1 Introduction

Since Mumford’s famous 1969 paper [12], it is well-known that the Chow group of 0-cycles \(A^nX\) on a complex variety X influences the cohomology group \(H^n(X,\mathbb {Q})\):

Theorem

(Mumford [12]) Let X be a smooth complete variety of dimension n defined over \(\mathbb {C}\). Suppose that \(A^nX_{\mathbb {Q}}\) is supported on a divisor. Then \(H^n(X,\mathbb {Q})\) is supported on a divisor, in particular \(H^n(X,\mathcal O_X)=0\).

In the 1992 paper [5], Esnault, Srinivas and Viehweg study the multiplicative behaviour of the Chow ring \(A^*X\) versus the multiplicative behaviour of various cohomology rings associated to X. We now state the part of their result that is relevant to us. For a given partition \(n=n_1+\cdots +n_r\) (with \(n_i\in \mathbb {N}_{>0}\)), let us consider the following properties:

  • (P1) There exists a Zariski open \(V\subset X\), such that intersection product induces a surjection

    $$\begin{aligned} A^{n_1}V_{\mathbb {Q}}\otimes A^{n_2}V_{\mathbb {Q}}\otimes \cdots \otimes A^{n_r}V_{\mathbb {Q}}\ \rightarrow \ A^{n}V_{\mathbb {Q}}\ ; \end{aligned}$$
  • (P2) There exists a Zariski open \(V\subset X\), such that cup product induces a surjection

    $$\begin{aligned} H^{n_1}(V,\mathbb {Q})\otimes H^{n_2}(V,\mathbb {Q})\otimes \cdots \otimes H^{n_r}(V,\mathbb {Q})\ \rightarrow \ H^n(V,\mathbb {Q})/N^1\ \end{aligned}$$

    (here \(N^*\) denotes the coniveau filtration);

  • (P3) Cup product induces a surjection

    $$\begin{aligned} H^{n_1}(X,\mathcal O_X)\otimes H^{n_2}(X,\mathcal O_X)\otimes \cdots \otimes H^{n_r}(X,\mathcal O_X)\ \rightarrow \ H^n(X,\mathcal O_X). \end{aligned}$$

    In these terms, what Esnault, Srinivas and Viehweg prove is the following:

Theorem

(Esnault et al. [5]) Let X be a smooth complete variety of dimension n over \(\mathbb {C}\). Then (P1) implies (P3), and (P2) implies (P3).

The implication from (P1) to (P3) is a kind of multiplicative variant of Mumford’s theorem, and the proof in [5] is motivated by Bloch’s proof of Mumford’s theorem using a “decomposition of the diagonal” argument ([2, 3], cf. also [4]). As noted in [5, remark 2], the generalized Hodge conjecture would imply that (P2) and (P3) are equivalent.Footnote 1

In this note, we show that the Esnault–Srinivas–Viehweg theorem can be extended from 0-cycles to arbitrary Chow groups. This is possible provided the variety X satisfies the Lefschetz standard conjecture B(X) (this is analogous to [10], where I extended Mumford’s theorem from 0-cycles to arbitrary Chow groups, provided B(X) holds):

Theorem

(\(=\)Theorem 1) Let X be a smooth projective variety of dimension n over \(\mathbb {C}\) that satisfies B(X). Suppose there exists a Zariski open \(V\subset X\), and \(j=j_1+\cdots + j_r\) with \(j_i\in \mathbb {N}_{>0}\) such that intersection product induces a surjection

$$\begin{aligned} A^{j_1}V_{\mathbb {Q}}\otimes A^{j_2}V_{\mathbb {Q}}\otimes \cdots \otimes A^{j_r}V_{\mathbb {Q}}\ \rightarrow \ A^jV_{\mathbb {Q}}. \end{aligned}$$

Then cup product induces a surjection

$$\begin{aligned} H^{j_1}(X,\mathcal O_X)\otimes H^{j_2}(X,\mathcal O_X)\otimes \cdots \otimes H^{j_r}(X,\mathcal O_X)\ \rightarrow \ H^j(X,\mathcal O_X). \end{aligned}$$

The proof of this theorem, which is very similar to the proof given by Esnault–Srinivas–Viehweg in [5], is an exercise in using the meccano of correspondences and the Bloch–Srinivas formalism.

It seems natural to wonder whether the converse to Theorem 1 might perhaps be true (this would be a multiplicative variant of Bloch’s conjecture). In [11], I prove this converse implication in some special cases for 0-cycles (i.e. \(j=n\)); the converse implication for \(j\not =n\) appears to be more difficult.

