} ss^{-n}A_M}{I \oplus sK}. \end{aligned}$$

By the hypothesis of the theorem concerning \(f\), it follows that the embedding \((1_M \times f) \circ \Delta _M: M \hookrightarrow M \times N\), which embeds \(M\) as \(\Gamma (f)\) in \(M \times N\), satisfies the hypothesis of Theorem 1.4 of [10], namely that \(H_1((1_M \times f)\circ \Delta _M; {\mathbb {Q}})\) is an isomorphism and \(H_2((1_M \times f) \circ \Delta _M; {\mathbb {Q}})\) is an epimorphism, hence by Theorem 1.4 of [10] we get that

$$\begin{aligned} \left( \frac{A_M\otimes A_N \oplus _{\widehat{\Delta _{\psi }}} ss^{-n}A_M}{I \oplus sK}, \bar{d}\right) \end{aligned}$$

is a CDGA model of \(M \times N\setminus \Gamma (f)\).

Comparing the semi-trivial CDGA structure as defined in Definition 4.1 of [10] on \((A_M\otimes A_N \oplus _{\widehat{\Delta _{\psi }}}ss^{-n}A_M)/(I \oplus sK)\) with the multiplication in the mapping cone given in this section, we conclude that the projection

$$\begin{aligned} (A_M\otimes A_N \oplus _{\widehat{\Delta _{\psi }}}ss^{-n}A_M, d) \rightarrow \left( \frac{A_M\otimes A_N \oplus _{\widehat{\Delta _{\psi }}}ss^{-n}A_M}{I\oplus sK}, \bar{d}\right) \end{aligned}$$

is a CDGA map. Moreover it is a quasi-isomorphism by Lemma 8.6 of [10].

Therefore \(( A_M\otimes A_N \oplus _{\widehat{\Delta _{\psi }}}ss^{-n}A_M, d)\) is a CDGA model of \( M\times N \setminus \Gamma (f)\). Hence by Lemma 4.6 \(\left( (A_M\otimes A_N)/(\Delta _{\psi }), \overline{d_{A_M} \otimes d_{A_N}}\right) \) is a CDGA model of \( M\times N \setminus \Gamma (f)\). \(\square \)

We finally prove Corollary 1.7 on formality of the complement of the graph of a map.

Proof of corollary. 1.7

Since \(f:M \rightarrow N\) is formal, \((H^*(N; {\mathbb {Q}}), 0)\buildrel {f^*}\over \rightarrow . (H^*(M; {\mathbb {Q}}), 0)\) is a model of \(f\).

Moreover, since \((H^*(N; {\mathbb {Q}}), 0)\) and \((H^*(M; {\mathbb {Q}}), 0)\) are oriented differential Poincaré duality algebras, by Theorems 1.6 and 1.8,

$$\begin{aligned} \left( \frac{H^*(M;\mathbb {Q})\otimes H^*(N;\mathbb {Q})}{(\Delta _f)}, 0\right) \cong (H^*(M\times N\setminus \Gamma (f);\mathbb {Q}), 0) \;\; \text{(as } \text{ CDGAs) } \end{aligned}$$

is a CDGA-model of \(M \times N\setminus \Gamma (f)\), and hence it is a formal space. \(\square \)

5 Examples and applications

In the last section we saw that if \(f:M \rightarrow N\) is a formal map of closed connected oriented formal manifolds of dimension \(n\) such that \(H^1(M;\mathbb {Q}) = 0 = H^1(N;\mathbb {Q}) = H^2(N;\mathbb {Q}),\) then \(M \times N\setminus \Gamma (f)\) is formal (Corollary 1.7). In Proposition 6.6 of [9] Lambrechts and Stanley proved that if \(M\) is a closed connected orientable manifold of dimension \(n\) such that \(M\times M\setminus \Gamma (1_M) = F(M,2)\) is formal then \(M\) is formal. We ask a similar question (this is the converse of Corollary 1.7):

Question 5.1

Given closed connected oriented formal manifolds \(M\) and \(N\) of dimension \(n\) and a continuous map \(f:M\rightarrow N\) such that \(M \times N\setminus \Gamma (f)\) is formal, is \(f\) necessarily a formal map?

The answer is in the negative. Here we construct such an example.

Example 5.2

Consider \( S^7\xrightarrow {f=(h,c)} S^4\times S^3,\) where \(h:S^7\rightarrow S^4\) is the Hopf map and \(c:S^7\rightarrow S^3\) is a constant map. It follows from Proposition 2.5 of [1] that any formal map \(S^7 \rightarrow S^4\times S^3\) must be nullhomotopic; but \(f\) is not nullhomotopic. Therefore \(f\) is not formal.

