math0104187/mrc.tex
1: 
2: 
3: 
4: \documentclass[12pt]{amsart}
5: 			 
6: \usepackage{amsfonts}
7: \usepackage{amssymb}
8: \usepackage{amscd, graphics}
9: \usepackage{epsfig}
10: \input xy
11: \xyoption{all}
12: 
13: \setlength{\parindent}{.4 in}
14: \setlength{\textwidth}{5.8 in}
15: \setlength{\topmargin} {-.3 in}
16: \setlength{\evensidemargin}{0 in}
17: \setlength{\oddsidemargin}{0 in}
18: \setlength{\footskip}{.3 in}
19: \setlength{\headheight}{.3 in}
20: \setlength{\textheight}{8.5 in}
21: \setlength{\parskip}{.1 in}
22: 
23: \newcommand{\marginlabel}[1]%
24:   {\mbox{}\marginpar{\raggedleft\hspace{0pt}\bfseries\sf#1}}
25: 
26: 
27: %%----------------------------------------------------------
28: %%  SHORTCUTS
29: %%----------------------------------------------------------
30: \def\ZZ{{\mathbb Z}}
31: \def\NN{{\mathbb N}}
32: \def\FF{{\mathbb F}}
33: \def\CC{{\mathbb C}}
34: \def\AA{{\mathbb A}}
35: \def\RR{{\mathbb R}}
36: \def\QQ{{\mathbb Q}}
37: \def\PP{{\textbf P}}
38: \def\OO{{\mathcal O}}
39: 
40: 
41: \def\cD{\mathcal{D}}
42: \def\cA{\mathcal{A}}
43: \def\cF{\mathcal{F}}
44: \def\cL{\mathcal{L}}
45: \def\cO{\mathcal{O}}
46: \def\cI{\mathcal{I}}
47: \def\I{\mathcal{I}}
48: \def\cM{\mathcal{M}}
49: \def\cZ{\mathcal{Z}}
50: \def\cU{\mathcal{U}}
51: \def\cQ{\mathcal{Q}}
52: \def\cE{\mathcal{E}}
53: 
54: \def\Mg{\cM_{g}}
55: \def\Mgn{\cM_{g,n}}
56: \def\Mg1{\cM_{g,g+1}}
57: \def\Pici{{\rm Pic}^{g-2i-1}(C)}
58: \def\hsigma{\widehat{\sigma}}
59: \def\iso{\simeq}
60: \def\reg{\rm reg}
61: \def\tensor{\otimes}
62: 
63: %%----------------------------------------------------------
64: %%  NEW OPERATORS
65: %%----------------------------------------------------------
66: \DeclareMathOperator{\Coker}{Coker}
67: \DeclareMathOperator{\Ker}{Ker}
68: \DeclareMathOperator{\depth}{depth}
69: \DeclareMathOperator{\Pic}{Pic}
70: \DeclareMathOperator{\Hom}{Hom}
71: \DeclareMathOperator{\HOM}{\mathcal{Hom}}
72: \DeclareMathOperator{\rank}{rank}
73: \DeclareMathOperator{\Spec}{Spec}
74: \DeclareMathOperator{\Cl}{Cl}
75: 
76: 
77: %%----------------------------------------------------------
78: %%  OTHER ENVIRONMENTS    
79: %%----------------------------------------------------------
80: 
81: %%\swapnumbers
82: 
83: \theoremstyle{plain}
84: \newtheorem*{introtheorem}{Theorem}   
85: \newtheorem*{introcorollary}{Corollary}
86: \newtheorem{theorem}{Theorem}[section]
87: \newtheorem{proposition}[theorem]{Proposition}
88: \newtheorem{corollary}[theorem]{Corollary}
89: \newtheorem{lemma}[theorem]{Lemma}
90: 
91: \theoremstyle{definition}
92: \newtheorem{definition}[theorem]{Definition}
93: \newtheorem{remark}[theorem]{Remark}
94: \newtheorem{example}[theorem]{Example}
95: \newtheorem{sketch}[theorem]{Sketch}
96: \newtheorem{conjecture}[theorem]{Conjecture}
97: \newtheorem{conjecture/question}[theorem]{Conjecture/Question}
98: \newtheorem{question}[theorem]{Question}  
99: 
100: \pagestyle{myheadings}
101: \theoremstyle{remark}
102: \newtheorem*{nctheorem}{Theorem}
103: 
104: 
105: %%----------------------------------------------------------
106: %%  TITLE PAGE INFO  
107: %%----------------------------------------------------------
108: \begin{document}
109: 
110: \title{Divisors on $\Mg1$ and the Minimal Resolution Conjecture
111: for points on canonical curves}
112: 
113: \author[G. Farkas]{Gavril Farkas}
114: \address{Department of Mathematics, University of Michigan, 
115: 525 East University, Ann Arbor, MI, 48109-1109}
116: \email{{\tt gfarkas@umich.edu}}
117: 
118: \author[M. Musta\c{t}\v{a}]{Mircea~Musta\c{t}\v{a}}
119: \address{Department of Mathematics, University of California,
120: Berkeley, CA, 94720 and Institute of Mathematics of
121: the Romanian Academy}
122: \email{{\tt mustata@math.berkeley.edu}}
123: 
124: \author[M. Popa]{Mihnea Popa} 
125: \address{Department of Mathematics, Harvard University, 
126: One Oxford Street, Cambridge, MA 02138}
127: \email{{\tt mpopa@math.harvard.edu}}
128: 
129: \subjclass{Primary 14H10; Secondary 13D02, 14F05.}
130: \keywords{Minimal free resolutions, difference varieties, moduli spaces of curves with 
131: marked points}
132: \maketitle
133: \markboth{G. FARKAS, M. MUSTA\c T\v A and M. POPA}
134: {\bf THE MINIMAL RESOLUTION CONJECTURE}
135: 
136: %%----------------------------------------------------------
137: %%  INTRODUCTION    
138: %%----------------------------------------------------------
139: \section*{\bf Introduction}
140: 
141: The Minimal Resolution Conjecture for points in projective space has 
142: attracted considerable attention in recent years, starting with the 
143: original \cite{lorenzini1}, \cite{lorenzini2} and continuing most notably 
144: with \cite{gaeta}, \cite{BG}, \cite{walter},
145: \cite{simpson}, \cite{popescu}, \cite{epsw}. The purpose
146: of this paper 
147: is to explain how a completely analogous problem can be formulated for 
148: sets of points on arbitrary varieties embedded in projective space, and 
149: then study in detail the case of curves. Similarly to 
150: the well-known analysis of syzygies of curves carried out by Green and 
151: Lazarsfeld (\cite{gl1}, \cite{gl2}, \cite{gl3}), we divide our work into
152: a study of resolutions of points on canonical curves and on curves of 
153: large degree. The central result of the paper states that the Minimal 
154: Resolution Conjecture is true on any canonical curve. In contrast, it 
155: always fails for curves embedded with large degree, although a weaker 
156: result, called the Ideal Generation Conjecture, holds also in this case.
157: These results turn out to have surprisingly deep connections with the 
158: geometry of difference varieties in Jacobians, special divisors on 
159: moduli spaces of curves with marked points, and moduli spaces of stable
160: bundles.  
161: 
162: Let $X$ be a projective variety over an algebraically closed field,
163: embedded by a (not necessarily complete) linear series. 
164: We begin by formulating a general version of the 
165: Minimal Resolution Conjecture (MRC), in analogy with the case of $\PP^n$,
166: predicting how the Betti numbers of 
167: a general subset of points of $X$ in the given embedding are related to the 
168: Betti numbers of $X$ itself. More precisely (cf. Theorem 
169: \ref{general_results} below),  
170: for a large enough general set of points $\Gamma$ on $X$, the Betti diagram 
171: consisting of the graded Betti numbers $b_{i,j}(\Gamma)$ is obtained from 
172: the Betti diagram of $X$ by adding two more nontrivial rows, at places 
173: well determined by the length of $\Gamma$. Recalling that the Betti diagram 
174: has the Betti number $b_{i,j}$ in the $(j,i)$-th position, and assuming 
175: that the two extra rows are indexed by $i=r-1$ and $i=r$, for some integer 
176: $r$, the MRC predicts that 
177: $$b_{i+1,r-1}(\Gamma)\cdot b_{i,r}(\Gamma)= 0,$$
178: i.e. at least one of the two Betti numbers on any "diagonal" is zero. 
179: As the difference $b_{i+1, r-1} - b_{i,r}$ can be computed exactly,
180: this implies a precise knowledge of the Betti numbers in these two rows.
181: Summing up, knowing the Betti diagram of $\Gamma$ would be the same as knowing 
182: the Betti diagram of $X$. A subtle question is however to understand
183: how the shape of the Betti diagram of $X$ influences whether MRC
184: is satisfied for points on $X$. An example illustrating this is given 
185: at the end of Section 1.
186: 
187: The Minimal Resolution Conjecture
188:  has been extensively studied in the case $X=\PP^n$. The
189: conjecture holds for $n\leq 4$ by results of
190: Gaeta, Ballico and Geramita, and Walter (see \cite{gaeta},
191: \cite{BG} and
192: \cite{walter}, respectively).
193: Moreover, Hirschowitz and Simpson 
194: proved in \cite{simpson}
195:  that it holds if the number of points is large enough with respect
196: to $n$. However, the conjecture does not hold in general:
197: it fails for every $n\geq 6$, $n\neq 9$ for almost $\sqrt{n}/2$
198: values of the number of points, by a result of Eisenbud, Popescu, Schreyer
199: and Walter (see \cite{epsw}). We refer to \cite{popescu} and
200: \cite{epsw} for a nice introduction and an account of the present status
201: of the problem in this case.
202: 
203: 
204: The main body of the paper is dedicated to a detailed study of MRC
205: in the case of curves. 
206: We will simply say that a curve \emph{satisfies MRC} in a given embedding 
207: if MRC is satisfied by a general set of points $\Gamma$ of any 
208: sufficiently large degree
209: (for the precise numerical statements see Section 1). We will also 
210: sometimes say that MRC holds for a line bundle $L$ if it 
211: holds for $C$ in the embedding given by $L$.  
212: Our main result says that MRC holds in the most significant case,
213: namely the case of canonically embedded nonhyperelliptic curves.
214: 
215: \begin{introtheorem}
216: If $C$ is a canonical curve, then $C$ satisfies MRC.
217: \end{introtheorem}
218: 
219: In contrast, under very mild assumptions on the genus, 
220: the MRC always fails in the case of curves of large degree, at well-determined 
221: spots in the Betti diagram (cf. Section 2 for precise details). The statement
222: $b_{2,r-1}\cdot b_{1,r}=0$, i.e. the case $i=1$, does hold though; this is 
223: precisely the Ideal Generation Conjecture, saying that the 
224: minimal number of generators of $I_\Gamma / I_X$ is as small as possible.
225: 
226: \begin{introtheorem}
227: (a) If $L$ is a very ample line bundle of degree $d\geq 2g$, then $IGC$ 
228: holds for $L$.
229: \newline
230: (b) If $g\geq 4$ and $L$ is a line bundle of degree $d\geq 2g+16$,
231: then there exists 
232: a value of $\gamma$ such that $C\subseteq\PP H^{0}(L)$ does not satisfy MRC 
233: for $i=\lfloor\frac{g+1}{2}\rfloor$. The same holds if $g\geq 15$ and
234:  $d\geq 2g+5$.
235: \end{introtheorem}
236: 
237: It is interesting to note that by the "periodicity" property of Betti
238: diagrams of 
239: general points on curves (see \cite{mustata} \S2), the theorem above
240: implies that 
241: on curves of high degree, MRC fails for sets of points of
242: arbitrarily large length.
243: This provides a very different picture from the case of projective space 
244: (cf. \cite{simpson}),
245: where asymptotically the situation is as nice as possible.
246: 
247: We explain the strategy involved in the proof of these results
248:  in some detail, as 
249: it appeals to some new geometric techniques in the study of
250:  syzygy related questions.
251: For simplicity we assume here that $C$ is a smooth curve embedded
252:  in projective space 
253: by means of a complete linear series corresponding to a very ample
254:  line bundle $L$
255: (but see \S2 for more general statements).
256:  A well-known geometric approach, developed  
257: by Green and Lazarsfeld in the study of syzygies of curves
258:  (see \cite{lazarsfeld} for a survey),
259: is to find vector bundle statements equivalent to the algebraic ones,
260:  via Koszul cohomology.
261: This program can be carried out completely in the case of MRC,
262:  and for curves we get a 
263: particularly clean statement. Assume that $M_L$ is the kernel of
264:  the evaluation map
265: $H^0(L)\otimes \OO_C\rightarrow L$ and $Q_L:=M_L^{*}$.
266: Then (cf. Corollary \ref{criterion} below) MRC holds 
267: for a collection of $\gamma\geq g$ general points on $C$
268:  if and only if the following 
269: is true:
270: $$h^0(\wedge^i M_L \otimes \xi)=0, ~{\rm for~all}~ i  ~{\rm and}~
271: \xi\in
272:  {\rm Pic}^{g-1+\lfloor{\frac{di}{n}}\rfloor}(C)~{\rm general}.\,\,\,\,(*)$$
273: Condition $(*)$ above is essentially the condition studied by Raynaud  
274: \cite{raynaud}, related to the existence of theta divisors for semistable 
275: vector bundles. In the particular situation of $\wedge^i M_L$, 
276: with $L$ a line bundle of large degree, it has been 
277: considered in \cite{popa} in order to produce base points for the determinant 
278: linear series on the moduli spaces $SU_C(r)$ of semistable bundles of rank $r$ 
279: and trivial determinant. A similar approach shows here the failure of
280: condition $(*)$ (and so of MRC) for $i=[\frac{g+1}{2}]$. On the
281: other hand, the fact that 
282: IGC holds is a
283:  rather elementary application of the Base Point Free Pencil Trick
284: \cite{ACGH} III \S3.
285: 
286: The case of canonical curves is substantially more involved, but in the end 
287: one is rewarded with a positive answer. As above, it turns out that MRC is 
288: equivalent to the vanishing:
289: $$h^0(\wedge^i Q\otimes \xi)=0, ~{\rm for~all}~i ~{\rm and~}~\xi\in
290:  {\rm Pic}^{g-2i-1}(C) 
291: ~{\rm general},$$
292: where $Q$ is the dual of the bundle $M$ defined by the evaluation sequence: 
293: $$0\longrightarrow M\longrightarrow H^0(\omega_C)\otimes \OO_C\longrightarrow 
294: \omega_C\longrightarrow 0.$$
295: As the slope of $\wedge^i Q$ is $2i\in \ZZ$, this is in turn equivalent to the 
296: fact that $\wedge^i Q$ has
297:  a theta divisor $\Theta_{\wedge^i Q}\in {\rm Pic}^{g-2i-1}(C)$. 
298: On a fixed curve, if indeed a divisor,
299: $\Theta_{\wedge^i Q}$ will be identified as being precisely 
300: the difference variety $C_{g-i-1}-C_i \subseteq\Pici$ (cf. \cite{ACGH} Ch.V.D), where 
301: $C_n$ is the $n$-th symmetric product of $C$.
302: This is achieved via a 
303: filtration argument and a cohomology class calculation similar to the 
304: classical Poincar\'e theorem
305:  (cf. Proposition \ref{1}). A priori though, on an arbitrary curve 
306: the nonvanishing locus
307:  $\{\xi~|~h^0(\wedge^i Q\otimes \xi)\neq 0\}$ may be the whole $\Pici$, 
308: in which case this identification 
309: is meaningless. We overcome this problem by working with all curves at once, that 
310: is by setting up a similar universal construction on the moduli space 
311: of curves with marked points $\Mg1$.
312:  Here we slightly oversimplify the exposition in order 
313: to present the main idea,
314:  but for the precise technical details see Section 3. 
315: We essentially consider the "universal nonvanishing locus" in $\Mg1$: 
316: $$\cZ = \{(C,x_1,\ldots ,x_{g-i},y_1,\ldots ,y_{i+1})|h^0(\wedge^i
317: Q_C\otimes\OO(x_1+\ldots 
318: +x_{g-i}-y_1-\ldots -y_{i+1}))\neq 0\}.$$
319: The underlying idea is that the difference line bundles $\OO_C(x_1+\ldots 
320: +x_{g-i}-y_1-\ldots -y_{i+1})$ in fact cover the whole $\Pici$ (i.e. 
321: $C_{g-i}-C_{i+1}= \Pici$), and so 
322: for any given curve $C$,
323:  $\cZ|_C$ is precisely the nonvanishing locus described above. 
324: The advantage of writing
325: it in this form is that we are led to performing a computation on $\Mg1$ rather
326: than on a universal Picard,
327:  where for example one does not have a canonical choice of 
328: generators for the Picard group.  
329: A ``deformation to hyperelliptic
330: curves'' argument easily implies that MRC holds for general canonical 
331: curves, so $\cZ$ is certainly a divisor.
332: We then show that $\cZ$ is the degeneracy locus of a morphism of vector
333: bundles of the 
334: same rank and compute its class using a Grothendieck-Riemann-Roch 
335: argument (cf. Proposition \ref{degeneracy_class}). 
336: 
337: On the other hand, one can define an (a priori different) divisor $D$ in
338: $\Mg1$ which 
339: is a global analogue of the preimage of $C_{g-i-1}-C_i$ in 
340: $C^{g-i}\times C^{i+1}$ via
341: the difference map. It is convenient to see $D$ as the locus of curves 
342: with marked points $(C,x_1,\ldots ,x_{g-i},y_1,\ldots ,y_{i+1})$ having a
343: $\mathfrak g^1_g$ 
344: which contains $x_1,\ldots , x_{g-i}$ in a fiber and $y_1, \ldots ,
345: y_{i+1}$ in a different fiber.
346: An equivalent formulation of the discussion above is that 
347: $D\subseteq\cZ$, and in order for MRC to hold
348:  for all canonical curves one should 
349: have precisely $D = \cZ$. As we show that $D$
350:  is reduced (cf. Proposition \ref{reduced}), it suffices 
351: then to prove that the class of $D$ coincides with that of $\cZ$.
352: To this end we consider the closure of $D$ in the 
353: compactification $\overline{\cM}_{g,g+1}$,
354:  where the corresponding boundary condition is defined
355: by means of limit linear series. The computation of the class of $D$ via 
356: this closure is essentially independent of the rest of 
357: the paper. 
358: It relies on degeneration and enumerative techniques 
359: in the spirit of \cite{HM} and \cite{EH1}.