Convention In this note, the word variety will refer to a quasi-projective irreducible algebraic variety over \(\mathbb {C}\), endowed with the Zariski topology. A subvariety is a (possibly reducible) reduced subscheme which is equidimensional. The Chow group of j-dimensional algebraic cycles on X with \(\mathbb {Q}\)-coefficients modulo rational equivalence is denoted \(A_jX\); for X smooth of dimension n the notations \(A_jX\) and \(A^{n-j}X\) will be used interchangeably. Caveat: note that what we denote \(A^jX\) is elsewhere often denoted \(CH^j(X)_{\mathbb {Q}}\). In an effort to lighten notation, we will often write \(H^jX\) or \(H_jX\) to indicate singular cohomology \(H^j(X,\mathbb {Q})\) resp. Borel–Moore homology \(H_j(X,\mathbb {Q})\).

For basics concerning algebraic cycles and their functorial behaviour, the curious reader is invited to consult [6]. For the formalism of correspondences, cf. [14, 15].

2 Preliminary

Let X be a smooth projective variety of dimension n, and \(h\in H^2(X,\mathbb {Q})\) the class of an ample line bundle. The hard Lefschetz theorem asserts that the map

$$\begin{aligned} L^{n-i}:H^i(X,\mathbb {Q})\rightarrow H^{2n-i}(X,\mathbb {Q}) \end{aligned}$$

obtained by cupping with \(h^{n-i}\) is an isomorphism, for any \(i< n\). One of the standard conjectures asserts that the inverse isomorphism is algebraic.

Definition 1

(Lefschetz standard conjecture) Given a variety X, we say that B(X) holds if for all ample h, and all \(i<n\) the isomorphism

$$\begin{aligned} (L^{n-i})^{-1}:H^{2n-i}(X,\mathbb {Q})\mathop {\rightarrow }\limits ^{\cong } H^i(X,\mathbb {Q}) \end{aligned}$$

is induced by a correspondence.

Remark 1

It is known that B(X) holds for the following varieties: curves, surfaces, abelian varieties [8, 9], threefolds not of general type [16], varieties motivated by a surface in the sense of Arapura [1] (this includes the Hilbert schemes of 0-dimensional subschemes of surfaces [1, Corollary 7.5]), n-dimensional varieties X which have \(A_i(X)_{}\) supported on a subvariety of dimension \(i+2\) for all \(i\le {n-3\over 2}\) [17, Theorem 7.1], n-dimensional varieties X which have \(H_i(X)=N^{\llcorner {i\over 2}\lrcorner }H_i(X)\) for all \(i>n\) [18, Theorem 4.2], products and hyperplane sections of any of these [8, 9].

It is known that B(X) implies that the Künneth components

$$\begin{aligned} \pi _j\in H^{2n-j}(X)\otimes H^j(X)\subset H^{2n}(X\times X) \end{aligned}$$

of the diagonal \(\Delta \subset X\times X\) are algebraic [8, 9]. Moreover, these Künneth components satisfy the following property:

Lemma 1

Let X be a smooth projective variety satisfying B(X), and let \(h\in H^2(X)\) be the class of an ample line bundle. For any \(j\le n\), there exists a cycle \(P_j\in A^j(X\times X)\) such that

$$\begin{aligned} \pi _j=(\tau \times \hbox {id})_*(\tau \times \hbox {id})^*(P_j)\ \ \in H^{2n}(X\times X) , \end{aligned}$$

where \(\tau :Y_j\rightarrow X\) denotes the inclusion of a dimension j complete intersection of class \([Y_j]=h^{n-j}\).

Proof

As mentioned above, B(X) ensures that \(\pi _j\) is algebraic [8, 9]. Consider now the isomorphism

$$\begin{aligned} L^{n-j}\times \hbox {id}:\ \ H^jX\otimes H^jX\ \xrightarrow {\cong }\ H^{2n-j}X\otimes H^jX\ \end{aligned}$$

(here we tacitly identify both sides with their images in \(H^*(X\times X)\)).

