Skip to main content
Log in

On the cohomology of moduli spaces of (weighted) stable rational curves

  • Published:
Mathematische Zeitschrift Aims and scope Submit manuscript

Abstract

We give a recursive algorithm for computing the character of the cohomology of the moduli space \({\overline{M}}_{0,n}\) of stable \(n\)-pointed genus zero curves as a representation of the symmetric group \(\mathbb{S }_n\) on \(n\) letters. Using the algorithm we can show a formula for the maximum length of this character. Our main tool is connected to the moduli spaces of weighted stable curves introduced by Hassett.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Similar content being viewed by others

Notes

  1. More precisely, when \(n\) is even, we have to consider the resolution of \((\mathbb P ^1)^n/\!\!/ \mathrm{PGL}(2)\) constructed by Kirwan [8], see Lemma 4.4 below.

References

  1. Ceyhan, Ö.: Chow groups of the moduli spaces of weighted pointed stable curves of genus zero. Adv. Math. 221(6), 1964–1978 (2009)

    Article  MathSciNet  MATH  Google Scholar 

  2. Faber, C., Pandharipande, R.: Tautological and non-tautological cohomology of the moduli space of curves. In: Farkas, G., Morrison, I. (eds.) Handbook of Moduli, Advanced Lectures in Mathematics, vol. 1, pp. 293–330. International Press, Boston (2012)

  3. Getzler, E.: Operads and moduli spaces of genus 0 Riemann surfaces. In: The moduli space of curves (Texel Island, 1994), pp. 199–230, Progr. Math., 129. Birkhäuser Boston, Boston, MA (1995)

  4. Getzler, E., Kapranov, M.M.: Modular operads. Compositio Math. 110(1), 65–126 (1998)

    Article  MathSciNet  MATH  Google Scholar 

  5. Hassett, B.: Moduli spaces of weighted pointed stable curves. Adv. Math. 173(2), 316–352 (2003)

    Article  MathSciNet  MATH  Google Scholar 

  6. Keel, S.: Intersection theory of moduli space of stable \(n\)-pointed curves of genus zero. Trans. Amer. Math. Soc. 330(2), 545–574 (1992)

    MathSciNet  MATH  Google Scholar 

  7. Kiem, Y.-H., Moon, H.-B.: Moduli spaces of weighted pointed stable rational curves via GIT. Osaka J. Math. 48(4), 1115–1140 (2011)

    MathSciNet  MATH  Google Scholar 

  8. Kirwan, F.: Partial desingularisations of quotients of nonsingular varieties and their Betti numbers. Ann. Math. (2) 122(1), 41–85 (1985)

    Article  MathSciNet  MATH  Google Scholar 

  9. Macdonald, I.G.: Symmetric Functions and Hall Polynomials, 2nd edn. The Clarendon Press, Oxford University Press, New York (1995)

    MATH  Google Scholar 

  10. Newell, M.J.: A theorem on the plethysm of \(S\)-functions. Quart. J. Math. Oxford Ser. (2) 2, 161–166 (1951)

    Article  MathSciNet  MATH  Google Scholar 

  11. Thaddeus, M.: Geometric invariant theory and flips. J. Am. Math. Soc. 9(3), 691–723 (1996)

    Article  MathSciNet  MATH  Google Scholar 

  12. Voisin, C.: Hodge Theory and Complex Algebraic Geometry I, Cambridge Studies in Advanced Mathematics, 76. Cambridge University Press, Cambridge (2002)

    Book  Google Scholar 

Download references

Acknowledgments

The second named author is supported in part by JSPS Grant-in-Aid for Young Scientists (No. 22840041). The authors thank the Max–Planck–Institut für Mathematik for hospitality. The authors also thank the referees for their thorough reading and helpful comments.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Jonas Bergström.