360: 
361: The results of both this and the computation
362:  of the class of $\mathcal{Z}$ are summarized 
363: in the following theorem. For the statement,
364:  we recall that ${\rm Pic}(\mathcal{M}_{g,n})_{\QQ}$
365: is generated by the class $\lambda$ of
366:  the Hodge bundle and the classes $\psi_j$, $1\leq j\leq n$,
367: where $\psi_j:= c_1(p_j^* \omega)$,
368:  with $\omega$ the relative dualizing sheaf on the universal 
369: curve $\mathcal{C}_g \rightarrow
370:  \mathcal{M}_g$ and $p_j : \Mgn \rightarrow \mathcal{C}_g$
371: the projection onto the $j$-th factor.
372: 
373: \begin{introtheorem}
374: The divisors $\cZ$ and $D$ defined above have the 
375: same class in $\Pic(\cM_{g,g+1})_{\QQ}$, namely
376: $$-\Bigl({g-1\choose i}-10{g-3\choose {i-1}}\Bigr)
377: \lambda+{{g-2}\choose i}\Psi_x
378: +{{g-2}\choose{i-1}}\Psi_y,$$
379: where $\Psi_x=\sum_{j=1}^{g-i}\psi_j$ and $\Psi_y=\sum_{j=g-i+1}^{g+1}\psi_j$. 
380: In particular $D = \cZ$.
381: \end{introtheorem}
382: 
383: As mentioned above, this implies that $\wedge^i Q$ always has a 
384: theta divisor, for all $i$, so equivalently that MRC holds
385: for an arbitrary canonical curve. We record the more 
386: precise identification of this theta divisor, which now follows
387: in general.
388: 
389: \begin{introcorollary}
390: For any nonhyperelliptic curve $C$, $\Theta_{\wedge^i Q} = C_{g-i-1}
391: - C_i$. 
392: \end{introcorollary}
393: 
394: In this particular form, our result answers positively a conjecture 
395: of R. Lazarsfeld.
396: It is worth mentioning that it also answers negatively a question that
397: was raised in connection with \cite{popa}, namely if $\wedge^i Q$ provide
398: base points for determinant linear series on appropriate moduli
399: spaces of vector bundles.
400: 
401: 
402: The paper is structured as follows. In Section 1 we give some 
403: equivalent formulations of the Minimal Resolution Conjecture and 
404: we describe the vector bundle setup used in the rest of the paper.
405: In Section 2 we treat the case of curves embedded with large degree,
406: proving IGC and showing that MRC fails. Section 3 is devoted to the 
407: main result, namely the proof of MRC for canonical curves, and here 
408: is where we look at the relationship with difference varieties and 
409: moduli spaces of curves with marked points. The divisor 
410: class computation in $\overline{\mathcal{M}}_{g,g+1}$, on which part of 
411: the proof relies, is carried out in Section 4 by means of limit linear
412: series.
413: 
414: \medskip
415: \noindent 
416: {\bf Acknowledgments.} We would like to thank D. Eisenbud and 
417: R. Lazarsfeld for sharing with us numerous useful ideas on the subject.
418: 
419: 
420: \section{\bf Several formulations of the Minimal Resolution Conjecture}
421: 
422: \subsection*{\bf Notations and conventions}
423: 
424: We work over an algebraically closed field $k$ which, unless
425: explicitly mentioned otherwise, has
426: arbitrary characteristic. Let $V$ be a vector space over $k$
427: with $\dim_kV=n+1$ and $S=Sym(V)\iso k[X_0,\ldots,X_n]$
428: the homogeneous coordinate ring of the corresponding
429: projective space $\PP V\iso\PP^n$. 
430: 
431: For a finitely generated graded $S$-module $N$,
432: the \emph{Betti~ numbers} $b_{i,j}(N)$ of $N$ are defined from the
433: minimal free resolution $F_{\bullet}$ of $N$ by
434: $$F_i=\oplus_{j\in\ZZ}S(-i-j)^{b_{i,j}(N)}.$$
435: The \emph{Betti ~diagram} of $N$ has in the $(j,i)$-th  position
436: the Betti number $b_{i,j}(N)$. The \emph{regularity}
437: ${\rm reg}(N)$ of $N\neq 0$
438: can be defined as the index of the last nontrivial row
439: in the Betti diagram of $N$ (see \cite{eisenbud},
440: 20.5 for
441: the connection with the cohomological definition).
442: 
443: We will use the computation of Betti numbers via
444: Koszul cohomology: $b_{i,j}(N)$ is the dimension over $k$
445: of the cohomology of the following piece of the
446: Koszul complex:
447: $$\wedge^{i+1}V\tensor N_{j-1}\longrightarrow
448: \wedge^iV\tensor N_j\longrightarrow\wedge^{i-1}V\tensor N_{j+1}$$
449: (see \cite{green} for details).
450: 
451: For an arbitrary subscheme $Z\subseteq \PP^n$, we denote by
452: $I_Z\subseteq S$ its saturated ideal and let $S_Z=S/I_Z$.
453: We denote by $P_Z$ and $H_Z$ the Hilbert polynomial and
454: Hilbert function of $Z$, respectively. The regularity 
455: ${\rm reg}(Z)$ of $Z$ is defined to be the regularity of
456: $I_Z$, if $Z\neq \PP^n$, and $1$ otherwise. Notice that
457: with this convention, in the Betti diagram of $Z$,
458: which by definition is the Betti diagram of $S_Z$,
459: the last nontrivial row is always indexed by ${\rm reg}(Z)-1$. 
460: 
461: For a projective variety $X$, a line bundle $L$ on $X$, and
462: a linear series $V\subseteq H^0(L)$ which generates $L$, 
463: we denote by $M_V$ the vector bundle which is the kernel of the
464: evaluation map 
465: $$0\longrightarrow M_V \longrightarrow V\tensor\cO_X
466:  \overset{ev}{\longrightarrow} 
467: L\longrightarrow 0.$$
468: When $V=H^0(L)$ we use the
469: notation $M_L:=M_V$. If $C$ is a smooth curve of genus $g\geq 1$,
470: and $\omega_C$ is the canonical line bundle, then 
471: $M_C$ denotes the vector bundle $M_{\omega_C}$. The dual vector bundles
472: will be denoted by $Q_V$, $Q_L$ and $Q_C$, respectively. 
473: Whenever there is no risk of confusion, we will simply write
474: $M$ and $Q$, instead of $M_C$ and $Q_C$.
475: 
476: 
477: \subsection*{\bf The Minimal Resolution Conjecture for points on
478:  embedded varieties.} 
479: In this section $X\subseteq\PP\,V\iso\PP^n$ is a fixed irreducible projective 
480: variety of positive dimension.
481:  We study the Betti numbers of a general set of $\gamma$ points $\Gamma
482: \subseteq X$.
483:  Since the Betti numbers are upper semicontinuous functions,
484: for every positive integer $\gamma$, there is an open subset $U_{\gamma}$ of
485: $X^{\gamma}\setminus\cup_{p\neq q}\{x\colon x_p=x_q\}$ such that
486: for all $i$ and $j$, $b_{i,j}(\Gamma)$ takes its minimum value
487: for $\Gamma\in U_{\gamma}$. Notice that as the regularity is bounded in terms
488: of $\gamma$, we are concerned with finitely many Betti numbers.
489: From now on, $\Gamma$ \emph{general} means $\Gamma\in U_{\gamma}$.
490: 
491: It is easy to determine the Hilbert function of a general set of points 
492: $\Gamma$ 
493: in terms of the Hilbert function of $X$ 
494: (see \cite{mustata}). We have the following:
495: 
496: \begin{proposition}\label{hilbert_function}
497: If $\Gamma\subseteq X$ is a general set of $\gamma$ points, then
498: $$H_{\Gamma}(t)={\rm min}\,\{H_X(t),\gamma\}.$$
499: \end{proposition}
500: 
501: 
502: 
503: To determine the Betti numbers of a general set of points $\Gamma$ is a 
504: much more subtle problem. If $\gamma$ is large enough, then 
505: the Betti diagram of $\Gamma$ looks as follows: in the upper part we
506: have the Betti diagram of $X$ and there are two extra nontrivial rows at
507: the bottom. Moreover, the formula in Proposition~\ref{hilbert_function}
508: gives an expression for the differences of the Betti numbers in these
509: last two rows. We record the formal statement in the following theorem
510: and for the proof we refer to \cite{mustata}.
511: 
512: \begin{theorem}\label{general_results}
513: Assume that $\Gamma\subseteq X$ is a general set of $\gamma$ points,
514: with $P_X(r-1)\leq\gamma<P_X(r)$ for some $r\geq m+1$, where $m={\rm reg}\,X$.
515: 
516: \item{\rm (i)} For every $i$ and $j\leq r-2$, we have $b_{i,j}(\Gamma)
517: =b_{i,j}(X)$.
518: \item{\rm (ii)} $b_{i,j}(\Gamma)=0$, for $j\geq r+1$ and there is an  
519: $i$ such that $b_{i,r-1}(\Gamma)\neq 0$.
520: \item{\rm (iii)} For every $j\geq m$, we have
521: $$b_{i,j}(\Gamma)=b_{i-1,j+1}(I_{\Gamma}/I_X)=b_{i-1,j+1}(\oplus_{l\geq 0}
522: H^0(\cI_{\Gamma/X}(l))).$$
523: \item{\rm (iv)} If $d=\dim\,X$, then for every $i\geq 0$, we have
524: $b_{i+1,r-1}(\Gamma)-b_{i,r}(\Gamma)=Q_{i,r}(\gamma)$, where
525: $$Q_{i,r}(\gamma)=\sum_{l=0}^{d-1}
526: (-1)^l{{n-l-1}\choose{i-l}}\Delta^{l+1}P_X(r+l)-
527: {n\choose i}(\gamma-P_X(r-1)).$$
528: \end{theorem}
529: 
530: 
531: We will focus our attention on the Betti numbers in the bottom two rows
532: in the Betti diagram of $\Gamma$. The equation in Theorem~\ref
533: {general_results} (iv) gives lower bounds for these numbers, namely
534: $b_{i+1,r-1}(\Gamma)\geq {\rm max}\,\{Q_{i,r}(\gamma),0\}$
535: and $b_{i,r}(\Gamma)\geq {\rm max}\,\{-Q_{i,r}(\gamma),0\}$.
536: 
537: \medskip
538: \begin{definition}
539: In analogy with the case $X=\PP^n$
540: (see \cite{lorenzini1} and \cite{lorenzini2}), we say that the 
541: \emph{Minimal Resolution Conjecture} (to which we refer from now on as MRC)
542: holds for a fixed value of $\gamma$ as above
543: if for every $i$ and every  general set $\Gamma$,
544:  $b_{i+1,r-1}(\Gamma)={\rm max}\,\{Q_{i,r}(\gamma),0\}$
545:  and $b_{i,r}(\Gamma)={\rm max}\,\{-Q_{i,r}(\gamma),0\}$.
546:  Equivalently, it says that 
547: $$b_{i+1,r-1}(\Gamma)\cdot b_{i,r}(\Gamma)
548: =0 {\rm ~for~ all~} i.$$
549: \end{definition}
550: 
551: 
552: This conjecture has been extensively studied in the case $X=\PP^n$,
553:  $L=\OO_{\PP^n}(1)$. 
554: It is known to hold for small values of $n$ ($n=2$, $3$ or $4$) and for
555: large values of $\gamma$, depending on $n$, but not in
556: general. In fact, it has been shown that for every $n\geq 6$, $n\neq 9$,
557: MRC fails for almost $\sqrt{n}/2$ values of $\gamma$ (see
558: \cite{epsw}, where one can find also a detailed account of the problem).
559: 
560: Note that the assertion in MRC holds obviously for $i=0$. The first
561: nontrivial case $i=1$ is equivalent by Theorem~\ref{general_results}
562: to saying that the minimal number of generators of $I_{\Gamma}/I_X$
563: is as small as possible. This suggests the following:
564: 
565: \medskip
566: \begin{definition}
567: We say that the \emph{Ideal Generation Conjecture}
568: (IGC, for short) holds for
569: $\gamma$ as above if for a general set of points $\Gamma\subseteq X$ of
570:  cardinality $\gamma$,
571: we have $b_{2,r-1}(\Gamma)\cdot b_{1,r}(\Gamma)=0$.
572: \end{definition}
573: 
574: \smallskip
575: 
576: 
577: \begin{example}(\cite{mustata})\label{case1}
578:  MRC holds for every $X$ when $\gamma=P_X(r-1)$,
579: since in this case $b_{i,r}(\Gamma)=0$ for every $i$.
580: Similarly, MRC holds for every $X$ when $\gamma= P_X(r)-1$,
581: since in this case $b_{1,r-1}=1$ and $b_{i,r-1}(\gamma)=0$ for $i\geq 2$.
582: \end{example}
583: 
584: 
585: 
586: We derive now a cohomological interpretation of MRC.
587: From now on we assume that $X$ is nondegenerate, so that
588: we have $V\subseteq H^0(\cO_X(1))$.
589: Using a standard Koszul cohomology argument, we can express the
590: Betti numbers in the last two rows of the Betti diagram of $\Gamma$
591: as follows.
592: 
593: 
594: \begin{proposition}\label{cohom}
595: With the above notation, we have in general for every $i\geq 0$ 
596: $$b_{i+1,r-1}(\Gamma)=h^0(\wedge^i M_V\tensor\cI_{\Gamma/X}(r)),$$
597: $$b_{i,r}(\Gamma)=h^1(\wedge^i M_V\tensor\cI_{\Gamma/X}(r)).$$
598: \end{proposition}
599: 
600: \begin{proof}
601: We compute the Betti numbers via Koszul cohomology,
602: using the formula in Theorem~\ref{general_results} (iii).
603: 
604: Consider the complex:
605: $$ \wedge^iV\tensor H^0(\I_{\Gamma/X}(r))\overset{f}{\longrightarrow}
606: \wedge^{i-1}V\tensor H^0(\I_{\Gamma/X}(r+1))
607: \overset{h}{\longrightarrow}\wedge^{i-2}V\tensor H^0(\I_{\Gamma/X}(r+2))$$
608: 
609: 
610: Since $H^0(\cI_{\Gamma/X}(r-1))=0$, it follows
611: that $\dim_k({\rm Ker}f)
612: =b_{i+1,r-1}(\Gamma)$ and $\dim_k({\rm Ker}\,h/{\rm Im}\,f)
613: =b_{i,r}(\Gamma)$.
614:  The exact sequence
615: $$0\longrightarrow M_V\longrightarrow V\tensor\OO_X\longrightarrow\OO_X(1)
616: \longrightarrow 0$$ induces long exact sequences
617: $$0\longrightarrow\wedge^i M_V\longrightarrow\wedge^i V\tensor\OO_X
618: \longrightarrow\wedge^{i-1} M_V\tensor\OO_X(1)\longrightarrow 0.\,\,\,\,
619: (*)$$
620: By tensoring with $\cI_{\Gamma/X}(r)$ and taking global sections,
621: we get the exact sequence
622: $$H^0(\wedge^i M_V\tensor\I_{\Gamma/X}(r))
623: \hookrightarrow\wedge^iV\tensor H^0(\cI_{\Gamma/X}(r))
624: \overset{f}{\longrightarrow}\wedge^{i-1}V\tensor H^0(\cI_{\Gamma/X}(r+1)).$$
625: This proves the first assertion in the proposition.
626: 
627: \smallskip
628: 
629: 
630: We have a similar exact sequence:
631: $$H^0(\wedge^{i-1} M_V\tensor\cI_{\Gamma/X}(r+1))
632: \hookrightarrow\wedge^{i-1} V\tensor H^0(\cI_{\Gamma/X}(r+1))
633: \overset{h}{\longrightarrow}
634: \wedge^{i-2}V\tensor H^0(\cI_{\Gamma/X}(r+2)).$$
635: Therefore $b_{i,r}(\Gamma)$ is the dimension
636: over $k$ of the cokernel of
637: $$g\,:\,\wedge^iV\tensor H^0(\cI_{\Gamma/X}(r))\longrightarrow
638: H^0(\wedge^{i-1}M_V\otimes\cI_{\Gamma/X}(r+1)).$$
639: 
640: Using again the exact sequence $(*)$, by tensoring with $\cI_{\Gamma/X}(r)$
641: and taking a suitable part of the long exact sequence, we get:
642: $$\wedge^iV\tensor H^0(\cI_{\Gamma/X}(r))\longrightarrow
643: H^0(\wedge^{i-1} M_V\tensor\cI_{\Gamma/X}(r+1))\longrightarrow$$
644: $$H^1(\wedge^i M_V\tensor\cI_{\Gamma/X}(r))\longrightarrow\wedge^iV
645: \tensor H^1(\cI_{\Gamma/X}(r)).$$
646: 
647: Since ${\rm reg}\,\Gamma\leq r+1$, we have ${\rm reg}\,\cI_{\Gamma/X}\leq r+1$
648: and therefore
649: $H^1(\cI_{\Gamma/X}(r))=0$. From the above exact sequence we see
650: that ${\rm Coker}\,g\simeq
651: H^1(\wedge^i M_V\tensor\cI_{\Gamma/X}(r))$, which proves
652: the second assertion of the proposition.
653: \end{proof}
654: 
655: \smallskip
656: 
657: 
658: 
659: \begin{remark}\label{higher}
660:  The higher cohomology groups 
661: $H^p(\wedge^i M_V\tensor\I_{\Gamma/X}(r))$, $p\geq 2$, always vanish. 
662: Indeed, using the exact sequences in the proof of the proposition,
663: we get $$h^p(\wedge^i M_V\tensor\I_{\Gamma/X}(r))=
664: h^1(\wedge^{i-p+1} M_V\tensor\I_{\Gamma/X}(r+p-1))=
665: b_{i-p+1,r+p-1}(\Gamma)=0.$$
666: Therefore we have $Q_{i,r}(\gamma)=\chi(\wedge^i M_V\tensor\I_{\Gamma/X}(r))$
667: and MRC can be interpreted as saying that for general $\Gamma$,
668: the cohomology of $\wedge^i M_V\tensor\I_{\Gamma/X}(r)$ is supported
669: in cohomological degree either zero or one.
670: \end{remark}
671: 
672: \smallskip
673: 
674: In the case of a curve $C$, 
675: MRC can be reformulated using 
676: Proposition~\ref{cohom}
677: in terms of general line bundles on $C$.