Since we have B(X), there exists a correspondence, say \(Q\in A^{j}(X\times X)\), such that

$$\begin{aligned} (L^{n-j}\times \hbox {id}) (Q\times \hbox {id})_*=\hbox {id}:\ H^{2n-j}X\otimes H^jX\ \rightarrow \ H^{2n-j}X\otimes H^jX. \end{aligned}$$

Since \(\pi _j\) is algebraic,

$$\begin{aligned} P_j:= (Q\times \hbox {id})_*(\pi _j)\in A^j(X\times X) \end{aligned}$$

is still algebraic, and has the requested property. \(\square \)

Remark 2

Lemma 1 implies in particular that for a variety satisfying B(X), the Künneth component \(\pi _j\) is represented by an algebraic cycle contained in \(Y_j\times X\), for a dimension j complete intersection \(Y_j\). This was also proven in [7] (and independently in [10, proof of Theorem 3.1], as I wasn’t aware of the Kahn–Murre–Pedrini reference at the time).

3 Main

We now prove the main theorem of this note:

Theorem 1

Let X be a smooth projective variety of dimension n over \(\mathbb {C}\) that satisfies B(X). Suppose there exists a Zariski open \(V\subset X\), and \(j=j_1+\cdots + j_r\) with \(j_i\in \mathbb {N}_{>0}\) such that intersection product induces a surjection

$$\begin{aligned} A^{j_1}V_{}\otimes A^{j_2}V_{}\otimes \cdots \otimes A^{j_r}V_{}\ \rightarrow \ A^jV_{}. \end{aligned}$$

Then cup product induces a surjection

$$\begin{aligned} H^{j_1}(X,\mathcal O_X)\otimes H^{j_2}(X,\mathcal O_X)\otimes \cdots \otimes H^{j_r}(X,\mathcal O_X)\ \rightarrow \ H^j(X,\mathcal O_X). \end{aligned}$$

Proof

Since B(X) holds, it follows from Lemma 1 that the Künneth component \(\pi _j\) can be written

$$\begin{aligned} \pi _j=(\tau \times \hbox {id})_*(\tau \times \hbox {id})^*(P_j)\ \ \in H^{2n}(X\times X)\ \end{aligned}$$

for some \(P_j\in A^j(X\times X)\), where \(\tau :Y_j\rightarrow X\) is the inclusion of a dimension j complete intersection.

Applying the Bloch–Srinivas argument, in the form of Proposition 1 below, to the cycle \(P_j\in A^j(X\times X)\), we find a decomposition

$$\begin{aligned} P_j=C_1\cdot \ldots \cdot C_r+\Gamma _1+\Gamma _2\ \ \in A^j(X\times X), \end{aligned}$$

where \(\Gamma _1, \Gamma _2\) are supported on \(D\times X\), resp. on \(X\times D\), for some divisor \(D\subset X\). This induces a decomposition of the Künneth component

$$\begin{aligned} \pi _j&=(\tau \times \hbox {id})_*(\tau \times \hbox {id})^*(C_1\cdot \ldots \cdot C_r)+ (\tau \times \hbox {id})_*(\tau \times \hbox {id})^*(\Gamma _1+\Gamma _2)\\&=(\tau \times \hbox {id})_*(\tau \times \hbox {id})^*(C_1\cdot \ldots \cdot C_r)+ \Gamma _1^\prime +\Gamma _2^\prime \ \ \in H^{2n}(X\times X), \end{aligned}$$

where \(\Gamma _2^\prime \) is still supported on \(X\times D\), and \(\Gamma _1^\prime \) is supported on \(Z\times X\), for some \(Z\subset X\) of dimension \(j-1\) (indeed, the general complete intersection \(Y_j\) will be in general position with respect to D; we then define Z to be \(D\cap Y_j\)).

Now we consider the action of \(\pi _j\) on \(H^j(X,\mathcal O_X)\). Since \(H^j(X,\mathcal O_X)=\hbox {Gr}^0_F H^j(X,\mathbb {C})\) (where F is the Hodge filtration), \(\pi _j\) acts as the identity on \(H^j(X,\mathcal O_X)\). On the other hand, it is clear that

$$\begin{aligned} (\Gamma _1^\prime )_*H^j(X,\mathcal O_X)=0 \end{aligned}$$

(by Lemma 3, the action of \(\Gamma _1^\prime \) factors over \(\hbox {Gr}^0_F H^j(Z,\mathbb {C})\), which is 0 for dimension reasons), and also that

$$\begin{aligned} (\Gamma _2^\prime )_*H^j(X,\mathcal O_X)=0 \end{aligned}$$

(by Lemma 3, the action of \(\Gamma _2\) factors over \(\hbox {Gr}^{-1}_F H^{j-2}(\widetilde{D},\mathbb {C})=0\), where \(\widetilde{D}\) is a resolution of singularities of D). To finish the argument, it only remains to analyze the action

$$\begin{aligned} \bigl ((\tau \times \hbox {id})_*(\tau \times \hbox {id})^*(C_1\cdot \ldots \cdot C_r)\bigr )_*:\ \ H^j(X,\mathcal O_X)\ \rightarrow \ H^j(X,\mathcal O_X). \end{aligned}$$