Now we show that \(S^7 \times (S^4\times S^3)\setminus \Gamma (f)\) is formal. Since \(S^3\xrightarrow {j} S^7\xrightarrow {h} S^4\) is a Hopf fibration we have the following commutative diagram [see example (2.68) page 82 of [5] or Chapter 15 of [3]

Here deg \((x)=4\), deg \((y)=7\), and deg \((z)=3\); \(d'x=0\), \(d'y=x^2\), \(d x=0\), \(d y=x^2\), \(d z = x;\) and \(d''z=0;\) the morphism \(\rho :(\Lambda (x,y),d')\rightarrow A_{PL}(S^4)\) is the minimal model of \(S^4\), \(\sigma :(\Lambda (x,y)\otimes \Lambda (z),d)\rightarrow A_{PL}(S^7)\) is a quasi-isomorphism, \(\xi :(\Lambda (z),d'')\rightarrow A_{PL}(S^3)\) is the minimal model of \(S^3\) , \(i:(\Lambda (x,y),d')\rightarrow (\Lambda (x,y)\otimes \Lambda (z),d)\) is a relative minimal CDGA and \(\eta : (\Lambda (x,y)\otimes \Lambda (z),d) \rightarrow (\Lambda (z),d'')\) is the projection.

From the above diagram we get the following commutative diagram

where \(i(x)=x,\; i(y)=y\;\) and \(i(z)=0\).

Let us define CDGAs

$$\begin{aligned} (B,0):=(\Lambda (e_4,e_3)/I_B, 0), \end{aligned}$$

where \(\deg \;(e_4)=4;\;\deg (e_3)=3\), \(I_B\) is the ideal generated by \(e_4^2,\) and

$$\begin{aligned} (A,d_A):=(\Lambda (u_4,u_3)/I_A, d_A), \end{aligned}$$

where \(\deg \;(u_4)=4;\;\deg (u_3)=3,\; I_A\) is the ideal generated by \(u_4^2, \; d_A\overline{u_3}=\overline{u_4},\) and \(d_A\overline{u_4}=0\). It is easily checked that \((B,0)\), and \((A,d_A)\) are oriented differential Poincaré duality algebras.

We define morphisms

$$\begin{aligned} \rho _B: (\Lambda (x,y)\otimes \Lambda (z),d'\otimes d'')\rightarrow (B, 0), \end{aligned}$$

where \( \rho _B(x)=\overline{e_4},\;\rho _B(y)=0,\;\rho _B(z)=\overline{e_3}\) and

$$\begin{aligned} \rho _A: (\Lambda (x,y)\otimes \Lambda (z),d)\rightarrow (A, d_A), \end{aligned}$$

where \(\rho _A(x)= \overline{u_4},\;\rho _A(y)=0,\;\rho _A(z)=\overline{u_3}.\)

Clearly \(\rho _B\) and \(\rho _A\) are quasi-isomorphisms.

Define \(\psi _f:(B,0)\rightarrow (A,d_A)\) by \( \psi _f(\overline{e_4})= \overline{u_4},\;\psi _f(\overline{e_3})= 0\) so that the following diagram is commutative.

Using the fact that if \(X\xrightarrow {(k,l)}Y\times Z\) is a continuous map then \(A_{PL}(k)\cdot A_{PL}(l): A_{PL}(Y)\otimes A_{PL}(Z) \rightarrow A_{PL}(X)\) is its model, we get the following commutative diagram

Therefore \(\psi _f\) is a model of \(f=(h,c)\). So, by Theorem 1.6

$$\begin{aligned} \left( \frac{A\otimes B}{(\bigtriangleup _{\psi _f})}, \overline{d_A \otimes 0}\right) \end{aligned}$$

is a CGDA model of \(S^7 \times (S^4\times S^3)\setminus \Gamma (f)\).

Note that \(\{a_1=1,\; a_2=\overline{e_3},\; a_3=\overline{e_4},\; a_4=\overline{e_4}\cdot \overline{e_3}\}\) is a homogeneous basis of \((B,0)\) with Poincaré dual basis \(\{a_1^*=\overline{e_4}\cdot \overline{e_3},\; a_2^*=\overline{e_4},\; a_3^*=\overline{e_3},\; a_4^*=1\}\), therefore by definition

$$\begin{aligned} \bigtriangleup _{\psi _f}&= \psi _f(1)\otimes \overline{e_4}\cdot \overline{e_3}-\psi _f(\overline{e_3})\otimes \overline{e_4}+\psi _f(\overline{e_4})\otimes \overline{e_3} - \psi _f(\overline{e_4}\cdot \overline{e_3})\otimes 1\\&= 1\otimes \overline{e_4}\cdot \overline{e_3}-\overline{u_4}\otimes \overline{e_3}. \end{aligned}$$