678: We will denote by $\lfloor x\rfloor$ and $\lceil x\rceil$
679: the integers defined by $\lfloor x\rfloor\leq x <\lfloor x\rfloor +1$
680: and $\lceil x \rceil -1<x\leq\lceil x\rceil$. 
681: 
682: \begin{corollary}\label{criterion}
683: Suppose that $C\subseteq\PP\,V$
684: is a nondegenerate, integral curve of arithmetic genus  $g$ 
685: and degree $d$.
686: We consider the following two statements:
687: 
688: \item{\rm (i)} For every  $i$ and for a general line bundle
689: $\xi\in \Pic^j(C)$, where $j=g-1+\lceil{\frac{di}{n}}\rceil$, we have
690: $H^1(\wedge^i M_V\tensor \xi)=0$. 
691: \item{\rm (ii)} For every $i$ and for a general line bundle
692: $\xi\in\Pic^j(C)$, where $j=g-1+\lfloor{\frac{di}{n}}\rfloor$, we have
693: $H^0(\wedge^i M_V\tensor \xi)=0$.
694: 
695: Then MRC holds for $C$ for every $\gamma\geq \max\{g, P_C({\rm reg}\,X)\}$
696:  if and only if
697: both {\rm (i)} and {\rm (ii)} are true. Moreover, if $C$ is locally Gorenstein,
698: then {\rm (i)} and {\rm (ii)} are equivalent.
699: \end{corollary}
700: 
701: 
702: \begin{proof} 
703: If $\gamma\geq g$, then for a general set $\Gamma$ of $\gamma$ points,
704: $\I_{\Gamma/C}$ is a general line bundle on $C$ of degree $-\gamma$.
705: Since in this case $\I_{\Gamma/C}(r)$ is a general line bundle of
706: degree $j=dr-\gamma$ and $d(r-1)+1-g\leq \gamma\leq dr+1-g$,
707: Proposition~\ref{cohom}
708: says that MRC holds for every $\gamma\geq \max\{g, P_C({\rm reg}\,C)\}$
709:  if and only if
710: for every $j$ such that $g-1\leq j\leq d+g-1$ and for
711: a general line bundle $\xi'\in\Pic^j(C)$,
712: either $H^0(\wedge^i M_V\tensor \xi')=0$ or
713: $H^1(\wedge^i M_V\tensor \xi')=0$. 
714: 
715: Since $\dim\,C=1$, we have $b_{i+1,r-1}(\Gamma)-b_{i,r}(\Gamma)=
716: d{n-1\choose i} -(\gamma-P_C(r-1)){n\choose i}$. It follows immediately
717: that $b_{i+1,r-1}(\Gamma)-b_{i,r}(\Gamma)\geq 0$ if and only if
718: $j\geq g-1+di/n$.
719: 
720: The first statement of the corollary follows now from the fact that
721: if $E$ is a vector bundle on a curve and $P$ is a point, then
722: $H^0(E)=0$ implies $H^0(E\tensor\OO(-P))=0$ and
723: $H^1(E)=0$ implies $H^1(E\tensor\OO(P))=0$. The last statement
724: follows from Serre duality and the isomorphism 
725: $\wedge^iQ_V\simeq\wedge^{n-i}M_V\otimes\cO_C(1)$.
726: \end{proof}
727: 
728: \begin{remark}\label{caseIGC}
729:  The corresponding assertion for IGC says that
730: $X$ satisfies IGC for every $\gamma\geq \max\{g, P_C({\rm reg}\,C)\}$
731:  if and only if both 
732: (i) and (ii) are true for $i=1$. Note that if $X$ is locally Gorenstein,
733: then by Serre duality condition (ii) for $i=1$ is equivalent to
734: condition (i) for $i=n-1$.
735: \end{remark}
736: 
737: \begin{remark}\label{half}
738: If $C$ is a locally  Gorenstein integral curve
739: such that $d/n\in\ZZ$, then in order to check MRC for all
740: $\gamma\geq \max\{g, P_C({\rm reg}\,C)\}$,
741:  it is enough to check condition (i) in Corollary~\ref
742: {criterion} only for $i\leq n/2$. Indeed, using Serre duality
743: and Riemann-Roch, we see that the conditions for $i$
744: and $n-i$ are equivalent.
745: \end{remark}
746: 
747: In light of Corollary~\ref{criterion}, we make the following:
748: 
749: \begin{definition}
750: If $C\subseteq \PP\,V$ is a nondegenerate integral curve of arithmetic genus
751: $g$ and regularity $m$, we say that $C$ satisfies MRC if 
752: a general set of $\gamma$ points on $C$ satisfies MRC
753: for every $\gamma\geq{\rm max}\{g, P_C(m)\}$. If $L$ is a very ample line
754: bundle on a curve $C$ as before, we say that $L$ satisfies MRC if
755: $C\subset\PP\,H^0(L)$ satisfies MRC. Analogous definitions are made for IGC.
756: \end{definition}
757: 
758: \begin{example}({\bf Rational quintics in $\PP^3$.})\label{quintics}
759: We illustrate the above discussion in the case of smooth rational quintic
760: curves in $\PP^3$. We consider two explicit examples, the first when the curve
761: lies on a (smooth) quadric and the second when it does not.
762: Let $X$ be given parametrically by
763: $(u,v)\in\PP^1\longrightarrow (u^5,u^4v,uv^4,v^5)\in\PP^3$, so that
764: it lies on the quadric $X_0X_3=X_1X_2$. The Betti diagram of $X$ is
765: \begin{center}
766: \renewcommand{\arraystretch}{1.25}
767: \begin{tabular}[l]{c|c}
768: \hline
769: 0 & 1 -- -- -- \\
770: 1 & -- 1 -- --  \\
771: 2 & -- -- -- -- \\
772: 3 & -- 4 6 2  \\
773: \end{tabular}
774: \end{center}
775: and if $\Gamma\subset X$ is a set of $28$ points, then the Betti diagram of
776: $\Gamma$ is
777: \begin{center}
778: \renewcommand{\arraystretch}{1.25}
779: \begin{tabular}[l]{c|c}
780: \hline
781: 0 & 1 -- -- -- \\
782: 1 & -- 1 -- -- \\
783: 2 & -- -- -- -- \\
784: 3 & -- 4 6 2 \\
785: 4 & -- -- -- -- \\
786: 5 & -- 3 4 1 \\
787: 6 & -- -- 2 2 \\
788: \end{tabular}
789: \end{center}
790: As $b_{3,5}(\Gamma)=1$ and $b_{2,6}(\Gamma)=2$, we see that MRC
791: is not satisfied by $X$ for this number of points.
792: 
793: Let now $Y$  be the curve given parametrically by
794: $(u,v)\in\PP^1\longrightarrow (u^5+u^3v^2, u^4v-u^2v^3,uv^4,v^5)\in\PP^3$.
795: In this case $Y$ does not lie on a quadric, and in fact, its
796: Betti diagram is given by
797: \begin{center}
798: \renewcommand{\arraystretch}{1.25}
799: \begin{tabular}[l]{c|c}
800: \hline
801: 0 & 1 -- -- -- \\
802: 1 & -- -- -- -- \\
803: 2 & -- 4 3 -- \\
804: 3 & -- 1 2 1 \\
805: \end{tabular}
806: \end{center}
807: If $\Gamma'\subset Y$ is a set of $28$ points, then the Betti diagram of
808: $\Gamma'$ is
809: \begin{center}
810: \renewcommand{\arraystretch}{1.25}
811: \begin{tabular}[l]{c|c}
812: \hline
813: 0 & 1 -- -- -- \\
814: 1 & -- -- -- -- \\
815: 2 & -- 4 3 -- \\
816: 3 & -- 1 2 1 \\
817: 4 & -- -- -- -- \\
818: 5 & -- 3 4 -- \\
819: 6 & -- -- 1 2 \\
820: \end{tabular}
821: \end{center}
822: which shows that MRC is satisfied for $Y$ and this number of points.
823: 
824: These two examples show the possible behavior with respect to the
825: MRC for smooth rational quintics in $\PP^3$. The geometric condition
826: of lying on a quadric translates into a condition on the splitting type
827: of $M_V=\Omega_{\PP^3}(1)\vert_X$. More precisely, it is proved in
828: \cite{ev} that if $X\subset\PP^3$ is a smooth rational quintic curve, then
829: $X$ lies on a quadric if and only if we have
830: $\Omega_{\PP^3}(1)\vert_X\simeq\cO_{\PP^1}(-3)\oplus\cO_{\PP^1}(-1)^{\oplus 2}$
831: (the other possibility, which is satisfied by
832: a general such quintic, is that $\Omega_{\PP^3}(1)\vert_X
833: \simeq\cO_{\PP^1}(-1)\oplus\cO_{\PP^1}(-2)^{\oplus 2}$).
834: Corollary~\ref{criterion} explains therefore the behaviour with respect
835: to MRC in the above examples.
836: \end{example}
837: 
838: 
839: 
840: \section{\bf Curves of large degree and a counterexample to MRC}
841: 
842: In this section we assume that $C$ is a smooth projective curve
843: of genus $g$ and $L$ is a very ample line bundle on $C$. Our aim is to 
844: investigate whether $C$ satisfies MRC, or at least IGC,  for every
845: $\gamma\geq g$, in the embedding given by the complete linear series $|L|$. 
846: As before, $m$ will denote the regularity of $C$.
847: 
848: \begin{example}\label{g_small}
849: If $g=0$ or $1$, then $C$ satisfies MRC for all $\gamma\geq P_C(m)$ 
850: in every embedding given by a complete linear series
851: (see \cite{mustata}, Proposition 3.1).
852: \end{example}
853: 
854: In higher genus we will concentrate on the study
855:  of MRC for canonical curves and 
856: curves embedded with high degree, in direct analogy with the syzygy questions 
857: of Green-Lazarsfeld (cf. \cite{gl1}, \cite{gl2}, \cite{gl3}).
858: The main conclusion of this section will be that, while 
859: IGC is satisfied in both situations, the high-degree embeddings
860: always fail to satisfy MRC at a well-specified spot in the Betti diagram. 
861: This is in contrast with our main result, proved in \S3, that MRC 
862: always holds for canonical curves, and the arguments involved here  
863: provide an introduction to that section. The common theme of the proofs is the
864: vector bundle interpretation of MRC described in \S1.
865: 
866: \medskip
867: \noindent 
868: {\bf Review of filtrations for $Q_L$ and $Q$ \cite{lazarsfeld}.}
869:  Here we recall a basic property of 
870: the vector bundles $Q_L$ which will be essential for our arguments. 
871: Let $L$ be a very ample line bundle on $C$
872: of degree $d$, and recall from \S1 that $Q_L$ is given by the defining sequence
873: $$0\longrightarrow L^{-1}\longrightarrow H^0(L)^{*}\otimes
874:  \OO_{C}\longrightarrow
875: Q_{L}\longrightarrow 0.$$
876: Assume first that $L$ is non-special and $x_{1},\ldots,x_{d}$ are the points
877: of a general hyperplane section of $C\subseteq\PP H^{0}(L)$.
878:  One shows (see e.g. \cite{lazarsfeld} \S1.4) 
879: that there exists an exact sequence 
880: \begin{equation}\label{large_degree_filtration}
881: 0\longrightarrow \underset{i\in \{1,\ldots, d-g-1\}}
882: {\bigoplus}\OO_{C}(x_{i})\longrightarrow Q_{L}
883: \longrightarrow \OO_{C}(x_{d-g}+ \ldots +x_{d})\longrightarrow 0.
884: \end{equation}
885: On the other hand, assuming that $C$ is nonhyperelliptic and $L=\omega_C$, if 
886: $x_{1},\ldots,x_{2g-2}$ are the points of a general hyperplane section,
887: the analogous sequence reads:
888: \begin{equation}\label{canonical_filtration}
889: 0\longrightarrow \underset{i\in \{1,\ldots, g-2\}}
890: {\bigoplus}\OO_{C}(x_{i})\longrightarrow Q
891: \longrightarrow \OO_{C}(x_{g-1}+ \ldots +x_{2g-2})\longrightarrow 0.
892: \end{equation}
893: 
894: 
895: \medskip
896: We start by looking at the case of curves embedded with large degree.
897:  The main results 
898: are summarized in the following:
899: 
900: \begin{theorem}\label{curves_of_large_degree}
901: {\rm (a)} If $L$ is a very ample line bundle of degree $d\geq 2g$, then $IGC$ 
902: holds for $L$.
903: \newline
904: {\rm (b)} If $g\geq 4$ and $L$ is a line bundle of degree $d\geq 2g+10$,
905:  then there exists 
906: a value of $\gamma$ such that $C\subseteq\PP H^{0}(L)$ does not satisfy MRC 
907: for $i=\lfloor\frac{g+1}{2}\rfloor$. The same holds if $g\geq 14$ and
908:  $d\geq 2g+5$.
909: \end{theorem}
910: 
911: \begin{proof}
912: (a)
913: Let $L$ be a very ample line bundle of degree $d\geq 2g$.
914: By Corollary \ref{criterion} and Serre duality, it is easy to see that IGC 
915: holds for $L$ if:
916: \medskip
917: \newline
918: (i) $h^{1}(Q_{L}\otimes \eta)=0$ for $\eta\in {\rm Pic}^{g-2}(C)$ general
919: \newline
920: and
921: \newline
922: (ii) $h^{0}(Q_{L}\otimes \eta)=0$ for $\eta\in {\rm Pic}^{g-3}(C)$ general.
923: 
924: Condition (i) is a simple consequence of the filtration
925:  (\ref{large_degree_filtration}). 
926: More precisely, if $x_{1},\ldots,x_{d}$ are the points
927: of a general hyperplane section of $C\subseteq\PP H^{0}(L)$,
928:  from the exact sequence
929: $$0\longrightarrow \underset{i\in \{1,\ldots, d-g-1\}}
930: {\bigoplus}\OO_{C}(x_{i})\longrightarrow Q_{L}
931: \longrightarrow \OO_{C}(x_{d-g}+ \ldots +x_{d})\longrightarrow 0.$$
932: we conclude that it would be enough to prove:
933: $$h^{1}(\eta(x_{i}))=0 {\rm ~and~} h^{1}(\eta(x_{d-g}+ \ldots +x_{d}))=0$$
934: for $\eta\in {\rm Pic}^{g-2}(C)$ general. Now for every
935:  $i\in \{1,\ldots ,d-g-1\}$, $\eta(x_{i})$
936: is a general line bundle of degree $g-1$, so $h^{1}(\eta(x_{i}))=0$.
937:  On the other hand
938: ${\rm deg}~\eta(x_{d-g}+ \ldots +x_{d})\geq2g-1$, so clearly
939:  $h^{1}(\eta(x_{d-g}+ \ldots +x_{d}))=0$.
940: 
941: For condition (ii) one needs a different argument.
942:  By twisting the defining sequence of $Q_{L}$: 
943: $$0\longrightarrow L^{-1}\longrightarrow
944: H^{0}(L)^{*}\otimes \OO_{C}\longrightarrow
945: Q_{L}\longrightarrow 0$$
946: by $\eta\in {\rm Pic}^{g-3}(C)$ general and taking cohomology,
947:  we see that (ii) holds if and only 
948: if the map
949: $$\alpha^{*}:H^{1}(L^{-1}\otimes \eta)\rightarrow
950: H^{0}(L)^{*}\otimes H^{1}(\eta)$$
951: is injective, or dually if and only if the cup-product map
952: $$\alpha: H^{0}(L)\otimes H^{0}(\omega_{C}
953: \otimes \eta^{-1})\rightarrow H^{0}(L\otimes \omega_{C}
954: \otimes \eta^{-1})$$
955: is surjective. We make the following:
956: \medskip
957: \newline
958: \emph{Claim.} $|\omega_{C}\otimes \eta^{-1}|$
959:  \emph{is a base point free pencil}.
960: 
961: \medskip
962: Assuming this for the time being, one can apply the
963:  Base Point Free Pencil Trick
964: (see \cite{ACGH} III \S3) to conclude that
965: $${\rm Ker~}\alpha= H^{0}(L\otimes \omega_{C}^{-1}\otimes \eta).$$
966: But $L\otimes \omega_{C}^{-1}\otimes \eta$ is a general line bundle of
967:  degree $d-g-1\geq g-1$  
968: and so $h^{1}(L\otimes \omega_{C}^{-1}\otimes \eta)=0$.
969:  By Riemann-Roch this means
970: $h^{0}(L\otimes \omega_{C}^{-1}\otimes \eta)=d-2g$.
971:  On the other hand $h^{0}(L)=d-g+1$,
972: $h^{0}(\omega_{C}\otimes \eta^{-1})=2$ and
973:  $h^{0}(L\otimes \omega_{C}\otimes \eta^{-1})=d+2$, so
974: $\alpha$ must be surjective.
975: 
976: We are only left with proving the claim. Since $\eta\in {\rm Pic}^{g-3}(C)$
977:  is general,
978: $h^{0}(\eta)=0$, and so we easily get:
979: $$ h^{0}(\omega_{C}\otimes \eta^{-1})=h^{1}(\eta)=g-1-(g-3)=2.$$
980: Also, for every $p\in C$, $\eta(p)\in {\rm Pic}^{g-2}(C)$ is general,
981:  hence still noneffective. Thus:
982: $$h^{0}(\omega_{C}\otimes \eta^{-1}(-p))=h^{1}(\eta(p))=g-1-(g-2)=1.$$
983: This implies that $|\omega_{C}\otimes \eta^{-1}|$ is base point free.
984: 
985: \medskip
986: (b) Here we follow an argument in \cite{popa} leading to the required
987: nonvanishing statement.
988: First note that it is clear from (\ref{large_degree_filtration})
989:  that for every $i$ with $1\leq i\leq d-g-1$
990: there is an inclusion
991: $$\OO_{C}(x_{1}+\ldots +x_{i})\hookrightarrow \wedge^{i}Q_{L},$$
992: where $x_{1},\ldots , x_{i}$ are general points on $C$.
993: This immediately implies that 
994: $$h^{0}(\wedge^{i}Q_{L}\otimes \OO_{C}(E_{i}-D_{i}))\neq 0,$$
995: where $E_{i}$ and $D_{i}$ are general effective divisors on $C$ of degree $i$.