Appendices

Appendix A: Cohomology of the blow-up

First, we recall the fact that the cohomology of the blow-up \(\widetilde{M}\) of a smooth projective variety \(M\) along a smooth subvariety \(Z\) of codimension \(l\) is given by

$$\begin{aligned} H^k(\widetilde{M})\cong H^k(M)\oplus \bigoplus _{i=1}^{l-1} \left( H^{k-2i}(Z)\otimes H^{2i}(\mathbb P ^{l-1})\right) ~, \end{aligned}$$
(6.1)

see e.g. [12, Theorem 7.31].

Let \(X\) be a smooth projective variety and \(Y=Y_1 \cup \cdots \cup Y_n\) be the union of smooth subvarieties of \(X\). Let \(\widetilde{X}\) be the blow-up of \(X\) along \(Y\). We assume that any non-empty intersection of irreducible components of \(Y\) is transversal. Then it is known that \(\widetilde{X}\) is obtained by a sequence of smooth blow-ups along the proper transforms of the irreducible components of \(Y\) in any order, see e.g. [7, Proposition 2.10]. We further assume that \(\mathrm{codim}_X Y_i=l\) for any \(i\). The aim of this appendix is to give a formula for \(H^*(\widetilde{X})\). For a subset \(I\) of \(\{1, \ldots , n\}\), we set \(Y_I:=\cap _{i\in I} Y_i\), which is a smooth subvariety of \(X\) by the assumption of transversality.

Proposition 6.1

Under the above assumptions, we have

$$\begin{aligned} H^*(\widetilde{X}) \cong H^*(X) \oplus \bigoplus _{I \subset \{1,\ldots ,n\}} \left( H^*(Y_I) \otimes \left( H^+(\mathbb P ^{l-1}) \right) ^{\otimes |I|}\right) ~, \end{aligned}$$
(6.2)

where \(I\) runs over all subsets of \(\{1,\ldots , n\}\) for which \(Y_I \ne \emptyset \).

Proof

Let

$$\begin{aligned} X_{n} \overset{\pi _n}{\longrightarrow } X_{n-1} \rightarrow \cdots \rightarrow X_1 \overset{\pi _1}{\longrightarrow } X_0 \end{aligned}$$

be the sequence of blow-ups which we define inductively as follows.

  1. (i)

    Let \(\pi _1: X_1\rightarrow X_0:=X\) be the blow-up along \(Y_1\).

  2. (ii)

    For \(i\ge 2\), we put \(\pi _{i}: X_{i} \rightarrow X_{i-1}\) to be the blow-up along the proper transform of \(Y_i\).

Note that \(X_n\) is isomorphic to \(\widetilde{X}\). For \(1\le j \le i-1\), we denote by \(Y_{i,j}\) the proper transform of \(Y_i\) under \(\pi _{j}\circ \cdots \circ \pi _1: X_{j} \rightarrow X_{0}\). In this notation, \(\pi _i: X_i\rightarrow X_{i-1}\) is the blow-up along \({Y}_{i, i-1}\). Then it follows from (6.1) that

$$\begin{aligned} H^*(X_i) \cong H^*(X_{i-1}) \oplus \left( H^*({Y}_{i,i-1}) \otimes H^+(\mathbb P ^{l-1}) \right) , \end{aligned}$$
(6.3)

as graded vector spaces. The Eq. (6.3) together with (6.4) given in Lemma 6.2 below implies (6.2).\(\square \)

Lemma 6.2

Under the same assumptions and notation as above, we have

$$\begin{aligned} H^*({Y}_{i,i-1})\cong H^*(Y_i) \oplus \bigoplus _{I\subset \{1,\ldots ,i-1\}} \left( H^*(Y_i \cap Y_I)\otimes \left( H^+ (\mathbb P ^{l-1}) \right) ^{\otimes |I|} \right) , \end{aligned}$$
(6.4)

where \(I\) runs over all the subsets of \(\{1,\ldots , i-1\}\) for which \(Y_i\cap Y_I \ne \emptyset \).

Proof

Note that the proper transform \(Y_{i,i-1}\) of \(Y_i\) is obtained by the sequence of blow-ups

$$\begin{aligned} Y_{i,i-1}\overset{\pi _{i-1}}{\longrightarrow } Y_{i,i-2}\rightarrow \cdots \rightarrow Y_{i,1} \overset{\pi _1}{\longrightarrow } Y_{i,0}:=Y_i~, \end{aligned}$$

and that the center of the blow-up \(\pi _j: Y_{i,j}\rightarrow Y_{i,j-1}\) is \(Y_{i,j-1}\cap Y_{j,j-1}\). In general, for \(I\subset \{1,\ldots , n\}\) and \(j<\min \{i\in I\}\) such that \(Y_{I,j}:=\cap _{i\in I}Y_{i,j}\ne \emptyset , Y_{I,j}\) is the blow-up of \(Y_{I,j-1}\) along \(Y_{I,j-1}\cap Y_{j,j-1}\). Using this structure and (6.1) recursively, we obtain

$$\begin{aligned} H^*({Y}_{i,j})\cong H^*(Y_i) \oplus \bigoplus _{I\subset \{1,\ldots ,j\}} \left( H^*(Y_i \cap Y_I)\otimes \left( H^+(\mathbb P ^{l-1}) \right) ^{\otimes |I|} \right) , \end{aligned}$$
(6.5)

where \(I\) runs over all the subsets of \(\{1,\ldots , j\}\) for which \(Y_I\cap Y_i \ne \emptyset \). The desired formula (6.4) is the case \(j=i-1\) in (6.5).\(\square \)

Appendix B: Characters of representations of symmetric groups

Let \(\mathbb{S }_n\) be the symmetric group on \(n\) letters. Let \(\Lambda :=\underset{\longleftarrow }{\lim }~\mathbb Z [x_1, \ldots , x_n]^{\mathbb{S }_n}\) be the ring of symmetric functions. It is well known that \(\Lambda \otimes \mathbb Q =\mathbb Q [p_1, p_2, \ldots ]\) where \(p_n\) are the power sums. We denote by \(\mathcal{P }(n)\) the set of partitions of \(n\), and for \(\lambda =(\lambda _1,\ldots ,\lambda _{l(\lambda )}) \in \mathcal{P }(n)\) we set \(p_{\lambda }:=\prod \nolimits _{i=1}^{l(\lambda )} p_{\lambda _i}\). We also set \(\Lambda ^{x, y}:=\Lambda ^x\otimes \Lambda ^y\), where \(\Lambda ^x\) and \(\Lambda ^y\) are the ring of symmetric functions in \(x=(x_1, x_2, \ldots )\) and \(y=(y_1, y_2, \ldots )\), respectively.

For a representation \(V\) of \(\mathbb{S }_n\), we define

$$\begin{aligned} \mathrm{ch}_n(V) :=\frac{1}{n!} \sum _{w \in \mathbb{S }_n} \mathrm{Tr}_V(w) p_{\rho (w)} \in \Lambda , \end{aligned}$$

where \(\rho (w)\in \mathcal{P }(n)\) is the cycle type of \(w\in \mathbb{S }_n\). Similarly we define, for an \(\mathbb{S }_k \times \mathbb{S }_{n-k}\) representation \(V\),

$$\begin{aligned} \mathrm{ch}_{k, n-k}(V) :=\frac{1}{k! }\frac{1}{(n-k)!} \sum _{(u,v) \in \mathbb{S }_k \times \mathbb{S }_{n-k}} \mathrm{Tr}_V\left( (u,v)\right) p_{\rho (u)}^x \, p_{\rho (v)}^y \in \Lambda ^{x, y}, \end{aligned}$$

where \(p_n^x\) and \(p_n^y\) are the power sums in the variable \(x\) and \(y\), respectively.