Therefore as a CDGA

$$\begin{aligned} \left( \frac{A\otimes B}{(\bigtriangleup _{\psi _f})}, \overline{d_A \otimes 0} = \overline{d}\right) =\;\;\; \mathbb {Q}(\overline{1\otimes 1})\xrightarrow {0} 0 \xrightarrow {0} 0 \xrightarrow {0} \mathbb {Q}(\overline{1\otimes \overline{e_3}}, \;\overline{\overline{u_3} \otimes 1}) \end{aligned}$$
$$\begin{aligned} \xrightarrow {\overline{d}} \mathbb {Q}(\overline{1\otimes \overline{e_4}}, \; \overline{\overline{u_4} \otimes 1}) \xrightarrow {0} 0 \xrightarrow {0}\mathbb {Q}(\overline{\overline{u_3} \otimes \overline{e_3}})\xrightarrow {\overline{d}}\mathbb {Q}(\overline{\overline{u_3}\otimes \overline{e_4}},\;\overline{\overline{u_4}\otimes \overline{e_3}},\; \overline{\overline{u_4}. \overline{u_3} \otimes 1}) \end{aligned}$$
$$\begin{aligned} \xrightarrow {\overline{d}} \mathbb {Q}(\overline{\overline{u_4} \otimes \overline{e_4}}) \xrightarrow {0} 0 \xrightarrow {0} \mathbb {Q}(\overline{\overline{u_4}. \overline{u_3}\otimes \overline{e_3}}) \xrightarrow {0} \mathbb {Q}(\overline{\overline{u_4}. \overline{u_3} \otimes \overline{e_4}}) \xrightarrow {0} 0\cdots \end{aligned}$$

Let \(c_3\in H^3(S^3;\mathbb {Q})\), \(c_4\in H^4(S^4;\mathbb {Q})\) and \(c_7\in H^7(S^7;\mathbb {Q})\) be bases of the respective vector spaces. Then \( \{b_1=1 \otimes 1,\; b_2=1 \otimes c_3,\; b_3=c_4 \otimes 1,\;b_4 = c_4 \otimes c_3\}\) form a homogeneous basis of \(H^*(S^4\times S^3; \mathbb {Q}) = H^*(S^4; \mathbb {Q}) \otimes H^*(S^3; \mathbb {Q})\) with Poincaré dual basis \(\{b_1^* = c_4 \otimes c_3,\; b_2^*=c_4 \otimes 1,\; b_3^*=1 \otimes c_3,\;b_4^*=1 \otimes 1\}.\) We know by Theorem 1.8 that

$$\begin{aligned} H^*(S^7 \times (S^4\times S^3)\!\setminus \! \Gamma (f);\mathbb {Q})=\frac{H^*(S^7;\mathbb {Q})\otimes H^*(S^4\times S^3;\mathbb {Q})}{( \bigtriangleup _f)}, \end{aligned}$$

where

$$\begin{aligned} \bigtriangleup _f&= f^*(1 \otimes 1)\otimes c_4\otimes c_3 - f^*(1 \otimes c_3)\otimes c_4 \otimes 1 + f^*(c_4 \otimes 1)\otimes 1 \otimes c_3 \\&- f^*(c_4\otimes c_3)\otimes 1\otimes 1\\&= 1\otimes c_4\otimes c_3. \end{aligned}$$

Therefore as a CDGA

$$\begin{aligned}&(H^*(S^7 \times (S^4\times S^3)\!\setminus \! \Gamma (f);\mathbb {Q}),\; 0) \\&=\;\mathbb {Q}(\overline{1 \otimes 1 \otimes 1})\xrightarrow {0}0\xrightarrow {0}0\xrightarrow {0}\\&\mathbb {Q}(\overline{1 \otimes 1\otimes c_3}) \\&\quad \xrightarrow {0} \mathbb {Q}(\overline{1\otimes c_4 \otimes 1})\xrightarrow {0}0\xrightarrow {0}0\xrightarrow {0}\mathbb {Q}(\overline{c_7\otimes 1 \otimes 1})\xrightarrow {0}0\xrightarrow {0}0 \\&\quad \xrightarrow {0} \mathbb {Q}(\overline{c_7\otimes 1 \otimes c_3})\xrightarrow {0}\mathbb {Q}(\overline{c_7\otimes c_4 \otimes 1})\xrightarrow {0}0\ldots \end{aligned}$$