996: On the other hand 
997: we use the fact (see e.g. \cite{ACGH} Ex. V. D) 
998: that every line bundle $\xi \in {\rm Pic}^{0}(C)$ can be written as a 
999: difference
1000: $$\xi=\OO_{C}(E_{\lfloor\frac{g+1}{2}\rfloor}-
1001: D_{\lfloor\frac{g+1}{2}\rfloor}),$$
1002: which means that 
1003: $$h^{0}(\wedge^{\lfloor\frac{g+1}{2}\rfloor}
1004: Q_{L}\otimes \xi)\neq 0,~\forall \xi 
1005: \in {\rm Pic}^{0}(C) {\rm ~general~}.$$
1006: Now by Serre duality:
1007: $$H^{0}(\wedge^{i}Q_{L}\otimes \xi)\cong H^{1}(\wedge^{i}M_{L}\otimes 
1008: \omega_{C}\otimes \xi^{-1})^{*},$$
1009: so that Corollary \ref{criterion} easily implies that $C$
1010:  does not satisfy MRC for 
1011: $i=\lfloor\frac{g+1}{2}\rfloor$ as long as $2g-2\geq g-1+ \frac{di}{d-g}$.
1012: A simple computation gives then the stated conclusion.
1013: \end{proof}
1014: 
1015: \begin{remark}
1016: Motivation for the argument in (b) above was quite surprisingly
1017: provided by the study \cite{popa} of 
1018: the base locus of the determinant linear series on the moduli space $SU_C(r)$ 
1019: of semistable bundles of rank $r$ and trivial determinant on a curve $C$. 
1020: In fact this argument produces 
1021: explicit base points for the determinant linear series under appropriate
1022:  numerical conditions.
1023: \end{remark}
1024: 
1025: \begin{remark}
1026: The technique in Theorem \ref{curves_of_large_degree} (b) can be extended
1027:  to produce 
1028: examples of higher dimensional varieties for which appropriate
1029:  choices of $\gamma$ 
1030: force the failure of MRC for general sets of $\gamma$ points.
1031:  More precisely, the 
1032: varieties in question are projective bundles $\PP E \rightarrow C$ over a
1033:  curve $C$, 
1034: associated to very ample 
1035: vector bundles $E$ on $C$ of arbitrary rank and large degree,
1036:  containing sub-line bundles of 
1037: large degree. Using the interpretation given in Proposition \ref{cohom},
1038:  the problem 
1039: is reduced to a cohomological question about the exterior powers
1040:  $\wedge^i M_E$, where 
1041: $M_E$ is defined analogously as the kernel of the evaluation map
1042: $$0\longrightarrow M_E \longrightarrow H^0(E)\otimes
1043:  \OO_X\longrightarrow E\longrightarrow 0.$$
1044: This question is then treated essentially as above,
1045:  and we do not enter into details. 
1046: Unfortunately once a bundle $E$ of higher rank is fixed,
1047:  this technique does not seem 
1048: to produce couterexamples for arbitrarily large values of $\gamma$,
1049:  as in the case of 
1050: line bundles. Such examples would be very interesting,
1051:  in light of the asymptotically nice
1052: behavior of general points in $\PP^n$ (cf. \cite{simpson}).
1053: \end{remark}
1054: 
1055: Finally we turn to the case of canonical curves with the goal of
1056:  providing an introduction 
1057: to the main result in Section 3. Let $C$ be 
1058: a nonhyperelliptic curve of genus $g$, $V=H^0(\omega_C)$
1059: and $C\hookrightarrow\PP V\iso\PP^{g-1}$ the canonical embedding.
1060: We note here that an argument similar to
1061:  Theorem \ref{curves_of_large_degree} (a) 
1062: immediately implies IGC for $C$. This will be later subsumed in the general 
1063: Theorem \ref{canonical_curves}.  
1064: 
1065: \begin{proposition}\label{igc_canonical}
1066: IGC holds for the canonical curve $C$.
1067: \end{proposition}
1068: \begin{proof}
1069: The argument is similar (and in fact simpler) to the proof of (ii) in 
1070: Theorem \ref{curves_of_large_degree} (a). 
1071: In this case, again by interpreting Proposition~\ref{cohom}
1072: (se Remark~\ref{caseIGC}
1073:  IGC holds if and only if
1074: $$H^{0}(Q\otimes \xi)=0~{\rm  for}~\xi\in {\rm Pic}^{g-3}(C) ~{\rm general}.$$
1075: This is in turn equivalent to the surjectivity of the multiplication map:
1076: $$H^{0}(\omega_{C})\otimes H^{0}(\omega_{C}\otimes \xi^{-1})\rightarrow
1077: H^{0}(\omega_{C}^{\otimes 2}\otimes \xi^{-1}),$$
1078: which is again a quick application  of the Base Point Free Pencil Trick.
1079: \end{proof}
1080: 
1081: The geometric picture in the present case of canonical curves
1082:  can be described a
1083: little more precisely. In fact, for $\xi\in {\rm Pic}^{g-3}(C)$, we have
1084: $$\mu(Q\otimes \xi)=g-1,$$
1085: where  $\mu(E):={\rm deg}\,(E)/{\rm rk}\,(E)$
1086: denotes in general the \emph{slope} of the vector bundle $E$.
1087: By standard determinantal results, the subset 
1088: $$\Theta_{Q}:=\{\xi~|~ h^{0}(Q\otimes \xi)\neq 0\}\subseteq{\rm Pic}^{g-3}(C)$$
1089: is either a divisor or the whole variety.
1090: The statement of IGC is then equivalent to saying that  
1091: $\Theta_{Q}$ is indeed a divisor
1092: in ${\rm Pic}^{g-3}(C)$ (one says that $Q$ \emph{has a theta divisor}). 
1093: A simple filtration argument based on the sequence (3) above shows that in fact
1094: $$\Theta_{Q}=C_{g-2}-C:=\{\OO_{C}(p_{1}+\ldots+p_{g-2}-q)~|~p_{1},\ldots ,
1095: p_{g-2}, q \in C\},$$
1096: which has already been observed by Paranjape and Ramanan in
1097: \cite{paranjape}
1098: A generalization of this observation to the higher exterior powers
1099:  $\wedge^iQ$ will be 
1100: the starting point for our approach to proving MRC for canonical curves in
1101:  what follows.
1102: 
1103: 
1104: 
1105: \section{\bf MRC for canonical curves}
1106: 
1107: In this section $C$ will be a canonical curve, i.e. a smooth curve
1108: of genus $g$ embedded in $\PP^{g-1}$ by the canonical linear series
1109: $|\omega_C|$ (in particular $C$ is not hyperelliptic). Our goal is to prove
1110: the following:
1111: 
1112: \begin{theorem}\label{canonical_curves}
1113: If $C$ is a canonical curve, then $C$ satisfies MRC.
1114: \end{theorem}
1115: 
1116: \begin{remark}
1117: In fact, since $C$ is canonically embedded, its regularity is
1118: $m=4$, and as $g\geq 3$ we always have $P_C(m)=7(g-1)\geq g$.
1119: Thus the statement means that MRC holds for every $\gamma \geq P_C(m)$.
1120: \end{remark}
1121: 
1122: The general condition required for a curve to 
1123: satisfy MRC which was stated in Corollary \ref{criterion}
1124: (see also Remark~\ref{half}) 
1125: takes a particularly clean form in the case of canonical embeddings.
1126: We restate it for further use.
1127: 
1128: \begin{lemma}\label{restatement}
1129: Let $C$ be a canonical curve. Then $C$ satisfies MRC if and only if, for all 
1130: $1 \leq i\leq \frac{g-1}{2}$ we have
1131: $$h^0(\wedge^i M\otimes \eta) =
1132: h^1(\wedge^i M\otimes \eta)=0, ~{\rm for}~ \eta\in 
1133: {\rm Pic}^{g+2i-1}(C)~{\rm general},$$ 
1134: or equivalently
1135: $$(*) ~~h^0(\wedge^i Q\otimes \xi) =h^1(\wedge^i Q\otimes \xi)=0, 
1136: ~{\rm for}~ \xi\in \Pici~
1137: {\rm general}.$$ 
1138: \end{lemma} 
1139: 
1140: \begin{remark}
1141: Note that $\mu(Q) = 2$, so $\mu(\wedge^i Q)=2i\in \ZZ$.
1142:  This means that the condition 
1143: $(*)$ in Lemma~\ref{restatement} is equivalent to saying that $\wedge^i Q$
1144:  has a theta divisor (in 
1145: $\Pici$), which we denote $\Theta_{\wedge^i Q}$.
1146:  In other words, the set defined by 
1147: $$\Theta_{\wedge^i Q}:= \{ \xi\in \Pici~|~h^0(\wedge^i Q\otimes \xi)\neq 0\}$$
1148: with the scheme structure
1149:  of a degeneracy locus of a map of vector bundles of the same 
1150: rank is an actual divisor as expected (cf. \cite{ACGH} II \S4).
1151: \end{remark}
1152: 
1153: 
1154: \subsection*{\bf Hyperelliptic curves.}  Note that the statement $(*)$
1155: in Lemma~\ref{restatement} makes sense even for hyperelliptic curves.
1156:  Again $Q$ is the dual of $M$, where $M$ is the 
1157: kernel of the evaluation map for the canonical line
1158: bundle.
1159: Therefore we will say slightly abusively 
1160: that MRC is satisfied for some smooth curve of genus
1161: $g\geq 2$ if $(*)$ is satisfied for all $i$, $1\leq i\leq (g-1)/2$.
1162: In fact, the hyperelliptic case is the only one for which we can give a direct
1163: argument.
1164: 
1165: \begin{proposition}\label{hyperelliptic}
1166: MRC holds for hyperelliptic curves.
1167: \end{proposition}
1168: 
1169: \begin{proof}
1170: We show that for every $i$,
1171: $h^{0}(\wedge^{i}Q\otimes \xi)=0$, if $\xi\in {\rm Pic}^{g-2i-1}(C)$
1172: is general.
1173: Since $C$ is hyperelliptic, we have a degree two morphism
1174: $f:C\rightarrow\PP^1$ and if $L=f^*(\OO_{\PP^1}(1))$,
1175: then $\omega_C=L^{g-1}$. Therefore the morphism
1176: $\widetilde{f}\,:\,C\longrightarrow\PP^{g-1}$ defined by $\omega_C$
1177: is the composition of the Veronese embedding $\PP^1\hookrightarrow
1178: \PP^{g-1}$ with $f$. Note that we have $M=\widetilde{f}^*(\Omega_{\PP^{g-1}}
1179: (1))$.
1180: 
1181: Since on $\PP^1$ we have the exact sequence:
1182: $$0\longrightarrow
1183: \OO_{\PP^1}(-1)^{\oplus(g-1)}\longrightarrow
1184:  H^0(\OO_{\PP^1}(g-1))\tensor\OO_{\PP^1}
1185: \longrightarrow\OO_{\PP^1}(g-1)\longrightarrow 0,$$
1186: we get $M\iso (L^{-1})^{\oplus(g-1)}$.
1187: Therefore for every $i$, we have 
1188: $$\wedge^{i}Q\iso(L^i)^{\oplus{{g-1}\choose i}}.$$
1189: Now if $\xi\in\Pic^{g-2i-1}(C)$ is general,
1190: then $\xi\tensor L^{i}$ is a general line bundle
1191: of degree $g-1$ and so $h^0(\wedge^{i}Q\tensor \xi)=0$.
1192: \end{proof}
1193: 
1194: 
1195: 
1196: \subsection*{\bf Theta divisors and difference varieties for a fixed curve.}
1197: We noted above that MRC is satisfied for $C$ if and only if
1198:  $\Theta_{\wedge^i Q}$ 
1199: is a divisor. We now identify precisely what the divisor should be,
1200:  assuming that 
1201: this happens. (At the end of the day this will hold for all canonical curves.)
1202: Recall that by general theory, whenever a divisor,
1203: $\Theta_{\wedge^i Q}$ belongs 
1204: to the linear series
1205: $|{{g-1} \choose {i}}\Theta|$, where we slightly abusively  
1206: denote by $\Theta$ a certain theta divisor on ${\rm Pic}^{g-2i-1}(C)$ (more
1207: precisely $\Theta_N$, where $N$ is a ${{g-1} \choose {i}}$-th root of 
1208: ${\rm det}(\wedge^i Q)$).
1209: 
1210: From now on we always assume that we are in this situation.
1211: The Picard variety
1212:  ${\rm Pic}^{g-2i-1}(C)$ contains a \emph{difference subvariety}
1213: $C_{g-i-1}-C_{i}$ defined as the image of the difference map
1214: $$\phi:C_{g-i-1}\times C_{i}\longrightarrow {\rm Pic}^{g-2i-1}(C)$$
1215: $$(x_1+\ldots +x_{g-i-1},y_1+\ldots +y_i)\rightarrow \mathcal{O}_C
1216: (x_1+\ldots +x_{g-i-1} -y_1 -\ldots - y_i).$$
1217: The geometry of the difference varieties has interesting links with the 
1218: geometry of the curve \cite{Rob} and \cite{ACGH} (see below). 
1219: The key observation is that our theta divisor 
1220: is nothing else but the difference variety above.
1221: 
1222: \begin{proposition}\label{1}
1223: For every smooth curve $C$ of genus $g$, we have
1224: $$C_{g-i-1} - C_{i}\subseteq\Theta_{\wedge^i Q}.$$
1225: Moreover, if $C$ is nonhyperelliptic and $\Theta_{\wedge^i Q}$
1226: is a divisor, then the above inclusion is an equality.
1227: \end{proposition}
1228: 
1229: We start with a few properties of the difference varieties, which for 
1230: instance easily imply that $C_{g-i-1} - C_{i}$ is a divisor. More 
1231: generally, we study the difference variety $C_a - C_b$, $a\geq b$, defined 
1232: analogously. Note that this study is suggested in a series of exercises 
1233: in [ACGH] Ch.V.D and Ch.VI.A in the case $a=b$, but the formula in V.D-3 
1234: there giving the cohomology class of $C_a - C_a$ is unfortunately incorrect,
1235:  as we first learned from R. Lazarsfeld. 
1236: The results we need are collected in the following:
1237: 
1238: \begin{proposition}\label{2}
1239: (a) Assume that $1\leq b\leq a\leq \frac{g-1}{2}$. Then the difference map:
1240: $$\phi: C_a \times C_b \longrightarrow C_a - C_b\subseteq{\rm
1241: Pic}^{a-b}(C)$$
1242: is birational onto its image if $C$ is
1243: nonhyperelliptic. When $C$ is hyperelliptic, $\phi$ has
1244:  degree ${{a} \choose {b}}\,2^b$ onto its image.
1245: \newline
1246: \noindent
1247: (b) If $C$ is nonhyperelliptic, the cohomology 
1248: class $c_{a,b}$ of $C_a - C_b$ in ${\rm Pic}^{a-b}(C)$ is given by
1249: $$c_{a,b} = {{a+b} \choose {a}}\,\theta^{g-a-b},$$
1250: where $\theta$ is the class of a theta divisor.
1251: \end{proposition}
1252: 
1253: Assuming this, the particular case $a = g-i-1$ and $b = i$ quickly
1254: implies
1255: the main result.
1256: 
1257: \begin{proof}(of Proposition \ref{1})
1258: From Proposition \ref{2} (b) we see that if $C$ is nonhyperelliptic, then
1259: the class of $C_{g-i-1} - C_{i}$ is given by:
1260: $$c_{g-i-1,i} = {{g-1} \choose {i}} \theta.$$
1261: On the other hand, as $\Theta_{\wedge^i Q}$ is associated to the
1262: vector bundle $\wedge^i Q$, if it is a divisor, then its cohomology class is 
1263: ${{g-1} \choose {i}}\theta$ (recall that $\Theta_{\wedge^i Q}$ has the same 
1264: class as ${{g-1}
1265: \choose {i}}\Theta$). As in this case both $\Theta_{\wedge^i Q}$ and
1266:  $C_{g-i-1} - C_{i}$ are
1267: divisors, in order to finish the proof of the proposition it is enough
1268: to prove the first statement.
1269:  
1270: To this end, we follow almost verbatim the argument in
1271:  Theorem \ref{curves_of_large_degree}
1272: (b). Namely, the filtration (2) in \S2 implies that for every $i\geq 1$
1273:  there is an inclusion:
1274: $$\OO_C(x_1+\ldots + x_i)\hookrightarrow \wedge^i Q,$$
1275: where $x_1,\ldots , x_i$ are general points on $C$. This means that 
1276: $$h^0(\wedge^i Q\otimes \OO_C(E_{g-i-1}-D_i))\neq 0$$
1277: for all general effective divisors $E_{g-i-1}$
1278:  of degree $g-i-1$ and $D_i$ of degree $i$,
1279: which gives the desired inclusion.
1280: \end{proof}
1281: 
1282: We are left with proving Proposition \ref{2}. This follows by more or less
1283: standard arguments in the study of Abel maps and Poincar\' e formulas for
1284: cohomology classes of images of symmetric products.
1285: 
1286: \begin{proof}(of Proposition \ref{2})
1287: (a) This is certainly well known (cf. \cite{ACGH} Ch.V.D), and we do not reproduce 
1288: the proof here.
1289: \newline
1290: \noindent
1291: (b) Assume now that $C$ is nonhyperelliptic, so that 
1292: $$\phi: C_a \times C_b \longrightarrow C_a - C_b$$
1293: is birational onto its image. For simplicity we will map 
1294: everything to the Jacobian of $C$, so fix a point $p_0 \in C$ and consider 
1295: the commutative diagram:
1296: $$\xymatrix{
1297: & C^{a+b} \ar[dl]_{\psi} \ar[dr]^{\alpha} \\
1298: C_a\times C_b \ar[r]^{\phi} & C_a - C_b \ar[r]^{-(a-b)p_0} & J(C)}$$
1299: where $C^{a+b}$ is the $(a+b)$-th cartesian product of the curve and the
1300: maps 
1301: are either previously defined or obvious.
1302: We will in general denote by $[X]$ the fundamental class of the compact
1303: variety $X$. Since $\psi$ clearly has degree $a!\cdot b!$, and since
1304: $\phi$ is birational by (a), we have:
1305: $$\alpha_{*}[C^{a+b}] = a!\cdot b!\cdot c_{a,b}.$$
1306: This means that it is in fact enough to prove that $\alpha_{*}[C^{a+b}]
1307: = (a+b)!\cdot \theta^{g-a-b}$, and note that 
1308: $(a+b)!\cdot\theta^{g-a-b}$ is the same as the class
1309: $u_{*}[C^{a+b}]$, where $u$ is the usual Abel map:
1310: $$u: C^{a+b}\longrightarrow J(C)$$
1311: $$(x_1,\ldots , x_{a+b})\rightarrow \OO_C(x_1 + \ldots + x_{a+b} - (a+b)p_0).$$
1312: The last statement is known as Poincar\'e's formula (see e.g. [ACGH] I
1313: \S5).