Using Lemmas 2 and 3, we find an inclusion

$$\begin{aligned} \bigl ((\tau \times \hbox {id})_*(\tau \times \hbox {id})^*(C_1\cdot \ldots \cdot C_r)\bigr )_*H^j(X,\mathcal O_X) \subset \bigl (C_1\cdot \ldots \cdot C_r\bigr )_*\hbox {Gr}_F^{n-j} H^{2n-j}(X,\mathbb {C}). \end{aligned}$$

Using Lemma 4, we find that

$$\begin{aligned} \bigl (C_1\cdot \ldots \cdot C_r\bigr )_*\hbox {Gr}_F^{n-j} H^{2n-j}(X,\mathbb {C})\!\subset \! \hbox {Im}\bigl (H^{j_1}(X,\mathcal O_X)\!\otimes \cdots \otimes \!H^{j_r}(X,\mathcal O_X) \!\rightarrow \!H^j(X,\mathcal O_X)\!\bigr ) , \end{aligned}$$

and so we are done. \(\square \)

Proposition 1

(Bloch–Srinivas style) Let X be a smooth projective variety of dimension n. Suppose there exists a Zariski open \(V\subset X\), and \(j=j_1+\cdots + j_r\) with \(j_i\in \mathbb {N}_{>0}\) such that intersection product induces a surjection

$$\begin{aligned} A^{j_1}V_{}\otimes A^{j_2}V_{}\otimes \cdots \otimes A^{j_r}V_{}\ \rightarrow \ A^jV_{}. \end{aligned}$$

Then for any \(a\in A^j(X\times X)\), there exists a decomposition

$$\begin{aligned} a= C_1\cdot \ldots \cdot C_r+\Gamma _1+\Gamma _2\ \ \in A^j(X\times X), \end{aligned}$$

where \(C_i\in A^{j_i}(X\times X)\), and \(\Gamma _1, \Gamma _2\) are supported on \(D\times X\) (resp. on \(X\times D\)), for some divisor \(D\subset X\).

Proof

To be sure, this is a variant of the argument of [4], exploiting the fact that \(\mathbb {C}\) is a universal domain. Let \(D_1\subset X\) denote the complement of V. Taking the smallest possible field of definition, we can suppose everything (X, V and the cycle a) is defined over a field \(k\subset \mathbb {C}\) which is finitely generated over its prime subfield. Then the inclusion \(k(X)\subset \mathbb {C}\) (which comes from \(\mathbb {C}\) being a universal domain) induces an injection

$$\begin{aligned} A^j(X_{k(X)})\ \rightarrow \ A^j(X_{\mathbb {C}}) \end{aligned}$$

[3, Appendix to Lecture 1]. On the other hand,

$$\begin{aligned} A^j(X_{k(X)})_{}= \varinjlim A^j(X\times U)_{}, \end{aligned}$$

where the limit is taken over opens \(U\subset X\) [3, Appendix to Lecture 1].

Given the cycle \(a\in A^j(X\times X)\), consider the restriction

$$\begin{aligned} a_{restr}\ \ \in A^j(X_{k(X)}). \end{aligned}$$

The assumption implies there exist cycles \(c_i\in A^{j_i}(X_{\mathbb {C}})\) such that

$$\begin{aligned} a_{restr}=c_1\cdot \ldots \cdot c_r+a_0\ \ \in A^j(X_{\mathbb {C}}), \end{aligned}$$

where \(a_0\) is supported on \(D_1\). Now, we extend k so that the cycles \(c_i\) are also defined over k (and k is still finitely generated over its prime subfield, so that \(k(X)\subset \mathbb {C}\)). Then using the injection \(A^j(X_{k(X)})\ \rightarrow \ A^j(X_{\mathbb {C}})\) cited above, we obtain the decomposition

$$\begin{aligned} a_{restr}=c_1\cdot \ldots \cdot c_r+a_0\ \ \in A^j(X_{k(X)}). \end{aligned}$$