Now we define a CDGA morphism

$$\begin{aligned} \eta :(H^*(S^7 \times (S^4\times S^3)\!\setminus \! \Gamma (f);\mathbb {Q}),\; 0) \rightarrow \left( \frac{A\otimes B}{(\bigtriangleup \psi _f)}, d_A \otimes 0\right) , \end{aligned}$$

by defining it on generators as follows:

$$\begin{aligned} \eta (\overline{1\otimes 1\otimes 1})&= \overline{1\otimes 1}, \eta (\overline{1 \otimes 1 \otimes c_3}) = \overline{1 \otimes \overline{e_3}}, \eta (\overline{1 \otimes c_4 \otimes 1}) = \overline{1\otimes \overline{e_4}} + \overline{\overline{u_4} \otimes 1}, \\ \eta (\overline{c_7 \otimes 1 \otimes 1})&= \overline{\overline{u_4}. \overline{u_3} \otimes 1}, \; \eta (\overline{c_7 \otimes 1 \otimes c_3}) = \overline{\overline{u_4}. \overline{u_3} \otimes \overline{e_3}}, \; \eta (\overline{c_7 \otimes c_4 \otimes 1}) \\&= \overline{\overline{u_4}. \overline{u_3} \otimes \overline{e_4}}. \end{aligned}$$

It is checked easily that \(\eta \) is a quasi-isomorphism. Therefore \((H^*(S^7 \times (S^4\times S^3)\!\setminus \! \Gamma (f);\mathbb {Q}),\; 0)\) is a CDGA model of \(S^7 \times (S^4\times S^3)\setminus \Gamma (f)\). Hence \(S^7 \times (S^4\times S^3)\setminus \Gamma (f)\) is a formal space.

Remark 5.3

It is known (see e.g. Lemma 6.3 of [9]) that if \(M\) is a simply connected closed manifold, if \(x\in M\), and if \(M\!\setminus \! \{x\}\) is formal, then \(M\) is formal.

Against this background we may ask the following question.

Question 5.4

Let \(f:M\rightarrow N\) be a continuous map between simply connected formal manifolds and let \(x\in M\). Suppose that \(f|_{M\setminus \{x\}} :M\!\!\setminus \!\!\{x\} \longrightarrow N\!\!\setminus \!\! \{f(x)\}\) is formal. Is it necessarily true that \(f\) is formal ?

The following example shows that the answer to this question is in the negative.

Example 5.5

Take \(M=S^3\), \(N=S^2\) and \(f\) the Hopf map \(h:S^3\rightarrow S^2\). Now \(S^3\!\setminus \!\!\{(0,0,0,1)\} \cong \mathbb {R}^3\) and \(S^2\!\setminus \!\! \{(0,0,1)\} \cong \mathbb {R}^2\). Since any continuous map form \(\mathbb {R}^3 \rightarrow \mathbb {R}^2\) is formal, \(h|_{S^3\setminus \{(0,0,0,1)\}}: S^3\!\setminus \!\! \{(0,0,0,1)\} \longrightarrow S^2\!\setminus \! \{(0,0,1)\}\) is formal. But \(h\) is not formal.

We end the paper by recording a few simple applications of Theorem 1.6 and Corollary 1.7:

Application 5.6

  1. 1.

    Let \(M\) and \(N\) be two simply connected closed formal manifolds of dimension \(n\). For any element \(y\in N\), since the constant map \(f_y: M\rightarrow N\) defined by \( f_y(x) = y\) for all \(x\in X\) is formal, by Corollary 1.7, \(M \times N\setminus \Gamma (f)\) is formal. But \(M \times N\setminus \Gamma (f)= M\times (N\!\setminus \!\{y\})\). Therefore \(M\times (N\!\setminus \!\{y\})\) is formal.

  2. 2.

    Let \(M\) be a simply connected closed formal manifold of dimension \(n\). Since the identity map \(f=1_M: M\rightarrow M\) is formal, by Corollary 1.7, \(M \times M\!\setminus \! \Gamma (f)\) is formal. But \(M \times M\setminus \Gamma (f)=F(M,2)\). Therefore \(F(M,2)\) is formal, which is one of the main results (Corollary 6.1) in [9].

  3. 3.

    Let \(M\) and \(N\) be two simply connected Lie groups of dimension \(n\). It is known that the minimal model of a Lie group is of the form \((\Lambda V ,0)\). Hence Lie groups are formal and continuous maps between Lie groups are also formal. Therefore, by Corollary 1.7, for a continuous map \(f:M\rightarrow N\), \(M \times N\!\setminus \! \Gamma (f)\) is formal.