1314: We are now done by the following lemma,  which essentially says that 
1315: adding or subtracting points is the same when computing cohomology
1316: classes. 
1317: \end{proof}
1318: 
1319: \begin{lemma}
1320: Let $u,\alpha : C^{a+b} \longrightarrow J(C)$ defined by:
1321: $$u(x_1 ,\ldots ,x_{a+b}) = \OO_C(x_1 + \ldots + x_{a+b} - (a+b)p_0)$$
1322: and 
1323: $$\alpha(x_1 ,\ldots , x_{a+b}) = \OO_C(x_1 + \ldots + x_a - x_{a+1} - \ldots -
1324: x_{a+b} - (a-b)p_0).$$
1325: Then $u_{*}[C^{a+b}] = \alpha_{*}[C^{a+b}]\in H^{2(g-a-b)}(J(C),
1326: \mathbb{Z})$.
1327: \end{lemma}
1328: \begin{proof}
1329: For simplicity, in this proof only, we will use additive divisor notation,
1330: although we actually mean the associated line bundles.
1331: Consider the auxiliary maps:
1332: $$u_0: C^{a+b}\longrightarrow J(C)^{a+b}$$
1333: $$(x_1, \ldots , x_{a+b})\rightarrow (x_1 - p_0 ,\ldots , x_{a+b} -
1334: p_0),$$
1335: $$\alpha_0: C^{a+b}\longrightarrow J(C)^{a+b}$$
1336: $$(x_1,\ldots , x_{a+b})\rightarrow (x_1  - p_0, \ldots , x_a - p_0, 
1337: p_0 - x_{a+1}, \ldots , p_0 - x_{a+b})$$
1338: and the addition map:
1339: $$a: J(C)^{a+b}\longrightarrow J(C)$$
1340: $$(\xi_1, \ldots , \xi_{a+b})\rightarrow \xi_1 + \ldots + \xi_{a+b}.$$
1341: Then one has:
1342: $$u = au_0 ~{\rm and}~\alpha = a\alpha_0 = a\mu u_0,$$
1343: where $\mu$ is the isomorphism
1344: $$\mu: J(C)^{a+b}\longrightarrow J(C)^{a+b}$$
1345: $$(\xi_1, \ldots ,\xi_{a+b})\rightarrow (\xi_1,\ldots,\xi_a , -\xi_{a+1},
1346: \ldots , -\xi_{a+b}).$$
1347: Now the statement follows from the more general fact that $\mu$ induces 
1348: an isomorphism on cohomology. This is in turn a simple consequence
1349: of the fact that the involution $x \rightarrow -x$ induces the identity on
1350: $H^{1}(J(C), \mathbb{Z})$.
1351: \end{proof}
1352: 
1353: \begin{remark}\label{hyperelliptic_case}
1354: The equality in Proposition~\ref{1} holds set-theoretically also for
1355: hyperelliptic curves. Indeed, we have  
1356: the inclusion $C_{g-i-1}-C_i\subseteq\Theta_{\wedge^iQ}$,
1357: and we have seen in the proof of
1358: Proposition~\ref{hyperelliptic} that $\Theta_{\wedge^iQ}$ is irreducible
1359: (with the reduced structure, it is just a translate of the usual theta
1360:  divisor).
1361: \end{remark}
1362: 
1363: \subsection*{\bf General canonical curves.}
1364: Since we have seen that MRC holds for hyperelliptic curves, a standard
1365: argument shows that it holds for general canonical curves. In fact, our
1366: previous result about the expected
1367: form of the theta divisors $\Theta_{\wedge^iQ}$ allows us
1368: to say something more precise about the set of curves in ${\mathcal M}_g$
1369: which might not satisfy MRC.
1370: 
1371: \begin{proposition}\label{generic}
1372: For every $i$, the set of curves
1373: $\{[C]\in {\mathcal M}_g\mid\Theta_{\wedge^iQ_C}=\Pic^{g-2i-1}(C)\}$
1374: is either empty or has pure codimension one. In particular,
1375: the same assertion is true for the set of curves
1376:  in ${\mathcal M}_g$
1377: which do not satisfy MRC.
1378: \end{proposition}
1379: 
1380: \begin{proof}
1381: As the arguments involved are standard we will just sketch the proof.
1382: 
1383: Start by considering, for a given $d\geq 2g+1$,   
1384: the Hilbert scheme $\mathcal H$ of curves in $\PP^{d-g}$ with  
1385: Hilbert polynomial $P(T)=d\,T+1-g$ and ${\mathcal U}\subseteq\mathcal{H}$
1386: the open subset corresponding to smooth connected nondegenerate curves.
1387:  
1388: Let $f\,:\,{\mathcal Z}\longrightarrow {\mathcal U}$
1389:  be the universal family over $\mathcal U$,
1390: which is smooth of relative dimension $1$,
1391:  and $\omega_{{\mathcal Z}/{\mathcal U}}
1392: \in\Pic({\mathcal Z})$ the relative cotangent bundle.
1393: By base change there is an exact sequence
1394: $$0\longrightarrow {\mathcal Q}^{\vee}
1395: \longrightarrow f^*f_*\omega_{{\mathcal Z}/
1396: {\mathcal U}}
1397: \longrightarrow 
1398: \omega_{{\mathcal Z}/{\mathcal U}}\longrightarrow 0,$$
1399: where ${\mathcal Q}$ is a vector bundle on $\cZ$ such that 
1400:  if $u\in {\mathcal U}$ corresponds to a curve
1401:  $C=\cZ_u$ (in a suitable embedding), then ${\mathcal Q}\vert_{{\mathcal Z}_u}
1402: \iso Q_C$.
1403: 
1404: The usual deformation theory arguments show that ${\mathcal H}$ is smooth
1405: and has dimension $(d-g+1)^2+4(g-1)$. Moreover, the universal family
1406: ${\mathcal Z}$ defines a surjective
1407: morphism $\pi\,:\,{\mathcal U}\longrightarrow
1408: \cM_g$ whose fibers are irreducible and have dimension $(d-g+1)^2+g-1$.
1409: It is immediate to see from this that $\cU$ is connected.
1410: 
1411: Fix now $l$ such that $d=(g-2i-1)+l(2g-2)\geq 2g+1$. Consider
1412: $\cU$ and $\cZ$ as above and let $\cF:=\wedge^i\cQ\otimes\omega_{\cZ/\cU}^{-l}
1413: \otimes p^*\cO_{\PP^{d-g}}(1)$, where $p$ is the composition of the
1414: inclusion $\cZ\hookrightarrow\cU\times\PP^{d-g}$ and the projection onto the
1415: second factor.
1416: 
1417: We consider also the closed subset of $\cU$:
1418: $${\mathcal D}_1=\{u\in\cU\mid h^0(\cF\vert_{\cZ_u})\geq 1\}.$$
1419: It is clear by definition that $\pi^{-1}([C])\subseteq{\mathcal D}_1$
1420: if and only if $\Theta_{\wedge^iQ_C}=\Pic^{g-2i-1}(C)$.
1421:  In particular,
1422:  Proposition~\ref{hyperelliptic} implies that ${\mathcal D}_1\neq \cU$.
1423: 
1424: ${\mathcal D}_1$ is the degeneracy locus of a morphism between two vector
1425:  bundles of the same rank. Indeed, if $H\subset\PP^{d-g}$ is a hyperplane,
1426: $\widetilde{H}=p^{-1}H$,
1427: and $r\gg 0$, then ${\mathcal D}_1$ is the degeneracy locus of 
1428: $$f_*(\cF\otimes\cO_{\mathcal Z}(r\widetilde{H}))
1429: \longrightarrow f_*(\cF\otimes \cO_{r\widetilde{H}}(r\widetilde{H})).$$
1430: Note that these are both vector bundles of rank $rd{{g-1}\choose i}$
1431: (we use base change and the fact that by Corollary~3.5 in 
1432: \cite{paranjape}, for every
1433: smooth curve $C$, the bundle $Q_C$ is semistable). We therefore conclude that
1434: ${\mathcal D}_1$ is a divisor on $\cU$.
1435: 
1436: 
1437: On the other hand, it is easy to see that the set
1438: $${\mathcal D}_2=\{u\in\cU\mid \cO_{\cZ_u}(1)\otimes\omega_{\cZ_u}^{-l}
1439: \in (\cZ_u)_{g-i-1}-(\cZ_u)_i\}$$ is closed.
1440: Moreover, Proposition~\ref{1}
1441: (see also Remark~\ref{hyperelliptic_case}) shows that
1442: ${\mathcal D}_2\subseteq {\mathcal D}_1$, 
1443: and if $\pi^{-1}([C])\not\subseteq{\mathcal D}_1$,
1444: then $\pi^{-1}([C])\cap {\mathcal D}_1=\pi^{-1}([C])\cap {\mathcal D}_2$.
1445: 
1446: Let $\mathcal S$ be the set of irreducible components of ${\mathcal D}_1$
1447: which are not included in ${\mathcal D}_2$. Using the fact that $\pi$
1448: has irreducible fibers, all of the same dimension, it is easy to see that
1449: if $Y\in{\mathcal S}$, then $\pi(Y)$ is closed in $\cM_g$,
1450: that it is in fact a divisor, and $Y=\pi^{-1}(\pi(Y))$. Moreover, the locus of
1451: curves in $\cM_g$ for which $\Theta_{\wedge^iQ}$ is not a divisor is
1452: $\cup_{Y\in{\mathcal S}}
1453: \pi(Y)$, which proves the proposition.
1454: \end{proof}
1455: 
1456: 
1457: 
1458: 
1459: \subsection*{\bf The class of the degeneracy locus on $\cM_{g,g+1}$}
1460: 
1461: We first fix the notation. We will denote by ${\mathcal M}_g^{0}$ the open
1462: subset of ${\mathcal M}_g$ which corresponds to curves with no
1463: nontrivial automorphisms.
1464: From now on we assume that $g\geq 4$, since for $g=3$ MRC is equivalent
1465: to IGC, which is the content of Proposition \ref{igc_canonical}.
1466: Thus $\cM_g^{0}$ is 
1467: nonempty and its complement has codimension 
1468: $g-2\geq 2$ (see \cite{HaMo} pg.37), so working with this subset 
1469: will not affect the answers we get for divisor class computations
1470: on $\cM_g$ or $\cM_{g,n}$.
1471: 
1472: In this case we have a universal family over $\cM_g^{0}$ denoted by
1473: ${\mathcal C}_g^{0}$
1474:  and for every $n\geq 1$, the open subset of
1475: $\cM_{g,n}$ lying over $\cM_g^{0}$
1476: (which we denote by $\cM_{g,n}^{0}$) is equal to the fiber product
1477: $(\times_{\cM_g^{0}}{\mathcal C}_g^{0})^n$
1478:  minus all the diagonals. We assume that $n\geq g+1$.
1479: 
1480: Consider the following cartezian diagram:
1481: \[
1482: \begin{CD}
1483: {\mathcal X}@>{q}>> \cM_{g,n}^{0}\\
1484: @VV{f}V@VV{h}V\\
1485: {\mathcal C}_g^{0}@>{p}>>\cM_g^{0}\\
1486: \end{CD}
1487: \]
1488: in which all the morphisms 
1489: and varieties are smooth and $p$ (hence also $q$) is proper.
1490: 
1491: Let $\omega\in\Pic({\mathcal C}_g^{0})$ be the relative canonical line
1492:  bundle for $p$, $E=p_*(\omega)$ the Hodge vector bundle and
1493:  $\cQ$ the rank $g-1$
1494: vector bundle on ${\mathcal C}^{0}_g$ such that $\cQ^{\vee}$ is the
1495: kernel of the evaluation map $p^*E\longrightarrow\omega$. For every
1496: $[C]\in\cM_g^{0}$, we have $\cQ\vert_{p^{-1}([C])}\simeq Q_C$.
1497: 
1498: The projection on the $j$-th factor $p_j\,:\,\cM_{g,n}^{0}
1499: \longrightarrow{\mathcal C}_g^{0}$ induces a section 
1500: $q_j\,:\,\cM_{g,n}^{0}\longrightarrow {\mathcal X}$ of $q$.
1501: If $E_j={\rm Im}(q_j)$, then $E_j$ is a relative divisor over $\cM_{g,n}
1502: ^{0}$. Consider the following vector bundle on ${\mathcal X}$:
1503: $$\cE=\wedge^if^*\cQ\left(\sum_{j=1}^{g-i}E_j-\sum_{j=g-i+1}^{g+1}E_j\right)$$
1504: and let
1505:  $\cZ=\{u\in\cM_{g,n}^{0}\vert ~h^0(\cE\vert_{{\mathcal X}_u})\geq 1\}$.
1506: 
1507: The algebraic set
1508: $\cZ$ comes equipped with a natural stucture of degeneracy locus. 
1509: Suppose for example that $Y$ is a sum of $m$ divisors $E_j$
1510: (possibly with repetitions), where $m\gg 0$. In this case, $\cZ$
1511: is the degeneracy locus of the morphism
1512:  $$\cF:=q_*(\cE\otimes\cO_{\mathcal X}(Y))
1513: \longrightarrow \cF':=q_*(\cE\otimes\cO_{\mathcal X}(Y)\vert_Y).$$
1514:  This scheme 
1515: structure does not depend on the divisor $Y$ we have chosen.
1516: In fact, it is the universal subscheme over which the $0$-th Fitting ideal
1517: of the first higher direct image of $\cE$ becomes trivial (see e.g.
1518: \cite{ACGH} Ch.IV \S3
1519:  for the proof of an analogous property). Note that $\cZ$ is a divisor
1520: and not the whole space, since by Proposition~\ref{generic} we know that
1521: for a general curve $C$, there is $\xi\in\Pic^{g-2i-1}(C)$
1522: such that $H^0(\wedge^iQ_C\otimes\xi)=0$ (note also 
1523:  that the difference map
1524: $C^{g-i}\times C^{i+1}\longrightarrow \Pic^{g-2i-1}(C)$ is surjective,
1525: cf. \cite{ACGH} Ch.V.D).
1526: 
1527: 
1528: We will use the notation $\lambda=c_1(h^*(E))$, $\psi_j=c_1(p_j^{*}(\omega))$,
1529: and $\Psi_x=\sum_{j=1}^{g-i}\psi_j$ and $\Psi_y=\sum_{j=g-i+1}^{g+1}\psi_j$.
1530: It is well known that $\lambda$ together with $\psi_j$, $1\leq j\leq n$,
1531: form a basis for $\Pic(\cM_{g,n}^{0})_{\QQ}$.
1532: 
1533: \begin{proposition}\label{degeneracy_class}
1534: With the above notation, for every $n\geq g+1$, the class of $\cZ$
1535: in $\Pic(\cM_{g,n}^{0})_{\QQ}$ is given by
1536: \begin{equation}
1537: [\cZ]=-\left({{g-1}\choose i}-10{{g-3}\choose{i-1}}\right)
1538: \lambda+{{g-2}\choose i}\Psi_x
1539: +{{g-2}\choose{i-1}}\Psi_y.
1540: \end{equation}
1541: \end{proposition}
1542: 
1543: 
1544: \begin{proof}
1545: Note that the pull-back of divisors induced by the projection to
1546:  the first $(g+1)$ components induces injective homomorphisms
1547: $\Pic(\cM_{g,g+1}^{0})_{\QQ}\hookrightarrow
1548: \Pic(\cM_{g,n}^{0})_{\QQ}$. From the universality of the scheme
1549: structure on $\cZ$ it follows that
1550:  the computation of $c_1(\cZ)$ is
1551: independent of $n$. Therefore we may assume
1552: that $n$ is large enough, so that in defining the
1553: scheme structure of $\cZ$ as above, we may take $Y=\sum_{j=g+2}^nE_j$.
1554: We introduce also the notation $\Psi_z=\sum_{j=g+2}^n\psi_j$.
1555: 
1556: As a degeneracy locus, the class of $\cZ$ is given by $c_1(\cF')-c_1(\cF)$.
1557: It is clear that we have $E_j\cap E_l=\emptyset$ if $j\neq l$
1558: and via $E_j\simeq\cM_{g,n}^{0}$, we have $\cO_{E_j}(-E_j)\simeq
1559: p_j^*(\omega)$. Since $\cQ^{\vee}$ is the kernel of the evaluation map for
1560: $\omega$, we get 
1561: \begin{equation}\label{first_formula}
1562: c_1(f^*Q)=f^*(c_1(\omega))-q^*(\lambda).
1563: \end{equation}
1564: 
1565: 
1566: Before starting the computation of $c_1(\cF)$ and $c_1(\cF')$, we record the
1567: following well-known formulas for Chern classes.
1568: 
1569:  \begin{lemma}\label{chern_classes}
1570: Let $R$ be a vector bundle of rank $n$ on a variety $X$
1571: and $L\in {\rm Pic}(X)$. 
1572: \item{\rm (i)} $c_1(R\otimes L)=c_1(R)+nc_1(L)$.
1573: \item{\rm (ii)} $c_2(R\otimes L)=
1574: c_2(R)+(n-1)c_1(R)c_1(L)+{n\choose 2}c_1(L)^2$.
1575: \item{\rm (iii)} $c_1(\wedge^iR)=
1576: {{n-1}\choose{i-1}}c_1(R)$, if $n\geq 2$ and $1\leq i
1577: \leq n$.
1578: \item{\rm (iv)} $c_2(\wedge^iR)=\frac12
1579: {{n-1}\choose{i-1}}\left({{n-1}\choose {i-1}}-1\right)
1580: c_1(R)^2+{{n-2}\choose{i-1}}c_2(R)$, if $n\geq 3$ and $1\leq i\leq n$.
1581: \end{lemma}
1582: 
1583: {}From the previous discussion and Lemma~\ref{chern_classes} (i)
1584:  and (iii), we get
1585: $$c_1(q_*(\cE\otimes\cO_{E_j}(Y)))=-{{g-2}\choose{i-1}}\lambda-
1586: {{g-2}\choose i}\psi_j,$$
1587: {}from which we deduce
1588: $$c_1(\cF')=-(n-g-1){{g-2}\choose{i-1}}\lambda-{{g-2}\choose i}\Psi_z.$$
1589: 
1590: In order to compute $c_1(\cF)$, we apply the Grothendieck-Riemann-Roch
1591: formula for $q$ and $\cE(Y)$
1592: (see \cite{fulton} 15.2).
1593:  Note that the varieties are smooth and $q$ is smooth and proper.