Let \(C_i\in A^{j_i}(X\times X)\) be any cycle restricting to \(c_i\), and let \(\Gamma _1\) be any cycle restricting to \(a_0\). Then using the limit property cited above, we find that the difference

$$\begin{aligned} a-C_1\cdot \ldots \cdot C_r-\Gamma _1 \ \ \in A^j(X\times X) \end{aligned}$$

restricts to 0 in \(A^j(X\times U)\), for some open \(U\subset X\). This means there exists a divisor \(D_2\subset X\) and a cycle \(\Gamma _2\) supported on \(D_2\) such that

$$\begin{aligned} a= C_1\cdot \ldots \cdot C_r+\Gamma _1+\Gamma _2\ \ \in A^j(X\times X). \end{aligned}$$

Taking D a divisor containing both \(D_1\) and \(D_2\), this proves the proposition. \(\square \)

Lemma 2

Let \(f:Y\rightarrow X\) be a proper morphism of smooth projective varieties, where \(\dim X=n\) and \(\dim Y=m\). Let \(C\in A^j(X\times X)\). Then

$$\begin{aligned} \bigl ( (f\times \hbox {id})^*C\bigr )_*= C_*f_*:\ \ H^iY\ \rightarrow \ H^{i+2(j-m)}X. \end{aligned}$$

Proof

This is purely formal, and surely well-known. Let \(p_1, p_2:X\times X\rightarrow X\) denote projection on the first (resp. second) factor. Let \(q_1, q_2\) denote projections from \(Y\times X\) to Y (resp. to X). For \(a\in H^iY\), we have

$$\begin{aligned} \bigl ( (f\times \hbox {id})^*C\bigr )_*(a)&= (q_2)_*\bigl ( (q_1)^*(a)\cdot (f\times \hbox {id})^*(C)\bigr )\\&= (p_2)_*\bigl ( (f\times \hbox {id})_*((q_1)^*(a)\cdot (f\times \hbox {id})^*(C)) \bigr ) \\&=(p_2)_*\bigl ( (f\times \hbox {id})_*(q_1)^*(a)\cdot C\bigr )\\&=(p_2)_*\bigl ( (p_1)^*f_*(a)\cdot C\bigr )=: C_*f_*(a)\ \ \in H^{i+2(j-m)}X.\\ \end{aligned}$$

Lemma 3

Let \(f:Y\rightarrow X\) be as in Lemma 2. Let \(D\in A^j(Y\times X)\). Then

$$\begin{aligned} \bigl ( (f\times \hbox {id})_*D\bigr )_*= D_*f^*:\ \ H^iX\ \rightarrow \ H^{i+2(j-m)}X. \end{aligned}$$

Proof

Just as Lemma 2, this is surely well-known. Keeping the notation of Lemma 2, for \(b\in H^iX\) we have

$$\begin{aligned} \bigl ( (f\times \hbox {id})_*D\bigr )_*(b)&= (p_2)_*\bigl ( (p_1)^*(b)\cdot (f\times \hbox {id})_*(D)\bigr )\\&= (p_2)_*\bigl ( (f\times \hbox {id})_*\bigl ( (f\times \hbox {id})^*(p_1)^*(b)\cdot D\bigr )\bigr )\\&=(p_2)_*(f\times \hbox {id})_*\bigl ( (q_1)^*f^*(b)\cdot D\bigr )\\&= (q_2)_*\bigl ( (q_1)^*f^*(b)\cdot D\bigr )=: D_*f^*(b)\ \ \in H^{i+2(j-m)}X. \\ \end{aligned}$$

\(\square \)

Lemma 4

(Esnault–Srinivas–Viehweg [5]) Let X be a smooth projective variety of dimension n. Let \(C_i\in A^{j_i}(X\times X), i=1,\ldots ,r\), with \(j=j_1+\cdots +j_r\). Then

$$\begin{aligned}\ (C_1\cdot \ldots \cdot C_r)_*\hbox {Gr}_F^{n-j}H^{2n-j}(X,\mathbb {C}) \!\subset \!\hbox {Im}\Bigl ( H^{j_1}(X,\mathcal O_X)\!\otimes \!\cdots \!\otimes \!H^{j_r}(X,\mathcal O_X) \!\rightarrow \!H^j(X,\mathcal O_X)\Bigr )\!. \end{aligned}$$