If \(V\) and \(W\) are representations of \(\mathbb{S }_n\) we put

$$\begin{aligned} \mathrm{ch}_{n}(V) * \mathrm{ch}_{n}(W):=\mathrm{ch}_{n}(V \otimes W). \end{aligned}$$

For any \(\lambda \in \mathcal{P }(k)\), if we put \(m_j(\lambda ):=\#\{ i \mid \lambda _i =j\}\) and

$$\begin{aligned} \frac{\partial }{\partial p^x_{\lambda }}:=\left( \prod _{i=1}^{k}\frac{1}{m_i(\lambda )!} \right) \frac{\partial }{\partial p^x_{\lambda _1}} \frac{\partial }{\partial p^x_{\lambda _2}} \cdots \frac{\partial }{\partial p^x_{\lambda _{l(\lambda )}}}, \end{aligned}$$

then

$$\begin{aligned} \mathrm{ch}_{k,n-k}\left( \mathrm{Res}^{\mathbb{S }_{n}}_{\mathbb{S }_{k} \times \mathbb{S }_{n-k}}(V)\right) =\sum _{\lambda \in \mathcal{P }(n-k)} \left( \frac{\partial }{\partial p^x_{\lambda }} \mathrm{ch}_{n}(V) \right) \, p^y_{\lambda }. \end{aligned}$$

If \(V_i\) are representations of \(\mathbb{S }_{n_i}\) for \(1 \le i \le k\) then

$$\begin{aligned} \mathrm{ch}_{\sum \nolimits _{i=1}^k n_i} \left( \mathrm{Ind}_{\mathbb{S }_{n_1} \times \cdots \times \mathbb{S }_{n_k}}^{\mathbb{S }_{\sum \nolimits _{i=1}^k n_i}} (V_1 \boxtimes V_2 \boxtimes \cdots \boxtimes V_k) \right) = \prod _{i=1}^k \mathrm{ch}_{n_i}(V_i). \end{aligned}$$

If \(\circ \) denotes plethysm between symmetric functions, and \(\sim \) denotes the wreath product, that is, \(\mathbb{S }_{n_1} \sim \, \mathbb{S }_{n_2}:=\mathbb{S }_{n_1} \ltimes (\mathbb{S }_{n_2})^{n_1}\) where \(\mathbb{S }_{n_1}\) acts on \((\mathbb{S }_{n_2})^{n_1}\) by permutation, then

$$\begin{aligned} \mathrm{ch}_{n_1 n_2}\left( \mathrm{Ind}_{\mathbb{S }_{n_1} \sim \, \mathbb{S }_{n_2}}^{\mathbb{S }_{n_1 n_2}} (V_1 \boxtimes \underbrace{ V_2 \boxtimes \cdots \boxtimes V_2 }_{n_1} \,) \right) = \mathrm{ch}_{n_1}(V_1) \circ \mathrm{ch}_{n_2}(V_2), \end{aligned}$$

see [9, Appendix A, p. 158].

Recall finally that irreducible representations of \(\mathbb{S }_n\) are indexed by \(\mathcal{P }(n)\). For \(\lambda \in \mathcal{P }(n)\), let \(V_{\lambda }\) be the irreducible representation corresponding to \(\lambda \) and define the Schur function

$$\begin{aligned} s_{\lambda }:=\mathrm{ch}_n(V_{\lambda })\in \Lambda . \end{aligned}$$

It is well-known that \(\{s_{\lambda }\}\), where \(\lambda \) runs over all the partitions, is a \(\mathbb Z \)-basis of \(\Lambda \).

Appendix C: An ordering of symmetric functions

For any partition \(\lambda =(\lambda _1,\lambda _2,\ldots ,\lambda _{l(\lambda )})\) we denote its dual partition by \(\lambda ^{\prime }=(\lambda _1^{\prime },\lambda _2^{\prime },\ldots )\), where \(\lambda _i^{\prime }:= |\{j:\lambda _j \ge i \}|\). If \(\mu =(\mu _1,\mu _2,\ldots ,\mu _{l(\mu )})\) is another partition we define the partitions \(\lambda +\mu \) as \((\lambda _1+\mu _1,\lambda _2+\mu _2,\ldots )\) and the partition \(\lambda \cup \mu \) as the reordering of \((\lambda _1,\ldots ,\lambda _{l(\lambda )},\mu _1,\mu _2,\ldots ,\mu _{l(\mu )})\). Note that \(\lambda \cup \mu =(\lambda ^{\prime }+\mu ^{\prime })^{\prime }\).

We introduce an ordering. For any partitions \(\lambda \) and \(\mu \) we say that \(\lambda > \mu \) if there is a \(k\) such that \(\lambda _i^{\prime }=\mu _i^{\prime }\) for \(1 \le i \le k-1\) and \(\lambda _k^{\prime }> \mu _k^{\prime }\).

Definition 8.1

For any symmetric function \(f=\sum \nolimits _{\lambda } a_{\lambda } s_{\lambda }\) we let \(w(f)\) be the maximal partition \(\lambda \) (w.r.t. \(>\)) such that \(a_\lambda \ne 0\).

For any partition \(\lambda \), we put \(h_{\lambda }:=\prod \nolimits _{i=1}^{l(\lambda )} s_{(\lambda _i)}\) and \(e_{\lambda }:=\prod \nolimits _{i=1}^{l(\lambda )} s_{(1^{\lambda _i})}\). The following is well known.

Lemma 8.2

There are integers \(a_{\lambda ,\mu }\) and \(b_{\lambda ,\mu }\) such that

$$\begin{aligned} s_{\lambda }=h_{\lambda }+\sum _{\mu < \lambda } a_{\lambda ,\mu }h_{\mu }=e_{\lambda ^{\prime }}+\sum _{\mu < \lambda ^{\prime }} b_{\lambda ,\mu }e_{\mu }. \end{aligned}$$

Lemma 8.3

For any symmetric functions \(f\) and \(g\) we have, \(w(fg)=w(f) \cup w(g).\)

Proof

Since \(h_{\mu } h_{\nu }=h_{\mu \cup \nu }\), it follows from Lemma 8.2 that there are integers \(c_{\lambda }\) such that \(fg=c_{w(f) \cup w(g)}h_{w(f) \cup w(g)}+\sum \nolimits _{\lambda < w(f) \cup w(g)} c_{\lambda } h_{\lambda }\) with \(c_{w(f) \cup w(g)}\ne 0\).\(\square \)

Due to the lack of a suitable reference we will show here how \(w\) behaves with respect to plethysm between symmetric functions.

Lemma 8.4

[9, p. 158] For any symmetric functions \(f\), \(g\) and \(h\) we have,

$$\begin{aligned} (f \circ h) (g \circ h)=(f \, g) \circ h. \end{aligned}$$

We denote by \((\cdot \,| \cdot )\) the standard inner product on \(\Lambda \) for which Schur functions are orthonormal.

Proposition 8.5

[10, Theorem I, IA] Say that \(s_{(k)} \circ s_{(m)}=\sum \nolimits _{\lambda } a_{\lambda } s_{\lambda }\), then, for any \(0 \le i \le k\),

$$\begin{aligned} (s_{(1^i)} \circ s_{(m-1)}) \, (s_{(k-i)} \circ s_{(m)}) = \sum _{\lambda } \sum _{\nu } a_{\lambda } \left( s_{(1^i)}s_{\nu } | s_{\lambda }\right) s_{\nu }. \end{aligned}$$
(8.1)

Similarly, say that \(s_{(1^k)} \circ s_{(m)}=\sum \nolimits _{\lambda } b_{\lambda } s_{\lambda }\), then, for any \(0 \le i \le k\),

$$\begin{aligned} (s_{(i)} \circ s_{(m-1)}) \, (s_{(1^{k-i})} \circ s_{(m)}) = \sum _{\lambda } \sum _{\nu } b_{\lambda } \left( s_{(1^i)}s_{\nu } | s_{\lambda }\right) s_{\nu }. \end{aligned}$$
(8.2)

Proposition 8.6

For any \(m \ge 1\) and partition \(\mu \) of \(k\) we have

$$\begin{aligned} w(s_{\mu } \circ s_{(m)})= {\left\{ \begin{array}{ll} ({(m-1)}^k) + \mu \quad \,\text{ if }\quad m\, \text{ odd } \\ ({(m-1)}^k) + \mu ^{\prime } \quad \,\text{ if }\quad m\,\text{ even }. \end{array}\right. } \end{aligned}$$
(8.3)

Proof

We first recall that \(l(s_{\mu } \circ s_{(m)}) \le |\mu |\), see [9, Example 9, p. 140]. Let us then continue by proving Eq. (8.3) for \(\mu =(k)\) and \(\mu =(1^k)\) by induction on \(m\). The equation clearly holds for \(m=1\). If \(l(s_{\nu })=k\) and \(l(s_{\lambda }) \le k\) then \((s_{(1^k)}s_{\nu }|s_{\lambda }) \ne 0\) implies that \(\lambda =\nu + (1^k)\). Moreover, if \(l(s_{\eta })<k\), \(l(s_{\lambda }) \le k\) and \((s_{(1^k)}s_{\eta }|s_{\lambda }) \ne 0\) then \(\lambda < \nu + (1^k)\) for any \(\nu \) such that \(l(\nu )=k\). Therefore, taking \(i=k\) in Eq. (8.1) and applying the fact that \(l(s_{(k)} \circ s_{(m)}) \le k\) we find by looking at the term with \(\lambda =w(s_{(k)}\circ s_{(m)})\) on the right hand side of formula (8.1) that \(w(s_{(k)} \circ s_{(m)})=w(s_{(1^k)} \circ s_{(m-1)}) + (1^k)\). Similarly, using Eq. (8.2) we find by induction on \(m\) that \(w(s_{(1^k)} \circ s_{(m)})=w(s_{(k)} \circ s_{(m-1)}) + (1^k)\). This completes the induction.

Finally, let us prove the statement for any \(\mu \) and \(m\). From Lemma 8.4 and from the formula for \(w(s_{(k)} \circ s_{(m)})\) it follows, for \(m\) odd, that

$$\begin{aligned} w(h_{\mu } \circ s_{(m)})= \bigcup _i w(s_{(\mu _i)} \circ s_{(m)})={((m-1)}^k) + \mu \end{aligned}$$

and similarly from the formula for \(w(s_{(1^k)} \circ s_{(m)})\), for \(m\) even, that

$$\begin{aligned} w(e_{\mu } \circ s_{(m)})= \bigcup _i w(s_{(1^{\mu _i})} \circ s_{(m)})={((m-1)}^k) + \mu . \end{aligned}$$

Using Lemma 8.2 it then follows, for \(m\) odd, that

$$\begin{aligned} w(s_{\mu } \circ s_{(m)})=w \left( (h_{\mu }+\sum _{\nu < \mu } a_{\nu }h_{\nu }) \circ s_{(m)} \right) =w(h_{\mu } \circ s_{(m)}) \end{aligned}$$

and, for \(m\) even, that

$$\begin{aligned} w(s_{\mu } \circ s_{(m)})=w \left( (e_{\mu ^{\prime }}+\sum _{\nu < \mu ^{\prime }} b_{\nu } e_{\nu }) \circ s_{(m)} \right) =w(e_{\mu ^{\prime }} \circ s_{(m)}). \end{aligned}$$

\(\square \)

Rights and permissions

Reprints and permissions

About this article

Cite this article

Bergström, J., Minabe, S. On the cohomology of moduli spaces of (weighted) stable rational curves. Math. Z. 275, 1095–1108 (2013). https://doi.org/10.1007/s00209-013-1171-8

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00209-013-1171-8

Mathematics Subject Classification (2000)

Navigation