1594: Moreover, since we assume $n\gg 0$, we have $R^jq_*(\cE(Y))=0$, for $j\geq 1$.
1595: Therefore we get
1596: $$ch(q_*(\cE(Y)))=q_*(ch(\cE(Y))\cdot td(f^*\omega^{-1})).$$
1597: From this we deduce
1598: $$(*)\,
1599: c_1(\cF)=q_*\left(\frac12 c_1(\cE(Y))^2-c_2(\cE(Y))-
1600: \frac12 f^*c_1(\omega)\cdot c_1(\cE(Y))
1601: +\frac1{12}{{g-1}\choose i}f^*c_1(\omega)^2\right). 
1602: $$
1603: 
1604: We compute now each of the classes involved in the above equation.
1605: In order to do this we need to know how to make the push-forward of the
1606: elementary classes on $\mathcal X$. We list these rules in the following: 
1607: 
1608: \begin{lemma}\label{rules}
1609: With the above notation, we have
1610: \item {\rm (i)} $q_*(f^*c_1(\omega)^2)=12\lambda$.
1611: \item {\rm (ii)} $q_*(q^*\lambda\cdot f^*c_1(\omega))=(2g-2)\lambda$.
1612: \item {\rm (iii)} $q_*q^*(\lambda^2)=0$.
1613: \item {\rm (iv)} $q_*(c_1(E_j)\cdot q^*\lambda)=\lambda$.
1614: \item {\rm (v)} $q_*(c_1(E_j)\cdot f^*c_1(\lambda))=\psi_j$.
1615: \item {\rm (vi)} $q_*q^*c_2(h^*E)=0$.
1616: \item {\rm (vii)} $q_*(c_1(E_j)^2)=-\psi_j$.
1617: \end{lemma}
1618: 
1619: \begin{proof}[Proof of Lemma~\ref{rules}]
1620: The proof of (i) is analogous to that of the relation $p_*(c_1(\omega)^2)=
1621: 12\lambda$ (see \cite{HaMo} \S 3E). The other formulas are straightforward.
1622: \end{proof}
1623: 
1624: Using Lemma~\ref{chern_classes} (i) and (iii) and the formula 
1625: (\ref{first_formula}) for
1626: $c_1(f^*Q)$, we deduce that 
1627: $$c_1(\cE(Y))={{g-2}\choose{i-1}}(f^*c_1(\omega)-q^*\lambda)
1628: +{{g-1}\choose i}\left(\sum_{j=1}^{g-i}c_1(E_j)-\sum_{j=g-i+1}^{g+1}c_1(E_j)
1629: +\sum_{j=g+2}^nc_1(E_j)\right).$$
1630: Applying Lemma~\ref{rules}, we get
1631: $$q_*(c_1(\cE(Y))^2/2)=\left((8-2g){{g-2}\choose{i-1}}^2
1632: -(n-2i-2){{g-2}\choose{i-1}}
1633: {{g-1}\choose i}\right)\lambda+$$
1634: $${{g-2}\choose{i-1}}{{g-1}\choose i}(\Psi_x-\Psi_y+\Psi_z)
1635: -\frac12{{g-1}\choose i}^2(\Psi_x+\Psi_y+\Psi_z).$$
1636: 
1637: 
1638: {}From the above formula for $c_1(\cE(Y))$ and Lemma~\ref{rules}, we get
1639: $$q_*\left(-\frac12 f^*c_1(\omega)\cdot c_1(\cE(Y))\right)
1640: =(g-7){{g-2}\choose{i-1}}\lambda
1641: -\frac12{{g-1}\choose i}(\Psi_x-\Psi_y+\Psi_z).$$
1642: 
1643: 
1644: Lemma~\ref{rules} (i) gives
1645: $$q_*\left(\frac1{12}{{g-1}\choose i}f^*c_1(\omega)^2\right)
1646: ={{g-1}\choose i}\lambda.$$
1647: 
1648: {}From the defining exact sequence of $\cQ^{\vee}$ we compute
1649: $$c_2(f^*\cQ)=q^*c_2(h^*E)+f^*c_1(\omega)\cdot(f^*c_1(\omega)-q^*\lambda).$$
1650: Using now Lemma~\ref{chern_classes} (ii) and (iv) and Lemma~\ref{rules},
1651:  we deduce
1652: $$q_*c_2(\cE(Y))= \left((8-2g){{g-2}\choose{i-1}}\left(
1653: {{g-2}\choose{i-1}}-1\right)+(14-2g){{g-3}\choose{i-1}}\right)\lambda-$$
1654: $$(n-2i-2){{g-2}\choose{i-1}}\left({{g-1}\choose i}-1\right)\lambda+
1655: {{g-2}\choose{i-1}}\left({{g-1}\choose i}-1\right)(\Psi_x-\Psi_y+\Psi_z)-$$
1656: $$\frac12{{g-1}\choose i}\left({{g-1}\choose i}-1\right)(\Psi_x
1657: +\Psi_y+\Psi_z).$$
1658: 
1659: 
1660: 
1661: Using these formulas and equation $(*)$, we finally obtain
1662: $$c_1(\cF)=\left((2g-14){{g-3}\choose{i-1}}-(n+g-2i-3){{g-2}\choose{i-1}}
1663: +{{g-1}\choose i}\right)\lambda-$$
1664: $${{g-2}\choose i}(\Psi_x+\Psi_z)
1665: -{{g-2}\choose{i-1}}\Psi_y.$$
1666: 
1667: Since the class of $\cZ$ is equal with $c_1(\cF')-c_1(\cF)$, we deduce
1668: the statement of the proposition.
1669: \end{proof}
1670: 
1671: \subsection*{\bf Proof of the main result.}
1672: We introduce next a divisor
1673:  $D$ on $\mathcal{M}_{g,g+1}$ which is a global analogue of the
1674:  preimage of $C_{g-i-1}-C_{i}$ under the difference map
1675:  $C^{g-i}\times C^{i+1}\rightarrow \mbox{Pic}^{g-2i-1}(C)$.
1676:  This is motivated by Proposition \ref{1}, and our goal is roughly 
1677:  speaking to prove a global version of that result. 
1678: 
1679: \begin{definition} For $g\geq 3$ and
1680:  $1\leq i\leq \frac{g-1}{2}$ we define the divisor $D$
1681: on $\mathcal{M}_{g,g+1}$ to be the locus of smooth pointed curves
1682:  $(C, x_1,\ldots ,x_{g-i}, y_1,\ldots y_{i+1})$
1683:  having a linear series $\mathfrak g^1_{g}$ containing 
1684: $x_1+\cdots +x_{g-i}$ in a fiber and $y_1+\cdots +y_{i+1}$ in another fiber.
1685: Note that this means that we can in fact write the line bundle
1686: $\OO_C(x_1+\ldots +x_{g-i}-y_1-\ldots-y_{i+1})$ 
1687: as an element in $C_{g-i-1}-C_i$.
1688: \end{definition}
1689: 
1690: We consider in what follows the divisor $\overline{\mathcal{Z}}$,
1691: the closure in ${\mathcal M}_{g,g+1}$ of the divisor 
1692: ${\mathcal Z}$ studied above (we take now $n=g+1$,  but
1693: as we mentioned, this does not affect the formula for its class).
1694: In Section 4 we prove that $D$ is reduced and that
1695:  $D\equiv_{\QQ-{lin}} \overline{\mathcal{Z}}$
1696: (cf. Theorem \ref{class_D}).
1697: This being granted we are in a position to complete the proof of Theorem 3.1:
1698: 
1699: \begin{proof}[Proof of Theorem~\ref{canonical_curves}] 
1700: Note that for $g=3$ our assertion is just the statement of Proposition 
1701: \ref{igc_canonical}. Thus we can assume $g\geq 4$.
1702: As mentioned above $D$ is reduced, and from Proposition \ref{1}
1703: we see that
1704:  $\mbox{supp}(D)\subseteq \mbox{supp}(\overline{\mathcal{Z}})$.
1705:  We get that $\overline{\mathcal{Z}}-D$ is effective,
1706:  and in fact $\overline{\mathcal{Z}}-D=h^*(E)$, where $E$ is an effective
1707:  divisor on $\mathcal{M}_g$ and $h:\mathcal{M}_{g,g+1}\rightarrow
1708:  \mathcal{M}_g$ is the projection. 
1709: Moreover, the map $h^*:\mbox{Pic}(\mathcal{M}_g)_{\QQ}\rightarrow
1710:  \mbox{Pic}(\mathcal{M}_{g,g+1})_{\QQ}$ is injective (cf. \cite{AC2}),
1711:  hence $E\equiv_{\QQ-lin}0$. Since 
1712: the Satake compactification of $\mathcal{M}_g$ has boundary
1713:  of codimension $2$ (see e.g \cite{HaMo}, pg.45) this implies $E=0$,
1714:  that is, $\overline{\mathcal{Z}}=D$.
1715:  Therefore $\Theta_{\wedge ^i Q_C}$ is a divisor in
1716:  $\mbox{Pic}^{g-2i-1}(C)$ and the identification
1717:  $\Theta_{\wedge ^i Q_C}=C_{g-i-1}-C_{i}$ holds for
1718:  \emph{every} nonhyperelliptic curve $C$.
1719: \end{proof}
1720: 
1721: 
1722: 
1723: \section{\bf A divisor class computation on $\Mg1$}
1724: 
1725: 
1726: In this section we compute the class of the divisor
1727:  $D$ on $\mathcal{M}_{g,g+1}$ defined in the previous section.
1728: We start by recalling a few facts about line bundles on
1729:  $\overline{\mathcal{M}}_{g,n}$. Let us fix $g\geq 3$, $n\geq 0$
1730:  and a set $N$ of $n$ elements.
1731:  Following \cite{AC2}, we identify $\overline{\mathcal{M}}_{g,n}$
1732:  with the moduli space
1733:  $\overline{\mathcal{M}}_{g,N}$
1734:  of stable curves of genus $g$ with marked points indexed by $N$.
1735:  We denote by $\pi_{q}:\overline{\mathcal{M}}_{g,N\cup\{q\}}\rightarrow
1736:  \overline{\mathcal{M}}_{g,N}$ the map forgetting the marked
1737:  point indexed by $q$.
1738:  For each $z\in N$ we define the tautological class $\psi_z=c_1(\mathbb L_z)\in \mbox{Pic}(\overline{\mathcal{M}}_{g,N})_{\mathbb Q}$,
1739: where $\mathbb L_z$ is the line bundle over $\overline{\mathcal{M}}_{g,N}$ whose fiber over the moduli point $[C,\{x_i\}_{i\in N}]$ is the cotangent space $T_{x_z}^*(C)$. Note that although we are
1740:  using an apparently different definition, these $\psi$ classes are the same 
1741:  as those which appear in the previous section.
1742: 
1743:  For $0\leq i\leq g$ and $S\subseteq N$, the boundary divisor
1744:  $\Delta_{i:S}$  corresponds to the closure in $\overline{\mathcal{M}}_{g,N}$
1745: of the locus of nodal curves
1746:  $C_1\cup C_2$, with $C_1$ smooth of genus $i$, $C_2$ smooth
1747:  of genus $g-i$, and such that the marked points sitting on
1748:  $C_1$ are precisely those labelled by $S$.
1749:  Of course $\Delta_{i:S}=\Delta_{g-i:S^{c}}$ and we set
1750:  $\Delta_{0:S}:=0$ when $|S|\leq 1$. We also consider the divisor $\Delta_{irr}$ consisting of irreducible pointed curves
1751: with one node. We denote by $\delta_{i:S}\in \mbox{Pic}(\overline{\mathcal{M}}_{g,n})_{\QQ}$ the class of $\Delta_{i:S}$ and by $\delta_{irr}$ that of $\Delta_{irr}$. 
1752:  It is well known that the Hodge class $\lambda, \delta_{irr}$, the $\psi_z$'s and the
1753:  $\delta_{i:S}$'s freely generate
1754:  $\mbox{Pic}(\overline{\mathcal{M}}_{g,n})_{\QQ}$ (cf. \cite{AC2}). 
1755: 
1756:  For a smooth curve $C$ and for a pencil $\mathfrak g^1_d$ on $C$, we say that an effective divisor $E$ on $C$ is in a fiber of the pencil if there exists 
1757: $E'\in \mathfrak g^1_d$ such that $E'-E$ is an effective divisor.
1758: 
1759: Recall that for $g\geq 3$ and
1760:  $0\leq i\leq \frac{g-1}{2}$ we have defined the divisor $D$
1761: on $\mathcal{M}_{g,g+1}$ to be the locus of curves
1762:  $(C, x_1,\ldots ,x_{g-i}, y_1,\ldots ,y_{i+1})$
1763:  having a linear series $\mathfrak g^1_{g}$ containing 
1764: $x_1+\cdots +x_{g-i}$ in a fiber and $y_1+\cdots +y_{i+1}$ in another fiber.
1765: We denote by $\overline{D}$ the closure of $D$ in
1766:  $\overline{\mathcal{M}}_{g,g+1}$.
1767: 
1768: The divisor $D$ comes equipped with a scheme structure induced by the forgetful map $\mathcal{G}\rightarrow \mathcal{M}_{g,g+1}$. Here $\mathcal{G}$ is the variety parametrizing objects $[C,\vec{x},\vec{y},l]$, where $\vec{x}=(x_1,\ldots,x_{g-i})$ and $\vec{y}=(y_1,\ldots, y_{i+1})$ are such that $[C,\vec{x},\vec{y}]\in \mathcal{M}_{g,g+1}$ and $l$ is a linear series $\mathfrak g^1_g$ on $C$ such that $l(-\sum_{j=1}^{g-i} x_j)\neq \emptyset$ and
1769: $l(-\sum_{j=1}^{i+1} y_j)\neq \emptyset$.
1770: It is well-known that $\mathcal{G}$ is smooth of pure dimension $4g-3$ (see e.g. \cite{AC1}, pg. 346). Note also that there is a natural action of $S_{g-i}\times S_{i+1}$  on $\mathcal{M}_{g,g+1}$ (and hence on $\mathcal{G}$) by permuting the components of $\vec{x}$ and $\vec{y}$ separately.
1771: 
1772: The main result of the section is the following:
1773: 
1774: \begin{theorem}\label{class_D}
1775: The divisor $D$ is reduced and its class in $\Pic(\overline{\mathcal{M}}_
1776: {g,g+1})_{\QQ}$ is
1777: $$[D]=-\Bigl({g-1\choose  i}
1778: -10{g-3\choose i-1}\Bigr)\ \lambda+{g-2\choose i}\Psi
1779: _x+{g-2\choose i-1}\Psi_y, $$
1780: where $\Psi_x=\sum_{j=0}^{g-i} \psi_{x_j}$
1781:  and $\Psi_y=\sum_{j=0}^{i+1} \psi_{y_j}$.
1782: \end{theorem}
1783: 
1784: \noindent
1785: We begin by proving the first part of Theorem~\ref{class_D}:
1786: 
1787: \begin{proposition}\label{reduced}
1788: The divisor $D$ is reduced.
1789: \end{proposition}
1790: 
1791:  \begin{proof}
1792:  Since the variety $\mathcal{G}$ introduced above is smooth, it suffices to show that the projection $\pi:\mathcal{G}\rightarrow
1793:  \mathcal{M}_{g,g+1}$ given by $\pi([C,\vec{x},\vec{y},l])=[C,\vec{x},\vec{y}]$,
1794: is generically injective. 
1795: 
1796: We pick a component $\mathcal{X}$ of $\mathcal{G}\times _{\mathcal{M}_{g,g+1}}\mathcal{G}$ whose general point corresponds to a marked curve $[C,\vec{x},\vec{y}]\in \mathcal{M}_{g,g+1}$ together with two {\emph{different}} base-point-free $\mathfrak g^1_g$'s on $C$, both containing $\vec{x}$ and $\vec{y}$ in different fibers. Clearly $\mbox{dim}(\mathcal{X})\geq 4g-4$ and if we show that $\mbox{dim}(\mathcal{X})\leq 4g-4$, then we are done.
1797: For a general point in $\mathcal{X}$ we denote by $f_1,f_2:C\rightarrow \PP^1$ the induced $g$-sheeted maps. We may assume that $f_1(\vec{x})=f_2(\vec{x})=0$ and $f_1(\vec{y})=f_2(\vec{y})=\infty.$
1798: The product map $f=(f_1,f_2):C
1799: \rightarrow \PP^1\times \PP^1$ is birational onto its image
1800:  and $\Gamma=f(C)$ will have points of multiplicity
1801:  at least $g-i$ and $i+1$ at $a=(0,0)$ and $b=(\infty,\infty)$ respectively.
1802: 
1803:  If $S=\mbox{Bl}_{\{a,b\}}(\PP^1\times \PP^1)$ we set $\gamma= gl+gm-(g-i)E_a-(i+1)E_
1804: b\in \mbox{Pic(S)}$, where $l$ and $m$ are pullbacks of
1805:  the rulings on $\PP^1\times \PP^1$,
1806:  and $E_a$, $E_b$ are the exceptional divisors. We denote by $V(S,\gamma)$ 
1807: the Severi variety of curves $Y\subset S$ homologous to $\gamma$. 
1808: The discussion above shows that $\mathcal{X}$ lies in the closure of the image of the rational map $V(S,\gamma)- ->(\mathcal{G}\times_{\mathcal{M}_{g,g+1}}\mathcal{G})/S_{g-i}\times S_{i+1}$
1809: obtained by projecting $S$ onto the two factors. Thus $\mbox{dim}(\mathcal{X})\leq \mbox{dim }V(S,\gamma)-\mbox{dim} \mbox{ Aut}(S).$
1810: 
1811: 
1812: On the other hand an argument identical to that in \cite{AC1}, Proposition 2.4, 
1813: shows that since $S$ is a regular surface, every irreducible component $M$ of $V(S,\gamma)$ 
1814: having dimension $\geq g+1$ is of the expected dimension provided by deformation theory, 
1815: that is, 
1816: $\mbox{dim}(M)=g-1-\gamma \cdot K_S$. 
1817: Therefore $\mbox{dim}(\mathcal{X})\leq g-1-\gamma \cdot K_S-
1818: \mbox{dim }\mbox{Aut}(S)=4g-4$.
1819: \end{proof}
1820: 
1821: 
1822: We will prove the second part of Theorem~\ref{class_D} using degeneration
1823: techniques and enumerative geometry.