Proof

This is shamelessly plagiarized from [5], who prove the \(j=n\) case. Let \( C\in A^j(X\times X)\) be any correspondence. The crucial observation is that the action

$$\begin{aligned} C_*:\ \ \hbox {Gr}_F^{n-j}H^{2n-j}(X,\mathbb {C})\ \rightarrow \ \hbox {Gr}_F^0 H^j(X,\mathbb {C}) \end{aligned}$$

only depends on the image of C under the composite map

$$\begin{aligned} \iota :\ \ A^j(X\times X)\ \rightarrow \ H^{2j}(X\times X,\mathbb {C})\ \rightarrow \ H^jX\otimes H^jX\ \rightarrow \ \hbox {Gr}_F^j H^jX\otimes \hbox {Gr}_F^0 H^jX\ \end{aligned}$$

(Here the second map is given by the Künneth decomposition, and the last map is induced by projection on the appropriate summands of the Hodge decomposition). Indeed, suppose \(C\in A^j(X\times X)\) is such that \(\iota (C)=0\), i.e. the Künneth part of type \(H^jX\otimes H^jX\) of C is contained in

$$\begin{aligned} \bigoplus _{i<j} \hbox {Gr}_F^i H^jX\otimes \hbox {Gr}_F^{j-i} H^jX\ \subset \ \hbox {Gr}_F^j (H^jX\otimes H^jX). \end{aligned}$$

Then, for \(a\in \hbox {Gr}_F^{n-j}H^{2n-j}(X,\mathbb {C})\) we find that

$$\begin{aligned} (p_1)^*(a)\cdot C\ \ \in \bigoplus _{i<j} \hbox {Gr}_F^{n-j+i} H^{2n}X\otimes \hbox {Gr}_F^{j-i}H^jX=0, \end{aligned}$$

and hence

$$\begin{aligned} C_*(a)=0\ \ \in H^j(X,\mathbb {C}). \end{aligned}$$

Next, we apply this observation to

$$\begin{aligned} C=C_1\cdot \ldots \cdot C_r\ \ \in A^j(X\times X), \end{aligned}$$

with \(C_i\in A^{j_i}X\). The Hodge decomposition then gives that

$$\begin{aligned} \iota (C)= & {} \iota (C_1)\cdot \ldots \cdot \iota (C_r)\in \ \hbox {Im}\bigl (\hbox {Gr}_F^{j_1} H^{j_1}X\otimes \cdots \otimes \hbox {Gr}_F^{j_r} H^{j_r}X\bigr )\\&\otimes \, \hbox {Im}\bigl (\hbox {Gr}_F^{0} H^{j_1}X\otimes \cdots \otimes \hbox {Gr}_F^{0} H^{j_r}X\bigr )\subset \hbox {Gr}_F^j H^jX\otimes \hbox {Gr}_F^0 H^jX . \end{aligned}$$

This proves the Lemma: suppose

$$\begin{aligned} \iota (C)=\sum _k C^k_{left}\otimes C^k_{right}\ \ \in \hbox {Gr}_F^j H^jX\otimes \hbox {Gr}_F^0 H^jX. \end{aligned}$$

Then reasoning as above, we find that

$$\begin{aligned} (\iota (C))_*(a)= \sum _k (p_2)_*\left( \left( a\cup C^k_{left}\right) \otimes C^k_{right} \right) =\sum _k \alpha _k C^k_{right}\ \ \in \hbox {Gr}_F^0 H^jX, \end{aligned}$$

where the \(\alpha _k\) are complex numbers (this is because \(H^{2n}X\) is one-dimensional and generated by the class of a point). \(\square \)

Remark 3

It is mainly the contrapositive of Theorem 1 that is useful (this is another remark made in [5] for their theorem). Indeed, suppose X and \(j=j_1+\cdots +j_r\) are such that

$$\begin{aligned} H^{j_1}(X,\mathcal O_X)\otimes H^{j_2}(X,\mathcal O_X)\otimes \cdots \otimes H^{j_r}(X,\mathcal O_X)\ \rightarrow \ H^j(X,\mathcal O_X)\ \end{aligned}$$

is not surjective (for example, because

$$\begin{aligned} \prod _{i=1}^r \dim H^{j_i}(X,\mathcal O_X)< \dim H^j(X,\mathcal O_X)\ ). \end{aligned}$$

Then by Theorem 1, likewise

$$\begin{aligned} A^{j_1}X_{}\otimes \cdots \otimes A^{j_r}X_{}\ \rightarrow \ A^jX_{}\ \end{aligned}$$

fails to be surjective (and the same holds for any open \(V\subset X\)).