1 Erratum to: Math. Z. (2012) 270:871–887 DOI 10.1007/s00209-010-0830-2

In the original publication [3], the Theorem 3.1 and Theorem 4.2 are erroneous.

  1. (i)

    First, the proof of Theorem 3.1 is incorrect: The fault is at the Step 2 of the proof. In the meantime, the result has been proved in [4] with better bounds.

  2. (ii)

    Second, I correct Theorem 4.2. At p. 884, the last row of the diagram A.1 should be tensored by \({\fancyscript{O}}_E(-2E)\). This error affects the subsequent computations from Lemma A.2 onward, which are used in the proof of the Theorem.

The corrected version is provided below.

2 The modified Wahl map

Recall that \(\hat{C}\in |{\fancyscript{L}}={\fancyscript{A}}^d|\) is a nodal curve with nodes \(\fancyscript{N}:=\{\hat{x}_1,\ldots ,\hat{x}_\delta \}\) on the polarized \(K3\) surface \((S,{\fancyscript{A}})\), such that \({\fancyscript{A}}\in \mathop {\mathrm{Pic}}\nolimits (S)\) is not divisible, \({\fancyscript{A}}^2=2(n-1)\). (Note that the article [3] deals only with \(K3\) surfaces with cyclic Picard group). Let \(\sigma :\tilde{S}\rightarrow S\) be the blow-up of \(S\) at \(\fancyscript{N}\), and denote by \(E^{a}\), \({a}=1,\ldots ,\delta \), the exceptional divisors, and \(E:=E^1+\cdots +E^\delta \). The normalization \(C\) of \(\hat{C}\) fits into

(1)

\(\tilde{u}\) is an embedding, and \(K_C=\sigma ^*{\fancyscript{L}}(-E)\otimes {\fancyscript{O}}_C\). The curve \(C\) carries the divisor

$$\begin{aligned} \varDelta :=x_{1,1}+x_{1,2}+\cdots +x_{\delta ,1}+x_{\delta ,2}, \end{aligned}$$

where \(\{x_{{a},1},x_{{a},2}\}=E^{a}\cap C\) is the pre-image of \(\hat{x}_a\in \hat{C}\) by \(\nu \).

In general, if \(V^{\prime }\) is a subscheme of some variety \(V\), \({\fancyscript{I}}_V(V^{\prime })\) or \({\fancyscript{I}}(V^{\prime })\) stands for its sheaf of ideals, and \(DV^{\prime }\subset V\times V\) denotes the diagonally embedded \(V^{\prime }\).

Let \((X,\varDelta _X)\) be an arbitrary smooth, irreducible curve together with \(\delta \) pairwise disjoint pairs of points \( \varDelta _X=\{\{x_{1,1},x_{1,2}\},\ldots ,\{x_{\delta ,1},x_{\delta ,2}\}\}\subset X. \) The exact sequence \(0\rightarrow {\fancyscript{I}}(DX)^2\rightarrow {\fancyscript{I}}(DX)\rightarrow K_X\rightarrow 0\) yields the Wahl map

$$\begin{aligned} w_X:H^0\left( X\times X,{\fancyscript{I}}(DX)\otimes K_{X\times X}\right) {\rightarrow } H^0\left( X,K_X^3\right) . \end{aligned}$$

The vector space \(H^0({\fancyscript{I}}(DX)\otimes K_{X\times X})\) splits into

$$\begin{aligned} H^0({\fancyscript{I}}(DX)\otimes K_{X\times X})\cap \mathop {\mathrm{Sym}}\nolimits ^2H^0(K_X) \oplus \mathop {\bigwedge }\limits ^{2}H^0(K_X), \end{aligned}$$

and \(w_X\) vanishes on the first direct summand, as it is skew-symmetric. Denote

$$\begin{aligned} {P}_{\varDelta _X}:={\mathop {\mathop {\bigcup }\limits ^{\delta }}\limits _{a=1}} \{x_{a,1},x_{a,2}\}\times \{x_{a,1},x_{a,2}\}\subset X\times X, \end{aligned}$$
(2)

and let \(w_{X,\varDelta _X}\) be the restriction of \(w_X\) to \(H^0\big ({\fancyscript{I}}({P}_{\varDelta _X})\cdot {\fancyscript{I}}(DX)\otimes K_{X\times X}\big ) \cap \mathop {\bigwedge }\limits ^{2}H^0(K_X)\). (Thus, \(w_{X,\varDelta _X}\) is a punctual modification of the usual Wahl map.) With this notation, we replace [3], Theorem4.2] by the following.

Theorem 1

(i) Let \((S,{\fancyscript{A}})\), \({\fancyscript{A}}^2\geqslant 6\), be as above. Consider a nodal curve \(\hat{C}\,{\in }\,|d{\fancyscript{A}}|\) with

$$\begin{aligned} \delta \leqslant \min \left\{ \frac{d^2{\fancyscript{A}}^2}{3(d+4)}, \delta _\mathrm{max}(n,d)\right\} \end{aligned}$$
(3)

nodes and let \((C,\varDelta )\) be as above (\(\delta _\mathrm{max}(n,d)\) is defined in [3], p. 872]; the minimum is the first expression, except a finite number of cases). Then, the homomorphism \(w_{C,\varDelta }\) is not surjective.

(ii) For generic a generic curve \(X\) of genus \(g\geqslant 12\) with generic markings \(\varDelta _X\), such that \(\delta \leqslant \frac{g-1}{2}\), the homomorphism \(w_{X,\varDelta _X}\) is surjective.

This is a nonsurjectivity property for the pair \((C,\varDelta )\), rather than for \(C\) itself. I conclude the note (see Sect. 4) with some evidence toward the nonsurjectivity of the Wahl map \(w_C\) itself, and comment on related work in [4].

3 Relationship between the Wahl maps of \(C\) and \({\tilde{S}}\)

Lemma 2

  • (i) The following diagram has exact rows and columns:

    (4)

    In other words, we have \({\fancyscript{I}}:= {\fancyscript{I}}_{C\times C}(D\varDelta )\cdot {\fancyscript{I}}_{C\times C}(DC)=w_C^{-1}(K_C(-\varDelta ))\). Also, the involution \(\tau _C\) which interchanges the factors of \(C\times C\) leaves \({\fancyscript{I}}\) invariant.

  • (ii) \(H^0\bigl ( C\times C, {\fancyscript{I}}\otimes K_{C\times C}) \bigr ) = w_C^{-1}\bigl (\;H^0(C,K_C^3(-\varDelta ))\;\bigr )\).

  • (iii) For \( \varLambda := \bigl \{ s-\tau _C^*(s)\mid s\in H^0\bigl ( {\fancyscript{I}}\otimes K_{C\times C}) \bigr ) \bigr \} \) holds

    $$\begin{aligned} \varLambda \mathop {=}\limits ^{(\star )} H^0\bigl ({\fancyscript{I}}\otimes K_{C\times C}\bigr ) \;\cap \; {\mathop {\bigwedge }\limits ^{2}}\,H^0(K_C) \mathop {\subset }\limits ^{(\star \star )} H^0\bigl ({\fancyscript{I}}_{C\times C}(DC)\otimes K_{C\times C}\bigr ). \end{aligned}$$

    (iv) \( w^{\prime }_{C}(\varLambda ) = w^{\prime }_{C}\bigl (\; H^0\bigl ({\fancyscript{I}}\otimes K_{C\times C}\bigr )\bigr ). \)

Proof

  1. (i)

    The middle column is exact because \({\fancyscript{I}}_{C\times C}(DC)\) is locally free. We check the exactness of the first row around each point \((o,o)\in D\varDelta \). Let \(u\) be a local (analytic) coordinate on \(C\) such that \(o=0\), and \(u_1,u_2\) be the corresponding coordinates on \(C\times C\). Then, the first row becomes \( 0\rightarrow \langle u_2-u_1\rangle ^2\rightarrow \langle u_2-u_1\rangle \cdot \langle u_1,u_2\rangle \rightarrow \langle u\rangle \cdot \mathrm{d}u\rightarrow 0, \) with \(\mathrm{d}u:=(u_2-u_1)\mathrm{mod} (u_2-u_1)^2\), which is exact. The second statement is obvious.

  2. (ii)

    We tensor (4) by \(K_{C\times C}\), and take the sections in the last two columns. An elementary diagram chasing yields the claim.

  3. (iii)

    Let us prove \((\star )\). The vector space \(\varLambda \) is contained in \({\mathop {\bigwedge }\limits ^{2}}\,H^0(K_C)\) by the very definition, and also in \(H^0\bigl ({\fancyscript{I}}\otimes K_{C\times C}\bigr )\) because the sheaf \({\fancyscript{I}}(C,\varDelta )\) is \(\tau _C\)-invariant. For the inclusion in the opposite direction, take \(s\) in the intersection. As \(s\in {\mathop {\bigwedge }\limits ^{2}}\;H^0(K_C)\), it follows \(\tau _C^*(s)=-s\), so \(s=1/2\cdot (s-\tau _C^*(s))\in \varLambda \). The inclusion \((\star \star )\) is obvious.

  4. (iv)

    Indeed, the Wahl map is anti-commutative: \(w_C({\sum }_{i}s_i\otimes t_i)=-w_C({\sum }_{i}t_i\otimes s_i)\).

\(\square \)

Lemma 3

Let \(\varXi :=\{(x_{{a},1},x_{{a},2}),(x_{{a},2},x_{{a},1})\mid {a}=1,\ldots ,\delta \}\subset C\times C\), and consider the sheaf of ideals \({\fancyscript{I}}(C,\varDelta ):={\fancyscript{I}}(\varXi )\cdot {\fancyscript{I}} \mathop {=}\limits ^{(2)} {\fancyscript{I}}({P}_\varDelta )\cdot {\fancyscript{I}}(DC) \subset {\fancyscript{I}}.\) Furthermore, denote

$$\begin{aligned} \varLambda (\varDelta ):=H^0({\fancyscript{I}}(C,\varDelta )\otimes K_{C\times C}) \,\cap \, \mathop {\bigwedge }\limits ^{2}\,H^0(K_C). \end{aligned}$$
(5)

Then, the following statements hold:

  • (i) \({\fancyscript{I}}(C,\varDelta )\) is \(\tau _C\)-invariant, so \(w^{\prime }_C\bigl (\,\varLambda (\varDelta )\,\bigr ) =w^{\prime }_C\bigl (\;H^0({\fancyscript{I}}(C,\varDelta )\otimes K_{C\times C})\;\bigr ).\)

  • (ii) \({\fancyscript{I}}(C,\varDelta )+{\fancyscript{I}}_{C\times C}(DC)^2={\fancyscript{I}}\), and \({\fancyscript{I}}(C,\varDelta )\cap {\fancyscript{I}}(DC)^2= {\fancyscript{I}}(\varXi )\cdot {\fancyscript{I}}(DC)^2\). Therefore, the various sheaves introduced so far fit into the commutative diagram

    (6)

Proof

  1. (i)

    The proof is identical to Lemma 2(iv).

  2. (ii)

    The inclusion \(\subset \) is clear. For the reverse, notice that \({\fancyscript{O}}_{C\times C}={\fancyscript{I}}(\varXi )+{\fancyscript{I}}\), so \({\fancyscript{I}}\subset {\fancyscript{I}}(C,\varDelta )+{\fancyscript{I}}^2 \subset {\fancyscript{I}}(C,\varDelta )+{\fancyscript{I}}(DC)^2\). The second claim is analogous.

\(\square \)

Now, we compare the Wahl maps of \(C\) and \({\tilde{S}}\). Let \(\rho :\!{\fancyscript{I}}_{\tilde{S}\times \tilde{S}}(D\tilde{S})\rightarrow {\fancyscript{I}}_{C\times C}(DC)\) be the restriction homomorphism, and \({\fancyscript{M}}:=\sigma ^*{\fancyscript{L}}(-E)\). The diagram below relates various objects involved in the definition of \(w_C\) and \(w_{\tilde{S}}\):

(7)

The rightmost column corresponds to the first-order expansions of the sections along \(E\) and at \(\varDelta \). By using \(0{\rightarrow }{\fancyscript{O}}_E(1){\rightarrow }{\fancyscript{O}}_{2E}{\rightarrow }{\fancyscript{O}}_E{\rightarrow }0\), we deduce that it fits into:

(8)

Along each \(E^{a}\subset \tilde{S}\), we consider local coordinates \(u,v\) as follows: \(v\) is the coordinate along \(E^{a}\), and \(u\) is a coordinate in the normal direction to \(E^{a}\) (so \(E^{a}\) is given by \(\{u=0\}\)). Moreover, we assume that \(C\) is given by \(\{v=0\}\) around the intersection points \(\{x_{{a},1},x_{{a},2}\}=E^{a}\cap C\). Then, any element \(\tilde{s}\in H^0({\fancyscript{M}})\) can be expanded as

$$\begin{aligned} \tilde{s}=\tilde{s}^{a}_0(v)+u\tilde{s}^{a}_1(v)+O\left( u^2\right) , \end{aligned}$$
(9)

and its image in \(H^0({\fancyscript{M}}\otimes {\fancyscript{O}}_{2E^{a}})\) is \(\tilde{s}^{a}_0(v)+u\tilde{s}^{a}_1(v)\). Finally, observe that the values of \(w_{\tilde{S}}\) are sections of \(\varOmega ^1_{\tilde{S}}\otimes {\fancyscript{M}}^2\), and the restriction of this latter to \(E\) fits into

$$\begin{aligned} 0\rightarrow \underbrace{{\fancyscript{O}}_E(3)} _{\mathrm{normal~component}} \rightarrow \varOmega ^1_{\tilde{S}}\otimes {\fancyscript{M}}^2|_{E}\rightarrow \underbrace{\varOmega ^1_E\otimes {\fancyscript{M}}^2_E={\fancyscript{O}}_E} _{\mathrm{tangential~component}} \rightarrow 0. \end{aligned}$$

Lemma 4

Let the notation be as in (9). We consider \(\tilde{e}= {\sum }_{i}\,\tilde{s}_i\wedge \tilde{t}_i\in \mathop {\bigwedge }\limits ^{2}\,H^0({\fancyscript{M}})\), and let \(e:=\rho (\tilde{e})={\sum }_{i}\, s_i\wedge t_i\). Then, the following statements hold:

  • (i) \(w_{\tilde{S}}(\tilde{e})\in H^0(\varOmega ^1_{\tilde{S}}(-E)\otimes {\fancyscript{M}}^2)\) if and only if:

    $$\begin{aligned} \left\{ \begin{array}{ccl} (\star ) &{}\quad &{} \mathop {\sum }\limits _{i}\, \bigl ( s_{i}(x_{{a},1}) t_{i}(x_{{a},2}) - t_{i}(x_{{a},1}) s_{i}(x_{{a},2}) \bigr ) =0,\quad \text {and}\\ (\star \star ) &{}\quad &{} \mathop {\sum }\limits _{i}\, \bigl ( \tilde{s}_{i,0}^{a}\tilde{t}_{i,1}^{a}-\tilde{t}_{i,0}^{a}\tilde{s}_{i,1}^{a}\bigr ) =0,\quad \forall \,a=1,\ldots ,\delta . \end{array} \right. \end{aligned}$$
  • (ii) \(\varLambda (E):= w_{\tilde{S}}^{-1}\bigl (H^0(\varOmega ^1_{\tilde{S}}(-E)\otimes {\fancyscript{M}}^2)\bigr ) \,\cap \,\mathop {\bigwedge }\limits ^{2}H^0({\fancyscript{M}})\) has the property

    $$\begin{aligned} w_{\tilde{S}}(\varLambda (E)) = w_{\tilde{S}}\bigl (\; w_{\tilde{S}}^{-1}\bigl (H^0(\varOmega ^1_{\tilde{S}}(-E) \otimes {\fancyscript{M}}^2)\bigr )\bigr ). \end{aligned}$$
  • (iii) \(\rho \big (\varLambda (E)\big )\subset \varLambda (\varDelta )\), where the right hand side is defined by (5).

Proof

  1. (i)

    The element \(w_{\tilde{S}}(\tilde{e})\) vanishes along \(E\) if and only if both its tangential and normal components along each \(E^{a}\subset E\) vanish. A short computation shows that the normal component is \((\star \star )\). The tangential component is \({\sum }_{i}\left( \tilde{s}^{a}_{i,0}\left( \tilde{t}^{a}_{i,0}\right) ^{\prime }-\tilde{t}^{a}_{i,0}\left( \tilde{s}^{a}_{i,0}\right) ^{\prime } \right) \). But \(\tilde{s}^{a}_{i,0},\tilde{t}^{a}_{i,0}\in H^0({\fancyscript{O}}_{E^{a}}(1)),\) that is they are linear polynomials in \(v\), so

    $$\begin{aligned} \tilde{s}^{a}_{i,0}\left( \tilde{t}^{a}_{i,0}\right) ^{\prime }-\tilde{t}^{a}_{i,0}\left( \tilde{s}^{a}_{i,0}\right) ^{\prime } = \tilde{s}^{a}_{i,0}\left( x_{{a},1}\right) \tilde{t}^{a}_{i,0}\left( x_{{a},2}\right) -\tilde{t}^{a}_{i,0}\left( x_{{a},1}\right) \tilde{s}^{a}_{i,0}\left( x_{{a},2}\right) , \end{aligned}$$

    up to a constant factor. Also, we have \(\tilde{s}^{a}_{i,0}(x_{{a},j})=\tilde{s}^{a}(x_{{a},j}) =s(x_{{a},j})\), and \((\star )\) follows.

  2. (ii)

    The vector space \(w_{\tilde{S}}^{-1}\bigl (H^0(\varOmega ^1_{\tilde{S}}(-E) \otimes {\fancyscript{M}}^2)\bigr )\) is invariant under the involution \(\tau _{\tilde{S}}\) of \(\tilde{S}\times \tilde{S}\) which switches the two factors. As \(w_{\tilde{S}}\) is anti-commutative, the claim follows as in Lemma 2.

  3. (iii)

    Take \(\tilde{e}\in \varLambda (E)\) and \(e:=\rho (\tilde{e})\). Then, \(e(x_{{a},1},x_{{a},2})=-e(x_{{a},2},x_{{a},1})\mathop {=}\limits ^{(\star )}0\), and also \(w_C(e)(x_{{a},j})\) equals the expression \((\star \star )\) at \(x_{{a},j}\) (so it vanishes), for \(j=1,2\). \(\square \)

Now. we consider the commutative diagram:

(10)

It is the substitute in the case of nodal curves for [3], diagram (4.2)].

Lemma 5

Assume \(\mathop {\mathrm{Pic}}\nolimits (S)=\mathbb {Z}{\fancyscript{A}}\). Then, \(\rho _\varDelta :\varLambda (E)\rightarrow \varLambda (\varDelta )\) is surjective.

Proof

The restriction \(H^0({\tilde{S}},{\fancyscript{M}})\rightarrow H^0(C,K_C)\) is surjective (see [3], LemmaA.1]), and the kernel of \( \mathop {\bigwedge }\limits ^{2}\,H^0(\tilde{S},{\fancyscript{M}})\rightarrow \mathop {\bigwedge }\limits ^{2}\,H^0(C,K_C) \) consists of elements of the form \(\tilde{t}\wedge (\tilde{s}_C\tilde{s}_E)\), where \(\tilde{t}\in H^0({\fancyscript{M}})\) and \(\tilde{s}_C,\tilde{s}_E\) are the canonical sections of \({\fancyscript{O}}_{\tilde{S}}(C)\) and \({\fancyscript{O}}_{\tilde{S}}(E)\), respectively. (See the middle column of (7).)

Consider \(e={\sum }_{i}\;(s_i\otimes t_i-t_i\otimes s_i) \in \varLambda (\varDelta )\), and let \(\tilde{e}={\sum }_{i}\;(\tilde{s}_i\otimes \tilde{t}_i-\tilde{t}_i\otimes \tilde{s}_i) \in \mathop {\bigwedge }\limits ^{2}\,H^0({\fancyscript{M}})\) be such that \(\rho (\tilde{e})=e\). The proof of 4(i) shows that, for all \({a}\), the tangential component of \(w_{\tilde{S}}(\tilde{e})|_{E^{a}}\) equals \(e(x_{{a},1},x_{{a},2})=0\), so \(w_{\tilde{S}}(\tilde{e})|_{E}\) is a section of \(\varOmega ^1_{E/{\tilde{S}}}\otimes {\fancyscript{M}}_E^2\cong {\fancyscript{O}}_E(3)\). Since \(w_{\tilde{S}}(\tilde{e})|_{E}\) vanishes at the points of \(\varDelta \), it is actually determined up to an element in \(H^0({\fancyscript{O}}_E(1))\). We claim that this latter can be canceled by adding to \(\tilde{e}\) a suitable element of the form \(\tilde{t}\wedge (\tilde{s}_C\tilde{s}_E)\). A short computation yields

$$\begin{aligned} w_{\tilde{S}}\left( \,\tilde{t}\wedge \left( \tilde{s}_C\tilde{s}_E\right) \right) |_E = \tilde{t}_E\cdot \left( \tilde{s}_C|_E\right) \cdot \left( \mathrm{d}\tilde{s}_E\right) |_E\in {\fancyscript{O}}_E(3), \end{aligned}$$

where \(\tilde{t}_E\in H^0({\fancyscript{O}}_E(1))\), \(\tilde{s}_C|_E\in H^0({\fancyscript{O}}_E(2))\) vanishes at \(\varDelta =E\cap C\), and \((\mathrm{d}\tilde{s}_E)|_E\in H^0({\fancyscript{O}}_E)\) (it is a section of \(\varOmega ^1_{\tilde{S}}|_E\) with vanishing tangential component). Thus, these two latter factors are actually (nonzero) scalars.

The previous discussion shows that \(\tilde{e}+\tilde{t}\wedge (\tilde{s}_C\tilde{s}_E)\in \varLambda (E)\) as soon as \(\tilde{t}\in H^0({\fancyscript{M}})\) satisfies \(\tilde{t}_E=-w_{\tilde{S}}(\tilde{e})|_E\in H^0({\fancyscript{M}}_E)\). According to Corollary 8, such an element \(\tilde{t}\) exists because the restriction \(H^0({\fancyscript{M}})\rightarrow H^0({\fancyscript{M}}_E)\) is surjective. \(\square \)

Proof of Theorem 1

(i) Case \(\mathop {\mathrm{Pic}}\nolimits (S)=\mathbb {Z}{\fancyscript{A}}\). If \(w_{C,\varDelta }\) is surjective; then, the homomorphism \(b\) in the diagram (10) is surjective too. Now, we follow the same pattern as in [3], p. 884,top]: \(b\) is the restriction homomorphism at the level of sections of

$$\begin{aligned} 0\rightarrow K_C\rightarrow \varOmega ^1_{\tilde{S}}\bigr |_C\otimes K_C^2(-\varDelta ) \rightarrow K_C^3(-\varDelta )\rightarrow 0, \end{aligned}$$

and its surjectivity implies that this sequence splits. This contradicts [3], Lemma4.1].

General case. It is a deformation argument. We consider

$$\begin{aligned} \begin{array}{l} {\fancyscript{K}}_n:= \left\{ \left( S,{\fancyscript{A}}\right) \mid {\fancyscript{A}}\in \mathop {\mathrm{Pic}}\nolimits (S)\text { is ample, not divisible},{\fancyscript{A}}^2=2(n-1)\right\} ,\\ {\fancyscript{V}}^d_{n,\delta }:=\left\{ \left( (S,{\fancyscript{A}}),\hat{C} \right) \mid \left( S,{\fancyscript{A}}\right) \in {\fancyscript{K}}_n,\; \hat{C}\in |d{\fancyscript{A}}|\text { nodal curve with }\delta \text { nodes} \right\} . \end{array} \end{aligned}$$

Then, the natural projection \(\kappa :{\fancyscript{V}}^d_{n,\delta }\rightarrow {\fancyscript{K}}_n\) is submersive onto an open subset of \({\fancyscript{K}}_n\). (See [3], Theorem 1.1(iii)] and the reference therein.)

Hence, for any \(((S,{\fancyscript{A}}),\hat{C})\in {\fancyscript{V}}^d_{n,\delta }\) there is a smooth deformation \(((S_t,{\fancyscript{A}}_t),\hat{C}_t)\) parameterized by an open subset \(T\subset {\fancyscript{K}}_n\). The points \(t\in T\) such that \(\mathop {\mathrm{Pic}}\nolimits (S_t)=\mathbb {Z}{\fancyscript{A}}_t\) are dense; for these \(w_{C_t,\varDelta _t}\) are nonsurjective. Since the nonsurjectivity condition is closed, we deduce that \(w_{C,\varDelta }\) is nonsurjective too.

(ii) Now let \((X,\varDelta _X)\) be a generic marked curve of genus at least 12. By [1], the Wahl map \(w_X\,{:}\,\mathop {\bigwedge }\limits ^{2}\,H^0(K_X)\rightarrow H^0(K_X^3)\) is surjective; thus, \(\widetilde{w}\,^{\prime }_X\,{:=}\,H^0(w^{\prime }_X\otimes K_{X\times X})\) in (6) is surjective as well (see lemma 2(ii)). As \(\delta \leqslant \frac{g-1}{2}\), the evaluation homomorphism \(H^0(K_X)\rightarrow K_{X}\otimes {\fancyscript{O}}_{\varDelta _X}\) is surjective for generic markings, so the same holds for

$$\begin{aligned} H^0(K_X)^{\otimes 2}\rightarrow \mathop {\mathop {\bigoplus }\limits ^{\delta }}\limits _{a=1}\left( K_{X,x_{a,1}}\oplus K_{X,x_{a,2}}\right) ^{\otimes 2}. \end{aligned}$$

The restriction to the anti-symmetric part (on both sides) yields the surjectivity of

$$\begin{aligned} \mathrm{ev}_\varXi :\mathop {\bigwedge }\limits ^{2}H^0(K_X)\rightarrow \mathop {\mathop {\bigoplus }\limits ^{\delta }}\limits _{a=1}K_{X,x_{a,1}}\otimes K_{X,x_{a,2}} =\mathop {\mathop {\bigoplus }\limits ^{\delta }}\limits _{a=1} K_{X\times X, \left( x_{a,1},\,x_{a,2}\right) }. \end{aligned}$$

(For \(s\in \mathop {\bigwedge }\limits ^{2}H^0(K_X)\), \(\mathrm{ev}_\varXi (s)\) takes opposite values at \((x_{a,1},x_{a,2})\) and \((x_{a,2},x_{a,1})\).)

The diagram (6) yields

(11)

A straightforward diagram chasing shows that \(w_{X,\varDelta _X}\) is surjective if

$$\begin{aligned} \mathrm{ev}_\varXi ^{\prime \prime }: \underbrace{H^0\bigl (\,{\fancyscript{I}}(DX)^2\cdot K_{X\times X}\,\bigr ) \,\cap \, {\mathop {\bigwedge }\limits ^{2}}\,H^0(K_X)}_{:=G} \rightarrow \underbrace{ {\mathop {\mathop {\bigoplus }\limits ^{\delta }}\limits _{a=1}} K_{X\times X,(x_{a,1},\,x_{a,2})}}_{:=H_{\varXi }} \end{aligned}$$

is so, or equivalently when the induced \(h_\varXi :\mathop {\bigwedge }\limits ^{\delta }\,G\rightarrow \mathop {\bigwedge }\limits ^{\delta }\,H_\varXi \) is nonzero. This is indeed the case for generic markings.

Claim \({\bigcap }_{\varDelta _X}\mathrm{Ker}(h_\varXi )=0\). (\(h_\varXi \) depends on \(\varDelta _X\).) Indeed, since \(\dim G\geqslant \delta \), we have

$$\begin{aligned} 0\ne \mathop {\bigwedge }\limits ^{\delta }\,G \subset \mathop {\bigwedge }\limits ^{\delta }\bigl (\mathop {\bigwedge }\limits ^{2}H^0(K_X)\bigr ) \subset H^0(K_{X\times X})^{\otimes \delta } = H^0\big ((X^2)^{\delta },K_{X\times X}\boxtimes \cdots \boxtimes K_{X\times X}\big ). \end{aligned}$$

The wedge is a direct summand of the tensor product (appropriate skew-symmetric sums), and \(h_\varXi \) is induced by the evaluation map

$$\begin{aligned} \mathrm{ev}^{\delta }: H^0\big ((X^2)^\delta ,K_{X\times X}\boxtimes \cdots \boxtimes K_{X\times X}\big ) \otimes {\fancyscript{O}} \rightarrow K_{X\times X}\boxtimes \cdots \boxtimes K_{X\times X} \end{aligned}$$

at \(\big ((x_{1,1},x_{1,2}),\ldots ,(x_{\delta ,1},x_{\delta ,2})\big )\in (X^2)^\delta \). If \(e\in \mathop {\bigwedge }\limits ^{\delta }G\) belongs to the intersection above, then \(e\in H^0\big (\mathrm{Ker}(\mathrm{ev}^{\delta })\big )=\{0\}\). Hence, for any \(e_1,\ldots ,e_\delta \in G\) with \(e_1\wedge \cdots \wedge e_\delta \ne 0\), there are markings \(\varDelta _X\) such that \(\mathrm{ev}_\varXi ^{\prime \prime }(e_1),\ldots ,\mathrm{ev}_\varXi ^{\prime \prime }(e_\delta )\) are linearly independent in \(H_\varXi \) (thus, they span it). \(\square \)

4 Multiple point Seshadri constants of \(K3\) surfaces with cyclic Picard group

This section is independent of the rest. Here we determine a lower bound for the multiple point Seshadri constants of \({\fancyscript{A}}\), which is necessary for proving Lemma 5.

Definition 6

(See [2], Section 6] for the original definition) The multiple point Seshadri constant of \({\fancyscript{A}}\) corresponding to \(\hat{x}_1,\ldots ,\hat{x}_\delta \in S\) is defined as

$$\begin{aligned} \varepsilon =\varepsilon _{S,\delta }({\fancyscript{A}}):=\inf _{Z} \frac{Z\cdot {\fancyscript{A}}}{{{\sum _{a=1}^{\delta }}\mathrm{mult}_{\hat{x}_a}(Z)}} = \sup \bigl \{ c\in \mathbb {R}\mid \sigma ^*{\fancyscript{A}}-cE\text { is ample on }{\tilde{S}}\bigr \}. \end{aligned}$$
(12)

The infimum is taken over all integral curves \(Z\subset S\) which contain at least one of the points \(\hat{x}_a\) above. Throughout this section, we assume that \(Z\in |z{\fancyscript{A}}|\), with \(z\geqslant 1\).

As the self-intersection number of any ample line bundle is positive, the upper bound \(\varepsilon \leqslant \frac{\sqrt{{\fancyscript{A}}^2}}{\sqrt{\delta }}\) is automatic. We are interested in finding a lower bound.

Theorem 7

Assume that \(\mathop {\mathrm{Pic}}\nolimits (S)=\mathbb {Z}{\fancyscript{A}}\), \({\fancyscript{A}}^2=2(n-1)\geqslant 4\), and \(\delta \geqslant 1\). Then, the Seshadri constant (12) satisfies \(\varepsilon \geqslant \frac{2{\fancyscript{A}}^2}{\delta +\sqrt{\delta ^2+4\delta (2+{\fancyscript{A}}^2)}},\) for any points \(\hat{x}_1,\ldots ,\hat{x}_\delta \in S\).

Our proof is inspired from [5], which treats the case \(\delta =1\).

Proof

We may assume that the points are numbered such that

$$\begin{aligned} \mathrm{mult}_{\hat{x}_a}(Z)\geqslant 2,\text { for }a=1,\ldots ,\alpha , \quad \mathrm{mult}_{\hat{x}_a}(Z)=1,\text { for }a=\alpha +1,\ldots ,\beta , \quad (\beta \leqslant \delta ). \end{aligned}$$

We denote \(p:=\mathop {\mathop {\sum }\limits ^{\alpha }}\limits _{a=1}\mathrm{mult}_{\hat{x}_a}(Z)\geqslant 2\alpha \) and \(m:=\mathop {\mathop {\sum }\limits ^{\delta }}\limits _{a=1}\mathrm{mult}_{\hat{x}_a}(Z)\leqslant p+\delta -\alpha \).

If \(\alpha =0\), then \(\frac{z\cdot {\fancyscript{A}}^2}{m}\geqslant \frac{{\fancyscript{A}}^2}{\delta }\) satisfies the inequality, so we may assume \(\alpha \geqslant 1\). A point of multiplicity \(m\) lowers the arithmetic genus of \(Z\) by at least \(\bigl (\begin{array}{c}{m}\\ {2}\end{array}\bigr )\); hence,

$$\begin{aligned} p_a(Z)=\frac{z^2{\fancyscript{A}}^2}{2}+1\geqslant \frac{1}{2}\mathop {\mathop {\sum }\limits ^{\alpha }}\limits _{a=1} \bigl (\mathrm{mult}_{\hat{x}_a}(Z)^2-\mathrm{mult}_{\hat{x}_a}(Z)\bigr ) \mathop {\mathop {\geqslant }\limits ^{\mathrm{Jensen}}}\limits _{\mathrm{inequality}} \frac{1}{2}\bigl (\frac{p^2}{\alpha }-p\bigr ), \end{aligned}$$

so \(p\leqslant \frac{\alpha +\sqrt{\alpha ^2+4\alpha (2+z^2{\fancyscript{A}}^2)}}{2}\). We deduce the following inequalities:

$$\begin{aligned} \frac{z{\fancyscript{A}}^2}{m}\geqslant & {} \frac{z{\fancyscript{A}}^2}{p-\alpha +\delta }\geqslant \underbrace{ \frac{z{\fancyscript{A}}^2}{\delta +\frac{\sqrt{\alpha ^2+4\alpha \left( 2+z^2{\fancyscript{A}}^2\right) }-\alpha }{2}} }_{\mathrm{decreasing~in}\,\alpha } \geqslant \underbrace{ \frac{z{\fancyscript{A}}^2}{\delta +\frac{\sqrt{\delta ^2+4\delta \left( 2+z^2{\fancyscript{A}}^2 \right) }-\delta }{2}} }_{\mathrm{increasing~in}\,z} \\\geqslant & {} \frac{2{\fancyscript{A}}^2}{\delta +\sqrt{\delta ^2+4\delta \left( 2+{\fancyscript{A}}^2\right) }}. \end{aligned}$$

\(\square \)

Corollary 8

\(H^0({\fancyscript{M}})\rightarrow H^0({\fancyscript{M}}_E)\) is surjective, for \({\fancyscript{A}}^2\geqslant 6\) and \(\delta \leqslant \frac{d^2{\fancyscript{A}}^2}{3(d+4)}\).

Proof

Indeed, it is enough to check that \(H^1({\tilde{S}},{\fancyscript{M}}(-E))= H^1({\tilde{S}},K_{\tilde{S}}\otimes {\fancyscript{M}}(-2E))\) vanishes. By the Kodaira vanishing theorem, this happens as soon as \({\fancyscript{M}}(-2E)=\sigma ^*{\fancyscript{A}}^d(-3E)\) is ample. The previous theorem implies that, in order to achieve this, is enough to impose \(\frac{3}{d}\leqslant \frac{2{\fancyscript{A}}^2}{\delta +\sqrt{\delta ^2+4\delta (2+{\fancyscript{A}}^2)}}\), which yields \(\delta \leqslant \frac{d^2({\fancyscript{A}}^2)^2}{3(d{\fancyscript{A}}^2+3{\fancyscript{A}}^2+6)}\). \(\square \)

5 Concluding remarks

(I) Evidence for the nonsurjectivity of \(\varvec{w_C}\) Theorem 1 is a nonsurjectivity property for the Wahl map of the pointed curve \((C,\varDelta )\), rather than that of the curve \(C\) itself.

Claim. In order to prove the nonsurjectivity of the Wahl map \(w_C\), is enough to have the surjectivity of the evaluation homomorphism

$$\begin{aligned} { H^0({\fancyscript{I}}(DC)^2\otimes K_{C\times C}) \rightarrow \mathop {\mathop {\bigoplus }\limits ^{\delta }}\limits _{a=1} K_{C\times C,(x_{a,1},x_{a,2})} \oplus K_{C\times C,(x_{a,2},x_{a,1})}. } \end{aligned}$$
(13)

(For \(\delta \) in the range (3), corollary 8 implies that \(K_C={\fancyscript{M}}_C\) separates \(\varDelta \), consequently \(\mathop {\bigwedge }\limits ^{2}H^0(K_C)\rightarrow {\bigoplus }_{a=1}^{\delta } K_{C\times C,(x_{a,1},x_{a,2})}\) is surjective. The surjectivity of (13) yields that of \(\mathrm{ev}_\varXi ^{\prime \prime }\) in (11), which is relevant for us.)

For the claim, observe that one has the following implications (see (4), (11)):

$$\begin{aligned} w_C\text { surjective }\;\Rightarrow \; w_C^{\prime }\text { surjective } \;\mathop {\mathop {\Rightarrow }\limits ^{(13)}}\limits _\mathrm{surj.}\; w_{C,\varDelta }\text { surjective, a contradiction.} \end{aligned}$$
(14)

The surjectivity of (13) is clearly a positivity property for \({\fancyscript{I}}(DC)^2\otimes K_{C\times C}\). We use again the Seshadri constants to argue why this is likely to hold. The \(\varXi \)-pointed Seshadri constants of the self-product of a very general curve \(X\) at very general points \(\varXi \) (as in Lemma 3) satisfy (see [6], p. 65 below Theorem 1.6, and Lemma 2.6]):

$$\begin{aligned} \varepsilon _{X\times X\!,\,\varXi }({\fancyscript{I}}(DX)^2\otimes K_{X\times X})\geqslant & {} 2(g-2)\varepsilon _{\mathbb {P}^2\!,\,g+\delta } ({\fancyscript{O}}_{\mathbb {P}^2}(1))\nonumber \\> & {} \frac{2(g-2)}{\sqrt{g+\delta }}\sqrt{1-\frac{1}{8(g+\delta )}}, \end{aligned}$$
(15)
$$\begin{aligned} \varepsilon _{X\times X\!,\,\varXi }({\fancyscript{I}}(DX)^4\otimes K_{X\times X})\geqslant & {} 4\cdot \frac{g-3}{2}\cdot \varepsilon _{\mathbb {P}^2\!,\,g+\delta } ({\fancyscript{O}}_{\mathbb {P}^2}(1))\nonumber \\> & {} \frac{2(g-3)}{\sqrt{g+\delta }}\sqrt{1-\frac{1}{8(g+\delta )}} =:\varphi (g,\delta ). \end{aligned}$$
(16)

The equation (16) implies (see [2], Proposition6.8]) that \({\big ({\fancyscript{I}}(DX)^2\otimes K_{X\times X}\big )}^2\) generates the jets of order \(\lfloor \varphi (g,\delta )\rfloor -2\) at \(\varXi \subset X\times X\). (We only need the generation of jets of order zero for \({\fancyscript{I}}(DX)^2\otimes K_{X\times X}\); also, note that \(\varphi (g,\delta )\) grows linearly with \(\sqrt{g}\) as long as \(\delta \) is small compared with \(g\) (see (3)).) This discussion suggests that \({\fancyscript{I}}(DX)^2\otimes K_{X\times X}\) is ‘strongly positive/generated.’ However, the passage to (13) above requires even more control.

(II) Related work In [4], the author extensively studies the properties of nodal curves on \(K3\) surfaces. Among several other results, he proves the nonsurjectivity of a marked Wahl map (different from the one introduced in here) for nodal curves on \(K3\) surfaces.