1824:  
1825: 
1826: \subsection*{Recap on limit linear series (cf. \cite{EH1})}
1827: We recall that for a smooth curve $C$,
1828:  a point $p\in C$ and a linear series $l=(L,V)$
1829:  with $L\in \mbox{Pic}^d(C)$ and $V\in G(r+1, H^0(L))$,
1830:  the \emph {vanishing sequence} of $l$ at $p$
1831:  is obtained by ordering the set $\{\mbox{ord}_p(\sigma)\}_{\sigma \in
1832:  V}$, and it is denoted by 
1833: $$a^l(p):0\leq a_0^l(p)<\ldots <a_r^l(p)\leq d.$$
1834: The \emph{weight} of $l$ at $p$ is defined as
1835:  $w^l(p):=\sum_{i=0}^r (a_i^l(p)-i)$. 
1836: 
1837:  Given a curve $C$ of compact type, a \emph{limit}
1838:  $\mathfrak g^r_d$ on $C$ is a collection of honest
1839:  linear series $l_Y=(L_Y,V_Y)\in G^r_d(Y)$ for each
1840:  component $Y$ of $C$, satisfying the compatibility
1841:  condition that if $Y$ and $Z$ are components of $C$ 
1842: meeting at $p$ then
1843: $$a_i^{l_Y}(p)+a_{r-i}^{l_Z}(p)\geq d\mbox{ }\mbox{ for }i=0,\ldots ,r.$$
1844: We note that limit linear series appear as limits of
1845:  ordinary linear series in $1$-dimensional families of
1846:  curves and there is a useful sufficient criterion
1847:  for a limit $\mathfrak g^r_d$ to be
1848:  \emph{smoothable} (cf. \cite{EH1}, Theorem 3.4).
1849: 
1850: \smallskip
1851: \noindent
1852: We will need the following enumerative result (cf. \cite{H}, Theorem 2.1):
1853: 
1854: \begin{proposition}\label{enumerative}
1855: Let $C$ be a general curve of genus $g$,
1856:  $d\geq \frac{g+2}{2}$ and $p\in C$ a general point.
1857: \begin{itemize}
1858: \item The number of $\mathfrak g^1_d$'s on $C$ containing
1859:  $(2d-g)q$ in a fiber, where $q\in C$ is an unspecified point, is
1860: $$b(d,g)=(2d-g-1)(2d-g)(2d-g+1)\frac{g!}{d!(g-d)!}\mbox{ }.$$
1861: \item If $\beta \geq 1$, $\gamma \geq 1$ are integers
1862:  such that $\beta+\gamma=2d-g$, the number of
1863:  $\mathfrak g^1_d$'s on $C$ containing $\beta p+\gamma q$
1864:  in a fiber for some point $q \in C$ is
1865: $$c(d,g,\gamma)=(\gamma^2(2d-g)-\gamma)\frac{g!}{d!\ (g-d)!}\ .$$
1866: \end{itemize}
1867: \end{proposition}
1868: 
1869: \noindent
1870: The following simple observation will be used repeatedly:
1871: 
1872: \begin{proposition}\label{glue}
1873: Fix $y,z\in N$ and denote by
1874:  $\pi_z:\overline{\mathcal{M}}_{g,N}\rightarrow
1875:  \overline{\mathcal{M}}_{g,N-\{z\}}$ the map forgetting
1876:  the marked point labelled by $z$. If $E$ is any divisor class
1877:  on $\overline{\mathcal{M}}_{g,N}$, then the $\lambda$ and the $\psi_x
1878: $ coefficients of $E$ are the same as those of
1879:  $(\pi_z)_*(E\cdot \delta_{0:yz})$ for all $x\in \{y,z\}^c.$
1880: \end{proposition}
1881: 
1882: \begin{proof}
1883:  We write $E$ uniquely as a combination of $\lambda$, tautological classes 
1884: $\psi_y,\psi_z$ and $\psi_x$ with $x\in \{y,z\}^c$ and boundary divisors. To express $(\pi_z)_*(E\cdot \delta_{0:yz})$ in $\mbox{Pic}(\overline{\mathcal{M}}_{g,N-\{z\}})_{\mathbb Q}$ we use that
1885: $(\pi_z)_*(\lambda\cdot \delta_{0:yz})=
1886: \lambda,\mbox{ }(\pi_z)_*(\psi_x\cdot \delta_{0:yz})
1887: =\psi_x \mbox{ for }x\in \{y,z\}^c$ and that $(\pi_z)_*(\psi_x\cdot \delta_{0:yz})=0$ for $x\in \{y,z\}.$ Moreover we have that $(\pi_z)_*(\delta_{i:S}\cdot \delta_{0:yz})$ is boundary in all cases
1888: except that $(\pi_z)_*(\delta_{0:yz}^2)=-\psi_y$ (cf. \cite{AC2}, Lemma 1.2
1889:  and \cite{logan}, Theorem 2.3). The conclusion follows immediately. 
1890: \end{proof}
1891: 
1892: 
1893:  By a succession of push-forwards, using Proposition \ref{glue}
1894:  we will reduce the problem of computing the class of $D$
1895:  to two divisor class computations in $\overline{\mathcal{M}}_{g,3}$. The main idea is to let all the points $x_j$ and then all the points $y_j$ come together and understand how the geometric condition defining $D$ changes under degeneration. Recall that by $\overline{D}$ we denote the closure of $D$ in $\overline{\mathcal{M}}_{g,g+1}$.
1896: 
1897:  We define the following sequence of divisors:
1898:  starting with $\overline{D}=D_{y_{i+1}}$, for $1\leq j\leq i$
1899:  we define inductively the divisors $D_{y_j}$ on
1900:  $\overline{\mathcal{M}}_{g,g-i+j}$ by
1901: $$D_{y_j}:=(\pi_{y_{j+1}})_*(\Delta_{0:y_jy_{j+1}}\cdot D_{y_{j+1}}).$$
1902: Loosely speaking, $D_{y_j}$ is obtained from $D_{y_{j+1}}$
1903:  by letting the marked points $y_j$ and $y_{j+1}$ come together.
1904: Then we define $D_{x_{g-i}}:=D_{y_1}$ and we let the
1905:  marked points $x_2,\ldots,x_{g-i}$ come together
1906: : for $2\leq j\leq g-i-1$
1907:  we define inductively the divisors $D_{x_j}$ on
1908:  $\overline{\mathcal{M}}_{g,j+1}$ by
1909: $$D_{x_j}:=(\pi_{x_{j+1}})_*(\Delta_{0:x_j x_{j+1}} \cdot D_{x_{j+1}}).$$
1910: Proposition~\ref{glue}
1911:  ensures that the $\psi_{x_1}$ and the $\lambda$ coefficients
1912: of $[\overline{D}]$ are the same as those of $[D_{x_2}]$.
1913: 
1914: \begin{proposition}\label{reduce}
1915:  The divisor $D_{x_2}$ is reduced and
1916:  it is the closure in $\overline{\mathcal{M}}_{g,3}$
1917:  of the locus of those smooth pointed curves $(C,x_1,x_2,y)$
1918:  for which there exists a $\mathfrak g^1_g$ with
1919:  $(i+1)y$ in a fiber and $x_1+(g-i-1)x_2$ in another fiber.
1920: \end{proposition}
1921: 
1922: \begin{proof}
1923:  For simplicity we will only prove that
1924:  $D_{y_i}$ is reduced and that it is the closure of the locus
1925:  of those smooth pointed curves $(C,x_1,\ldots,x_{g-i},y_1,\ldots,y_i)$
1926:  for which $x
1927: _1+\cdots +x_{g-i}$ and $y_1+\cdots +y_{i-1}+2y_i$ are
1928:  in different fibers of the same $\mathfrak g^1_g$.
1929:  Then by iteration we will get a similar statement for $D_{x_2}$.
1930: 
1931:  Let $(X=C\cup_q \PP^1,x_1,\ldots ,x_{g-i}, y_1,\ldots ,y_{i+1})$
1932:  with $y_i, y_{i+1}\in \PP^1$ be a general point in a
1933:  component of $D_{y_{i+1}}\cap \Delta_{0:y_i y_{i+1}}$. A standard dimension count shows that $C$ must be smooth. There exists a limit $\mathfrak g^1_g$ on $X$,
1934:  say $l=(l_C,l_{\PP^1})$, together with sections
1935:  $\sigma_{\PP^1} \in V_{\PP^1}$ and
1936:  $\sigma_C, \tau_C \in V_C$, such that
1937:  $\mbox{div}(\tau_C)\geq x_1+\cdots +x_{g-i}$,
1938:  $\mbox{div}(\sigma_C)\geq y_1+\cdots +y_{i-1}$,
1939:  $\mbox{div}(\sigma_{\PP^1})\geq y_i+y_{i+1}$ and moreover
1940:  $\mbox{ord}_q(\sigma_{\PP^1})+\mbox{ord}_{q}(\sigma_C)\geq g$
1941:  (apply \cite{EH1}, Proposition 2.2). 
1942: 
1943: Clearly $\mbox{ord}_q(\sigma_{\PP^1})\leq g-2$, hence
1944:  $\mbox{div}(\sigma_C)\geq 2q+y_1+\cdots +y_{i-1}$.
1945:  The contraction map $\pi_{y_{i+1}}$ collapses $\PP^1$
1946:  and identifies $q$ and $y_i$, so the second part of the claim follows.
1947: 
1948:  To conclude that $D_{y_i}$ is also reduced we use
1949:  that both $D_{y_{i+1}}$ and $\Delta_{0:y_i y_{i+1}}$
1950:  are reduced and that they meet transversally.
1951:  This is because the limit $\mathfrak g^1_g$ we found on $X$
1952:  is smoothable in such a way that all ramification is kept
1953:  away from the nodes (cf. \cite{EH1}, Proposition 3.1), hence the tangent spaces to $D_{y_i}$ and $\Delta_{0:y_iy_{i+1}}$ at the intersection point $(X,\vec{x},\vec{y})$ cannot be equal.
1954: \end{proof}
1955: 
1956: \smallskip
1957: 
1958:  In a similar way, by letting first all $x_j$ with
1959:  $1\leq j\leq g-i$ and then all $y_j$ with $2\leq j\leq i+1$
1960:  coalesce, we obtain a reduced divisor $D_{y_2}$ on
1961:  $\overline{\mathcal{M}}_{g,3}$ which is the closure of the locus
1962:  of smooth curves $(C,x,y_1,y_2)$ having a $\mathfrak g^1_g$
1963:  with $(g-i)x$ and $y_1+iy_2$ in different fibers.
1964:  Moreover, the $\lambda$ and the $\psi_{y_1}$ coefficients of
1965:  $[\overline{D}]$ coincide with those of $[D_{y_2}]$.
1966:  Once more applying Proposition~\ref{glue} it follows that the $\lambda$
1967:  and the $\psi_{y_1}$ coefficients of $[D_{y_2}]$ are the same
1968:  as those of 
1969:  $(\pi_x)_*([D_{y_2}]\cdot \delta_{0:xy_2})$. Similarly, 
1970: the $\psi_{x_1}$ coefficient of $[D_{x_2}]$ is the same as that
1971:  of $(\pi_y)_*([D_{x_2}]\cdot \delta_{0:x_2y})$.
1972:  
1973: \begin{proposition}\label{sum_Y}
1974:  We have that 
1975: $$(\pi_x)_*(D_{y_2}\cdot \Delta_{0:xy_2})=\sum_{j=0}^i
1976: Y_j,$$
1977: where for $j<i$ the reduced divisor $Y_j$ is the closure in
1978:  $\overline{\mathcal{M}}_{g,2}$ of the locus of curves
1979:  $(C,y_1,y_2)$ having a $\mathfrak g^1_{g-j}$ with
1980:  $(g-2j-1)y_2+y_1$ in a fiber, while the reduced divisor $Y_i$ consists of curves
1981:  $(C,y_1,y_2)$ with
1982: a $\mathfrak g^1_{g-i}$ having $(g-2i)y_2$ in a fiber
1983:  (and no condition on $y_1$).
1984: \end{proposition}
1985: 
1986: \begin{proof}
1987:  Once again, let $(X=C\cup_q \PP^1,x,y_1,y_2)$
1988:  be a point in $D_{y_2}\cap \Delta_{0:xy_2}$, with $y_1\in C$
1989:  and $x,y_2\in \PP^1$. Then there exists a limit
1990:  $\mathfrak g^1_g$, say $l=(l_C,l_{\PP^1})$ on $X$ together
1991:  with sections $\sigma_{\PP^1},\tau_{\PP^1}\in
1992:  V_{\PP^1}$ and $\sigma_C\in V_C$ such that 
1993: $\mbox{div}(\sigma_{\PP^1})\geq iy_2, \mbox{div}
1994: (\tau_{\PP^1})\geq (g-i)x, \mbox{div}(\sigma_C)\geq y_1$
1995:  and moreover $\mbox{ord}_q(\sigma_{\PP^1})
1996: +\mbox{ord}_q(\sigma_C)\geq g$.
1997: 
1998:  The Hurwitz formula on $\PP^1$ and the condition defining 
1999:  a limit linear series give that
2000:  $w^{l_C}(q)\geq w^{l_{\PP^1}}(x)+w^{l_{\PP^1}}(y_2)\geq g-2$.
2001:  On the other hand, since $(X,x,y_1,y_2)$ moves in a family
2002:  of dimension $\geq 3g-2$ it follows that $(C,q)$ also moves in a
2003:  family of dimension $\geq 3g-3$ in $\mathcal{M}_{g,1}$
2004:  (i.e. codimension $\leq 1$). Since according to \cite{EH2}, Theorem 1.2,
2005:  the locus of pointed curves $[C,q]\in \mathcal{M}_{g,1}$ carrying a $\mathfrak g^1_g$ having $w(q)\geq g$ has codimension $\geq 2$, we get $w^{l_C}(q)\leq g-1$. There are two possibilities:
2006: 
2007:  \textbf{i)} $w^{l_C}(q)=g-2$. Let us denote $j=a_0^{l_C}(q)$,
2008:  hence $a_1^{l_C}(q)=g-1-j$ and $a_k^{l_C}(q)+a_{1-k}^{l_{\PP^1}}(q)=g$
2009:  for $k=0,1$. Therefore
2010:  $j+1=a_0^{l_{\PP^1}}(q)\leq \mbox{ord}_q(\tau_{\PP^1})\leq i.$
2011:  Moreover, since $\mbox{ord}_{q}(\sigma_{\PP^1})
2012: \leq g-i\leq g-j-1$, we obtain that $\mbox{ord}_q(\sigma_C)\geq j+1$,
2013:  hence $\mbox{div}(\sigma_C)\geq y_1+(g-1-j)q$,
2014:  that is, $l_C(-jq)$ is a $\mathfrak g^1_{g-j}$
2015:  on $C$ with $(g-2j-1)q+y_1$ in a fiber, or equivalently $
2016: [C,y_1,q]\in Y_j$, where $0\leq j\leq i-1$.
2017: 
2018: 
2019: To see that conversely $\bigcup_{j=0}^{i-1}Y_j\subseteq
2020:  (\pi_{x})_*(D_{y_2}\cdot \Delta_{0:xy_2})$ we pick a general
2021:  pointed curve $(C,y_1,q)$ having a $\mathfrak g^1_{g-j}$ with
2022:  $(g-2j-1)q+y_1$ in a fiber and we construct a Harris-Mumford admissible
2023:  covering $f:X'\rightarrow B$ of degree $g$,
2024:  where $X'$ is a curve semistably equivalent to $X$ defined as above, and
2025:  $B=(\PP^1)_1\cup_{t}(\PP^1)_2$ is the transversal
2026:  union of two lines (see Fig. 1): we take $f_{|C}:C\rightarrow (\PP^1)_1$
2027:  to be the degree $g-j$ covering 
2028: such that $(g-2j-1)q+y_1\subseteq f_{|C}^*(t)$, while
2029:  $f_{|\PP^1}:\PP^1\rightarrow (\PP^1)_2$
2030:  is the degree $g-j-1$ map containing $(g-i)x$ and $iy_2$
2031:  in different fibers and with $(g-2j-1)q$ in the fiber over $t$.
2032:  It is clear that there
2033:  is a unique such $\mathfrak g^1_{g-j-1}$ on $\PP^1$.
2034:  Furthermore, at $y_1$ we insert a rational curve $R$ mapping
2035:  isomorphically onto $(\PP^1)_2$ and at the remaining $j$ points
2036:  in $f_{|C}^{-1}(t)-\{y_1,q\}$ we insert rational curves mapping 
2037: with degree $1$ onto $(\PP^1)_2$ while at the $g-j$ points
2038:  in $f_{|\PP^1}^{-1}(t)-\{q\}$ we insert
2039:  copies of $\PP^1$ mapping isomorphically onto $(\PP^1)_1$. 
2040: We denote the resulting curve by $X'$. If $y_1'=f_{|R}^{-1}(f(y_2))$, then
2041: $(X',x,y_1',y_2)$ is stably equivalent to
2042:  $(X,x,y_1,y_2)$ and $iy_2+y_1'$ and $(g-i)x$ appear in distinct fibers of the $g$-sheeted map $f:X'\rightarrow B$. Thus we get that $[X,x,y_1,y_2]\in D_{y_2}\cap \Delta_{0:xy_2}$.
2043: 
2044: 
2045: \begin{figure}[ht]
2046: \begin{center}
2047: \mbox{\epsfig{file=cov.eps,width=8cm,height=6cm,angle=0}}
2048: \end{center}
2049: \label{}
2050: \caption{}
2051: \end{figure}
2052: 
2053: 
2054: 
2055: 
2056: 
2057: 
2058: %\begin{figure}[!ht]\label{Boundary Map}
2059:  %       \scalebox{.4}{\includegraphics{covering.eps}}
2060: %\end{figure}        
2061: 
2062: \smallskip
2063: 
2064: \textbf{ii)} $w^{l_C}(q)=g-1$.
2065:  We denote $a^{l_C}(q)=a^{l_{\PP^1}}(q)=(j,g-j)$.
2066:  Since $\mbox{ord}_q(\tau_{\PP^1})\leq i$ we get that
2067:  $j\leq i$. Now $w^{l_C}(q)=g-1$ is already a
2068:  codimension $1$ condition on $\mathcal{M}_{g,1}$,
2069:  so it follows that $\mbox{ord}_q(\sigma_C)=j$,
2070: hence $\mbox{div}(\sigma_{\PP^1})\geq (g-j)q+iy_2$. This yields $i=j$ and
2071: $\mbox{div}(\sigma_{\PP^1})=(g-i)q+iy_2$.
2072:  We thus get that $[C,y_1,q]\in Y_i$. 
2073: 
2074:  Conversely, given $(C,y_1,q)\in \mathcal{M}_{g,2}$ together
2075:  with a $\mathfrak g^1_{g-i}$ on $C$ with $(g-2i)q$ in a fiber,
2076:  we construct a degree $g$ admissible covering
2077:  $f:X'\rightarrow (\PP^1)_1\cup_t(\PP^1)_2$,
2078: which will prove that $[C,q,y_2]\in (\pi_x)_*(D_{y_2}\cdot \Delta_{0:xy_2})$:
2079: we first take $f_{|C}:C\rightarrow (\PP^1)_1$ of degree
2080:  $g-i$ with $(g-2i)q\subseteq f_{|C}^*(t)$.
2081:  Then $f_{|\PP^1}:\PP^1\rightarrow (\PP^1)_2$
2082:  is of degree $g-i$, completely ramified at $x$ and with
2083:  $f_{|\PP^1}^{-1}(t)=(g-2i)q+iy_2$. At $y_2\in \PP^1$ we insert a 
2084: rational curve $R$ which we map $i:1$ to $(\PP^1)_1$
2085:  such that we have total ramification both at $y_2$
2086:  and at the point $y_2'\in R$ characterized by
2087: $f_{|C}(y_1)=f_{|R}(y_2')$. Finally, at each of the points in
2088:  $f_{|C}^{-1}(t)-\{q\}$ we insert a $\PP^1$
2089:  which we map isomorphically onto $(\PP^1)_2$. 
2090: 
2091: Thus we have proved that $\mbox{supp}(\pi_x)_*(D_{y_2}\cdot \Delta_{0:xy_2})=\cup_{j=0}^i \mbox{supp}(Y_j)$. The conclusion now follows if we notice that $D_{y_2}$ is reduced and all admissible coverings we constructed are smoothable, hence $D_{y_2}\cdot \Delta_{0:xy_2}$ is reduced too.
2092: \end{proof}
2093: 
2094: We have thus reduced the problem of computing $[D]$ to that of computing the class of all divisors $Y_j$ on $\mathcal{M}_{g,2}$ for $0\leq j\leq i$.
2095: 
2096: \begin{proposition}\label{class_Y}
2097: For $0\leq j\leq i$ we have the following relations
2098:  in $\rm{Pic}$$(\mathcal{M}_{g,2})_{\mathbb Q}$:
2099: $$Y_j\equiv_{lin}a_j\lambda+b_{1j}\psi_{y_1}+b_{2j}
2100: \psi_{y_2},\mbox{ }\mbox{   where}$$
2101: 
2102: $$a_j=-\frac{g-2j}{g}{g \choose j}+\frac{10(g-2j)}{g-2}{g-2 \choose j-1}
2103:  \mbox{ }\mbox{ }\mbox{ for all }0\leq j\leq i,$$
2104: $$b_{1j}=\frac{g-2j-1}{g-1}{g-1 \choose j}\mbox{ when }
2105: j\leq i-1, b_{1i}=0, b_{2i}=\frac{(g-2i)^3-(g-2i)}{2g-2}{g \choose i}$$
2106:  $$b_{2j}=\frac{(g-2j-1)(g^3-g^2-4g^2j+4j^2g+2jg-2j)(g-2)!}{2j!(g-1)!}
2107:  \mbox{ for }j\leq i-1.$$
2108: \end{proposition}
2109: 
2110: \begin{proof}
2111:  We will compute the class of $Y_j$
2112:  when $j\leq i-1$. The class of $Y_i$ is computed similarly.
2113:  Let us write the following relation in $\mbox{Pic}(\overline{\mathcal{M}}_{g,2})_{\mathbb Q}$:
2114: $$Y_j\equiv_{lin} a_j\lambda+b_{1j}\psi_{y_1}+
2115: b_{2j}\psi_{y_2}-c_j\delta_{0:y_1y_2}+(\mbox{ other boundary terms }).$$
2116: We use the method of test curves to determine the coefficients $a_j,b_{1j}$ and $b_{2j}$, that is, we intersect the classes appearing on both sides of the previous relation with curves inside $\overline{\mathcal{M}}_{g,2}.$ By computing intersection numbers we obtain linear relations between the coefficients $a_j,b_{1j},b_{2j}$.
2117: 
2118: 
2119: By Proposition 4.5 we have that 
2120: \begin{equation}
2121: Z_j:=(\pi_{y_2})_*(Y_j\cdot \Delta_{0:y_1 y_2})
2122: \equiv_{lin} a_j\lambda+c_j\psi_{y_1}+(\mbox{ boundary }).
2123: \end{equation}
2124: Using  the same reasoning as in Proposition~\ref{reduce},
2125:  we obtain that $Z_j$ is the closure in $\overline{\mathcal{M}}_{g,1}$
2126:  of the locus of curves $(C,y_1)$ carrying a
2127:  $\mathfrak g^1_{g-j}$ with $(g-2j)y_1$ in a fiber.
2128: 
2129: In order to determine the coefficient $c_j$ we intersect both sides 
2130: of $(5)$ with a general fiber $F$ of the map
2131:  $\overline{\mathcal{M}}_{g,1}\rightarrow
2132:  \overline{\mathcal{M}}_{g}$: we get that
2133:  $c_j=Z_j\cdot F/\psi_{y_1}\cdot F=b(g-j,g)/(2g-2)$ (cf. Proposition \ref{enumerative}). 
2134: 
2135:   To determine $b_{1j}$ and $b_{2j}$ we use two
2136:  test curves in $\overline{\mathcal{M}}_{g,2}$:
2137:  first, we fix a general curve $C$ of genus $g$ and we obtain
2138:  a family $C_{[1]}=\{(C,y_1,y_2)\}_{y_1\in C}$,
2139:  by fixing a general point $y_2\in C$ and letting $y_1$ vary on $C$.
2140:  From $(5)$, clearly
2141:  $C_{[1]}\cdot Z_j=(2g-1)b_{1j}+b_{2j}-c_j$.
2142:  On the other hand, according to Proposition~\ref{enumerative}
2143: $C_{[1]}\cdot Z_j=c(g-j,g, 1)$.
2144: 
2145:  For a new relation between $b_{1j}$ and $b_{2j}$
2146:  we use the test curve $C_{[2]}=\{(C,y_1,y_2)\}_{y_2\in C}$
2147:  in $\overline{\mathcal{M}}_{g,2}$,
2148:  where this time $y_1$ is a fixed general point
2149:  while $y_2$ varies on $C$. We have the equation $(2g-1)b_
2150: {2j}+b_{1j}-c_j=C_{[2]}\cdot Z_j=c(g-j,g,g-2j-1)$, and since
2151:  $c_j$ is already known we get in this way both $b_{1j}$ and $b_{2j}$.
2152: 
2153:  We are only left with the computation of $a_j$.
2154:  From \cite{EH2}, Theorem 4.1 we know that the class of
2155:  $Z_j$ is a linear combination  of the Brill-Noether
2156:  class and of the class of the divisor of Weierstrass points,
2157:  that is, $Z_j\equiv_{lin} \mu
2158: BN+\nu \mathcal{W}$, where
2159: $$BN:=(g+3)\lambda-\frac{g+1}{6}\delta_{irr}-\sum_{i=1}^{g-1}i(g-i)
2160: \delta_{i:y_1}\mbox{ and }$$
2161: $$\mathcal{W}:=-\lambda+\frac{g(g+1)}{2}\psi_{y_1}-
2162: \sum_{i=1}^{g-1}{g-i+1\choose 2}\delta_{i:y_1}.$$
2163: We already know that $\nu=2c_j/(g(g+1))$.
2164:  To determine $\mu$ we use the following test curve in
2165:  $\overline{\mathcal{M}}_{g,1}$: we take a general curve $B$
2166:  of genus $g-1$ and a general $2$-pointed elliptic curve $(E,0,y_1)$.
2167:  We consider the family $\overline{B}=
2168: \{X_q=B\cup_{q\sim 0}E,y_1\}_{q\in B}$ obtained by identifying
2169:  the variable point $q\in B$ with the fixed point $0\in E$.
2170:  We easily get $\overline{B}\cdot \psi_{y_1}=
2171: \overline{B}\cdot \lambda=0$,
2172:  $\overline{B}\cdot \delta_{1:y_1}=-\mbox{deg}K_B=4-2g$, 
2173: while $\overline{B}$ vanishes on all the other boundaries.
2174:  On the other hand $\overline{B}\cdot Z_j$ is the number of
2175:  limit $\mathfrak g^1_{g-j}$'s on the curves $X_q$ having vanishing
2176:  $g-2j$ at the fixed point $y_1\in E$. If $l=(l_B,l_E)$ is such a linear 
2177: series, then using again the additivity of the Brill-Noether numbers
2178:  (cf. \cite{EH1}, Proposition 4.6) and the assumption that $y_1-0\in
2179:  \mbox{Pic}^0(E)$ is not torsion, we obtain that $w^{l_B}(q)=g-2j$,
2180:  so either $a^{l_B}(q)=(1,g-2j)$ or $a^{l_B}(q)=(0,g-2j+1)$.
2181:  Thus $\overline{B}\cdot Z_j=b(g-j-1,g-1)+b(g-j,g-1)$ and we
2182:  can write a new relation enabling us to compute $a_j$.
2183: \end{proof}
2184: 
2185: 
2186: We can now complete the proof of Theorem~\ref{class_D}:
2187: \begin{proof}[Proof of Theorem~\ref{class_D}]
2188:  Let us write
2189:  $D\equiv_{lin}A\lambda+B_1\Psi_x+B_2\Psi_y$, where
2190:  $\Psi_x:=\sum_{j=1}^{g-i}\psi_{x_j}$ and $\Psi_{y}:=
2191: \sum_{j=1}^{i+1}\psi_{y_j}$. As noticed before, the
2192:  $\{\lambda, \Psi_y\}$-part of $[D]$ and the $\{\lambda, \psi_{y_1}\}$-part
2193:  of $\sum_{j=0}^i [Y_j]$ coincide, hence using Proposition \ref{class_Y}
2194: $$A=\sum_{j=0}^i a_j=-{g-1 \choose i}+10{g-3 \choose i-1}\mbox{ and }
2195:  B_2=\sum_{j=0}^i b_{1j}={g-2\choose i-1}.$$ 
2196: 
2197:  Finally, to determine $B_1$ one has to compute
2198:  the $\psi_{x_1}$ coefficient of the divisor $D_{x_2}$
2199:  on $\overline{\mathcal{M}}_{g,3}$. Arguing in a way that
2200:  is entirely similar to Proposition \ref{sum_Y} we obtain that
2201:  $B_1={g-2 \choose i}$.
2202: \end{proof}
2203:  
2204:  
2205: 
2206: 
2207: 
2208: %----------------------------------------
2209: %%  REFERENCES
2210: %%----------------------------------------------------------
2211: %%\bibliographystyle{amsalpha}
2212: %%\bibliography{MRC}
2213: 
2214: \providecommand{\bysame}{\leavevmode\hbox to3em{\hrulefill}\thinspace}
2215: 
2216: \begin{thebibliography}{EMS}
2217: 
2218: \bibitem[AC1]{AC1} 
2219: E. Arbarello and M. Cornalba, \emph{Footnotes to a paper of Beniamino 
2220: Segre}, Math. Ann. \textbf{256} (1981), 341--362.
2221: 
2222: \bibitem[AC2]{AC2} 
2223: E. Arbarello and M. Cornalba, \emph{Calculating
2224:  cohomology groups of moduli spaces of curves via
2225:  algebraic geometry}, Inst. Hautes Etudes Sci. Publ. Math.
2226:  No. 88 (1998), 97--127.
2227: 
2228: \bibitem[ACGH]{ACGH}
2229: E. Arbarello, M. Cornalba, P.A. Griffiths and J. Harris,
2230: \emph{Geometry of algebraic curves}, Grundlehren 267, Springer, 1985.
2231: 
2232: \bibitem[BG]{BG}
2233: E. Ballico and A.V. Geramita, \emph{The minimal free resolution of the
2234: ideal of $s$ general points in $\PP^3$}, Proceedings of the 1984
2235: Vancouver conference in algebraic geometry, 1--10, CMS Conf. Proc. \textbf{6},
2236: Amer. Math. Soc., Providence, RI, 1986.
2237: 
2238: \bibitem[Ei]{eisenbud}
2239: D. Eisenbud, \emph{Commutative algebra with a view
2240: toward algebraic geometry}, Springer, 1995.
2241: 
2242: \bibitem[EH1]{EH1}
2243: D. Eisenbud and J. Harris,
2244:  \emph{Limit linear series: basic theory}, Invent. Math. \textbf{85} (1986),
2245: no. 2, 337--371.
2246: 
2247: \bibitem[EH2]{EH2}
2248: D. Eisenbud and J. Harris, \emph{Irreducibility of some families
2249:  of linear series with Brill-Noether number $-1$},
2250:  Ann. Scient. Ec. Norm. Sup. (4), \textbf{22} (1989), no. 1, 33--53.
2251: 
2252: \bibitem[EP]{popescu}
2253: D. Eisenbud and S. Popescu, \emph{Gale duality
2254: and free resolutions of ideals of points}, Invent. Math. \textbf{136}
2255: (1999), no. 2, 419--449.
2256: 
2257: \bibitem[EPSW]{epsw}
2258: D. Eisenbud, S. Popescu, F.--O. Schreyer, Ch. Walter,
2259: \emph{Exterior algebra methods for the Minimal Resolution Conjecture}, 
2260: preprint 2000, math.AG/0011236.
2261: 
2262: \bibitem[EV]{ev}
2263: D. Eisenbud and A. Van de Ven, \emph{On the normal bundles of smooth
2264: rational space curves}, Math. Ann. \textbf{256} (1981), 453--463.
2265: 
2266: \bibitem[Fu]{fulton}
2267: W. Fulton, \emph{Intersection theory}. Second edition.
2268: Springer-Verlag, Berlin, 1998.
2269: 
2270: \bibitem[Ga]{gaeta}
2271: F. Gaeta, \emph{A fully explicit resolution of the ideal
2272: defining $N$ generic points in the plane}, preprint 1995.
2273: 
2274: \bibitem[Gr]{green}
2275: M. Green, \emph{Koszul cohomology and geometry}, in 
2276: \emph{Lectures on Riemann surfaces}, World Scientific Press, Singapore, 
2277: 177--200.
2278: 
2279: \bibitem[GL1]{gl1}
2280: M. Green and R. Lazarsfeld, \emph{On the projective normality of complete 
2281: linear series on an algebraic curve}, Invent. Math. \textbf{83} (1986), 73--90.
2282: 
2283: \bibitem[GL2]{gl2}
2284: M. Green and R. Lazarsfeld,
2285:  \emph{A simple proof of Petri's theorem on canonical 
2286: curves}, in \emph {Geometry Today}, Progress in Math. Birkhauser, 1986.
2287: 
2288: \bibitem[GL3]{gl3}
2289: M. Green and R. Lazarsfeld, \emph{Some results on the syzygies of finite sets
2290: and algebraic curves}, Compositio Math. \textbf{67} (1988), 301--314.
2291: 
2292: \bibitem[Ha]{H}
2293: J. Harris, \emph{On the Kodaira dimension of the moduli
2294:  space of curves. The even genus case},
2295: Invent. Math. \textbf{75} (1984), no.3, 437--466.
2296: 
2297: \bibitem[HaMo]{HaMo}
2298: J. Harris and I. Morrison, \emph{Moduli of curves}, Springer-Verlag,
2299: New York 1998.
2300: 
2301: \bibitem[HaMu]{HM}
2302: J. Harris and D. Mumford, \emph{On
2303:  the Kodaira dimension of the moduli space of curves},
2304: Invent. Math. \textbf{67} (1982), no.1, 23--88.
2305: 
2306: \bibitem[HS]{simpson}
2307: A. Hirschowitz and C. Simpson,
2308:  \emph{La r\'esolution minimale de l'arrangement d'un 
2309: grand nombre de points dans $\PP^n$}, Invent. Math. \textbf{126} (1996), no.3, 467--503. 
2310: 
2311: \bibitem[Lz1]{lazarsfeld}
2312: R. Lazarsfeld, \emph{A sampling of vector bundle techniques in
2313: the study of linear series}, in Lectures on Riemann Surfaces, \emph{World
2314: Scientific Press}, Singapore 1989, pp. 500--559.
2315: 
2316: \bibitem[Lz2]{Rob}
2317: R. Lazarsfeld, private communication.
2318: 
2319: \bibitem[Log]{logan} 
2320: A. Logan, \emph{Moduli spaces of curves with marked points}, Ph.D. thesis, Harvard University, 1999.
2321: 
2322: \bibitem[Lo1]{lorenzini1}
2323: A. M. Lorenzini, \emph{On the Betti numbers of points in projective
2324: space}, Ph.D. thesis, Queen's University, Kingston, Ontario, 1987.
2325: 
2326: \bibitem[Lo2]{lorenzini2}
2327: A. M. Lorenzini, \emph{The minimal resolution conjecture},
2328: J. Algebra \textbf{156} (1993), no. 1, 5--35. 
2329: 
2330: \bibitem[Mu]{mustata}
2331: M. Musta\c{t}\v{a}, 
2332: \emph{Graded Betti numbers of general finite
2333: subsets of points on projective varieties}, Le Matematiche, vol. LIII, 1998,
2334: 53--81.
2335: 
2336: \bibitem[PR]{paranjape}
2337: K. Paranjape and S. Ramanan,
2338:  \emph{On the canonical ring of a curve}, Algebraic geometry
2339: and commutative algebra, vol. II, 503--516, Kinokuriya, 1988.
2340: 
2341: \bibitem[Po]{popa}
2342: M. Popa, \emph{On the base locus of the generalized theta divisor},  
2343: C. R. Acad. Sci. Paris \textbf{329} (1999), S\'erie I, 507--512.
2344: 
2345: \bibitem[Ra]{raynaud}
2346: M. Raynaud, \emph{Sections des fibr\'es vectoriels sur une courbe}, Bull.
2347: Soc. Math. France \textbf{110} (1982), 103--125.
2348: 
2349: \bibitem[Wa]{walter}
2350: Ch. Walter, \emph{The minimal free resolution of the homogeneous ideal
2351: of $s$ general points in $\PP^4$}, Math. Zeitschrift \textbf{219} (1995),
2352: no. 2, 231--234.
2353: 
2354: 
2355: \end{thebibliography}
2356: 
2357: \end{document}
2358: 
2359: