math0701662/ceg.tex
1: \documentclass[12pt]{article}
2: 
3: %%%%%%%%%%
4: %
5: % VERSAO ATUAL (23 January 2007)
6: %
7: %%%%%%%%%%
8: 
9: \usepackage{graphicx}
10: \usepackage{amscd}
11: \usepackage{theorem}
12: 
13: \hyphenation{mo-du-lo}
14: \hyphenation{mo-du-li}
15: \hyphenation{co-ni-cs}
16: \hyphenation{de-ve-lo-ped}
17: 
18: \newfont{\eulerfraktur}{eufm10 scaled\magstep1}
19: \newcommand{\TS}{\mbox{\eulerfraktur {X}}}
20: 
21: \newfont{\Bbb}{msbm10 scaled\magstep1}
22: \def\AA{\mbox{\Bbb A}}
23: \newcommand{\CC}{{\mbox{\Bbb C}}}
24: \newcommand{\EE}{{\mbox{\Bbb E}}}
25: \newcommand{\PP}{{\mbox{\Bbb P}}}
26: \newcommand{\GG}{{\mbox{\Bbb G}}}
27: \newcommand{\NN}{{\mbox{\Bbb N}}}
28: \newcommand{\QQ}{{\mbox{\Bbb Q}}}
29: \newcommand{\LL}{{\mbox{\Bbb L}}}
30: \newcommand{\RR}{{\mbox{\Bbb R}}}
31: \newcommand{\WW}{{\mbox{\Bbb W}}}
32: \newcommand{\DD}{{\mbox{\Bbb D}}}
33: \newcommand{\ZZ}{{\mbox{\Bbb Z}}}
34: 
35: \def\bdm{\begin{displaymath}}
36: \def\be{\begin{equation}}
37: \def\bfe{{\bf e}}
38: \def\bfu{{\bf u}}
39: \def\ka{\kappa_}
40: \def\bfv{{\bf v}}
41: \def\bfx{{\bf x}}
42: \def\bfy{{\bf y}}
43: \def\cw{\overline{wt(2)}}
44: 
45: \def\CMF{\overline{{\cal M}}_g}
46: \def\CMS{\overline{M}_g}
47: \def\CMSP{\overline{M}_{g,1}}
48: \def\dex{\frac{\d}{\d x}}
49: \def\dey{\frac{\d}{\d y}}
50: \def\d{{\partial}}
51: \def\edm{\end{displaymath}}
52: \def\ee{\end{equation}}
53: 
54: \def\fl{f{\hskip 1pt}l}
55: \def\f{~(\ref}
56: 
57: \def\hb{\lambda}
58: \def\KK{{\bf K}}
59: \def\kk{{\bf k}}
60: \def\lra{{\longrightarrow}}
61: 
62: \def\norma{\|.\|}
63: \def\ora{\overrightarrow}
64: 
65: \def\cocoa{{\hbox{\rm C\kern-.13em o\kern-.07em C\kern-.13em o\kern-.15em A}}}
66: \def\wh{\hat}
67: \def\wgt{{\rm wt}}
68: \renewcommand{\:}{\colon\,}
69: \newcommand{\x}{\times}
70: \newcommand{\ol}{\overline}
71: \newcommand{\wt}{\widetilde}
72: \newcommand{\w}{\omega}
73: 
74: \def\proof{{\bf Proof.}\,\,\,}
75: 
76: \def\qed{\hfill\vrule height4pt width4pt depth0pt}
77: \def\rds{{\omega_\pi}}
78: 
79: 
80: \def\gf{\pi:\TS\lra S}
81: \def\ggp{{\overline{\eta}}}
82: \def\uu{(u_1,\ldots,u_n)}
83: \def\vv{(v_1,\ldots,v_n)}
84: 
85: \def\xx{(x_1,\ldots,x_n)}
86: 
87: \def\dmyb{(\pi^*\bigwedge^g\pi_*\rds)^\vee}
88: \def\myb{\pi^*\bigwedge^g\pi_*\rds}
89: \def\wesp{{\otimes\frac{g(g+1)}{2}}}
90: \def\N{\frac{g(g+1)}{2}}
91: \def\umega{{\underline{\omega}}}
92: \def\hy{hyperelliptic\,\,}
93: 
94: \theoremstyle{change} 
95: \newtheorem{claim}{\hskip-0.2cm}[section]
96: 
97: \begin{document}
98: 
99: \title{Special ramification loci 
100: on the double product of a general curve
101: \thanks{2000 Mathematics Subject Classification: 
102: 14H10, 14H15, 14N10.}}
103: 
104: 
105: \author{{\small{{ C. CUMINO $^{(\natural)}$}, 
106: E. ESTEVES $^{(\flat)}$  AND  
107: {L. GATTO} $^{(\natural)}$ }}
108: \thanks{Work partially sponsored by MURST 
109: (Progetto Nazionale ``Geometria delle Variet\`a Proiettive'' 
110: Coordinatore Sandro Verra), and supported by 
111: GNSAGA-INDAM. The second author was also supported
112: by CNPq, Proc. 478625/03-0 and 301117/04-7, 
113: and CNPq/FAPERJ, Proc. E-26/171.174/2003.}\\
114: { }\\
115: {\normalsize $^{(\natural)}$ Dipartimento di 
116: Matematica, Politecnico di Torino,}\\
117: {\normalsize Corso Duca degli Abruzzi 24, 
118: 10129 Torino -- (ITALY)}\\{æ}
119: \\{\normalsize $^{(\flat)}$ Instituto Nacional de 
120: Matem\'atica Pura e Aplicada,}\\
121: {\normalsize Estrada Dona Castorina 110} \\ 
122: {\normalsize 22460-320 Rio de Janeiro RJ -- (BRAZIL)}}
123: 
124: \date{ }
125: 
126: \maketitle
127: 
128: \begin{abstract}
129: \noindent
130: Let $C$ be a general connected, smooth, projective 
131: curve of 
132: positive genus $g$. For each integer $i\geq 0$ we 
133: give formulas for the number of pairs 
134: $(P,Q)\in C\times C$ off the diagonal such that 
135: $(g+i-1)Q-(i+1)P$ is linearly equivalent to an 
136: effective divisor, and the number of pairs 
137: $(P,Q)\in C\times C$ off the diagonal such that 
138: $(g+i+1)Q-(i+1)P$ is linearly equivalent to a moving 
139: effective divisor.
140: \end{abstract}
141: 
142: \section{Introduction}\label{sec1}
143: 
144: Let $C$ be a general connected, smooth, projective 
145: curve of genus $g>0$. Put $C^2:=C\times C$, and let 
146: $\Delta\subset C^2$ be the diagonal. For each 
147: integer $i\geq 0$ consider the 
148: following loci on $C^2$:
149:      \begin{eqnarray*}
150:        D_i&:=&\{(P,Q)\in C^2-\Delta\,|\,
151:        h^0({\cal O}_C((g+i-1)Q-(i+1)P))>0\},\\
152:        E_i&:=&\{(P,Q)\in C^2-\Delta\,|\,
153:        h^0({\cal O}_C((g+i+1)Q-(i+1)P))>1\}.       
154:      \end{eqnarray*}
155: Our Proposition \ref{A2CC} 
156: claims that $D_i$ and $E_i$ are 
157: finite, and our main result, 
158: Theorem \ref{thm6DiazCuk}, gives 
159: formulas for the number of points in $D_i$ and $E_i$.
160: 
161: A formula for the number of points in $D_i$ appeared 
162: already as Lemma~6.3 on page 24 
163: of the seminal work by 
164: Diaz \cite{Diaz/Exc}, where the unnecessary extra 
165: hypotheses that $g\geq 2$ and $i\geq 2$ are made. 
166: Diaz used this formula to compute the class in 
167: the moduli space of genus-$g$ stable curves 
168: $\CMF$ of the closure $\ol{\cal D}_g$ 
169: of the locus of smooth 
170: curves $C$ having a Weierstrass point $P$ 
171: of type $g-1$, i.e. such that 
172: $h^0({\cal O}_C((g-1)P))\geq 2$.
173: 
174: Later on, Cukierman \cite{Cukie} gave a formula for 
175: the class in $\CMF$ of the closure $\ol{\cal E}_g$ 
176: of the locus of smooth 
177: curves $C$ containing a Weierstrass point $P$ of 
178: type $g+1$, 
179: i.e. such that $h^0({\cal O}_C((g+1)P))\geq 3$. 
180: He did not follow in Diaz's 
181: footsteps for this formula, but 
182: rather observed that the union 
183: $\ol{\cal D}_g\cup\ol{\cal E}_g$ is the branch 
184: locus of the Weierstrass divisor on the 
185: ``universal'' curve over $\CMF$, and used a Hurwitz 
186: formula with singularities to compute the class of 
187: this branch divisor.
188:  
189: Had Cukierman followed in Diaz's footsteps, he 
190: would probably have found he needed a formula for 
191: the number of points in $E_i$. We give this formula 
192: here.
193: 
194: In fact, in a sense to be explained below, it is 
195: slightly easier to obtain the number of points in 
196: $E_i$ than in $D_i$, though we obtain both in a 
197: quite integrated form here. 
198: To obtain these numbers, the 
199: natural procedure is to use Porteous formula 
200: to compute the virtual classes of certain 
201: natural ramification schemes $D_i^+$ and $E_i^+$ of 
202: maps of vector bundles on $C^2$; see 
203: Subsection \ref{srs}. Set-theoretically, $D_i^+$ and 
204: $E_i^+$ are given exactly as $D_i$ and $E_i$, but 
205: without the restriction that the pair $(P,Q)$ lies 
206: off $\Delta$. 
207: 
208: The problem is that $D_i^+$ and $E_i^+$ are both 
209: larger than $D_i$ and $E_i$. Indeed, $E_i^+$ is the 
210: union of $E_i$ with the set of points $(P,P)$ such 
211: that $P$ is a Weierstrass point of $C$ and, worse, 
212: $D_i^+$ is the union of $D_i$ and the whole diagonal 
213: $\Delta$. Since $E_i^+$ is finite, 
214: Porteous formula does give an 
215: expression for the number of points in $E_i^+$, with 
216: weights, and thus at least an upper 
217: bound for the number of points in $E_i$. But it 
218: does not a priori give any information on $D_i$.
219: 
220: To compute the number of points in $D_i$ and $E_i$, 
221: we use the fact that, by the Riemann-Roch Theorem, 
222: the union of $D_i$ and $E_i$ is the locus $SW_i$ of 
223: pairs $(P,Q)\in C^2-\Delta$ such that $Q$ is a 
224: special ramification point of the complete linear 
225: system $H^0(\omega_C((i+1)P))$, where $\omega_C$ 
226: is the canonical bundle of $C$. 
227: 
228: We give $SW_i$ a scheme structure as follows. First, 
229: we consider the ramification divisor 
230: $Z_i\subset C^2$ of the 
231: family of linear systems $H^0(\omega_C((i+1)P))$ 
232: parameterized by $P\in C$. Our Proposition \ref{prop5} 
233: implies that 
234: $Z_i$ contains $\Delta$ with multiplicity exactly 
235: $g$. Set $W_i:=Z_i-g\Delta$. Our 
236: Proposition \ref{prop7} 
237: gives an expression for the cycle $[W_i]$, and 
238: our Proposition \ref{lem} claims that $W_i$ is 
239: nonsingular. Furthermore, in Subsection \ref{srs} we 
240: observe that the branch divisor of $W_i$ with 
241: respect to the projection $p_1\: C^2\to C$ over the 
242: first factor has support $SW_i$. We give 
243: $SW_i$ the structure of this branch divisor. 
244: 
245: The advantage of considering $SW_i$ is 
246: that it is quite easy to compute its degree. 
247: Indeed, $[SW_i]$ is the second Chern class 
248: of the bundle of first-order relative jets of $p_1$ 
249: with coefficients in ${\cal O}_{C^2}(W_i)$. 
250: Having an expression for $[W_i]$ we derive very 
251: quickly an expression for $\int_{C^2}[SW_i]$ in 
252: Proposition \ref{prop07}
253: 
254: Now, giving $D_i$ and $E_i$ the subscheme 
255: structures induced from $D_i^+$ and $E_i^+$, our 
256: Proposition \ref{A2CC} shows that, as 0-cycles, 
257:         $$[D_i]+[E_i]=[SW_i].$$
258: Actually, $D_i$ and $E_i$ are reduced. Indeed, in 
259: the proof of Theorem \ref{thm6DiazCuk} we show 
260: that the weight of $(P,Q)$ in $[SW_i]$ is at most 2, 
261: and the maximum weight is achieved if and only if 
262: $(P,Q)\in D_i\cap E_i$. 
263: 
264: Now, as we already know $\int_{C^2}[SW_i]$, 
265: it is enough to compute either $\int_{C^2}[D_i]$ or 
266: $\int_{C^2}[E_i]$. As mentioned above, we compute 
267: the latter. In fact, we can get $\int_{C^2}[E_i^+]$ 
268: using Porteous formula, and a local analysis, done 
269: in Proposition \ref{A2CC}, 
270: shows that the weight of $(P,P)$ 
271: in $[E_i^+]$ is equal to $g+1$ for each Weierstrass 
272: point of $C$. Thus $\int_{C^2}[E_i]$ follows.
273: 
274: In a second article \cite{CEG}, we 
275: show how the knowledge of the number of points in 
276: $E_i$ can be used to compute the class of 
277: $\ol{\cal E}_g$ in $\ol{\cal M}_g$. 
278: This computation is not 
279: straightforward as, following in Diaz's footsteps, 
280: we have to determine 
281: the limits of 
282: special Weierstrass points of type $g+1$ on stable 
283: curves with just one node. This is the main result 
284: of \cite{CEG}. 
285: 
286: The limits of special 
287: Weierstrass points of type $g-1$ were computed by 
288: Diaz, using admissible covers. However, the same 
289: method does not apply to points of type $g+1$. For 
290: those we apply in \cite{CEG} 
291: the theory of limit linear series in 
292: a rather new way, using 2-parameter families. 
293: Actually, as in the present article, we use an 
294: integrated approach in \cite{CEG} that 
295: yields simultaneously 
296: the limits of special Weierstrass points 
297: of both types, and also formulas for 
298: the classes of
299: both $\ol{\cal D}_g$ and $\ol{\cal E}_g$.
300: 
301: Here is a layout of the article. In 
302: Section \ref{sec2} we 
303: review the theory of linear systems and ramification 
304: on 
305: a
306: smooth curve $C$, introduce the linear 
307: systems 
308: we will consider in the remaining of the article, 
309: $H^0(\omega_C((i+1)P))$ for $P\in C$, and prove a 
310: preliminary result about them. In 
311: Section \ref{sec3}, 
312: assuming $C$ is general, we obtain through 
313: degeneration methods results that bound the order 
314: sequence of $H^0(\omega_C((i+1)P))$ at any point of 
315: $C$. In Section \ref{sec4}, 
316: we describe the structure 
317: of the ramification divisor $Z_i\subset C^2$ of the 
318: family of linear systems $H^0(\omega_C((i+1)P))$ 
319: parameterized by $P\in C$. 
320: Finally, in Section \ref{sec5} we define the loci 
321: $D_i$ and $E_i$ and compute their number of points, 
322: through the study of the locus $SW_i$ of special 
323: ramification points of the family 
324: $H^0(\omega_C((i+1)P))$ for $P\in C$.
325: 
326: We thank Nivaldo Medeiros for discussions on related 
327: topics. Also, we acknowledge the use of \cocoa 
328: \cite{cocoa} for some of the computations.
329: 
330: \section{Setup}\label{sec2}
331: 
332: \begin{claim} 
333: {\rm ({\it Linear systems and ramification})
334: Let $C$ be a {\it smooth curve}, that is, 
335: a projective, connected,
336: smooth scheme of dimension 1 over $\CC$. Denote by $\w_C$
337: its canonical sheaf. Let $g:=h^0(C,\w_C)$, 
338: the genus of $C$.
339: 
340: Let $V$ be a $\CC$-vector space of sections
341: of a line bundle $\cal L$ on $C$. We call $V$ a
342: {\it linear system}. The linear system is
343: called {\it complete} if $V=H^0(C,{\cal L})$.
344: Let $r:=\dim V-1$ and $d:=\deg{\cal L}$. We call
345: $r$ the {\it rank} of $V$ and $d$ its {\it degree}. We say
346: as well that $\dim V$ is the {\it dimension} of $V$.
347: 
348: For each point $P$ of $C$, and each integer $j\geq 0$,
349: let $V(-jP)$ denote the vector subspace of $V$ of sections
350: of $\cal L$ that vanish with order at least
351: $j$ at $P$. We say that $j$ is an {\it order} of
352: $V$ at $P$ if $V(-jP)\neq V(-(j+1)P)$. There are
353: $r+1$ orders, which, in an increasing sequence, will be
354: denoted by
355:        $$\epsilon_0(V,P),\,\epsilon_1(V,P),\,\dots,\,
356:        \epsilon_r(V,P).$$
357: 
358: For each integer $\ell\geq 0$ and 
359: each line bundle ${\cal M}$ on
360: $C$, let ${\cal J}^\ell_C({\cal M})$ 
361: be the bundle of {\it jets},
362: or {\it principal parts}, of $\cal M$ 
363: truncated in order $\ell$.
364: Consider the map of rank-$r$ bundles,
365:       $$V\otimes{\cal O}_C\longrightarrow
366:       {\cal J}^r_C({\cal L}),$$
367: locally obtained by differentiating up to 
368: order $r$ the sections
369: of $\cal L$ in $V$. The {\it wronskian} $w_V$ of 
370: $V$ is the (nonzero) section of
371:       $${\cal L}^{\otimes r+1}\otimes
372:       \w_C^{\otimes r(r+1)/2}$$
373: induced by taking determinants in the
374: above map of bundles.
375: 
376: For each point $P$ of $C$, the {\it weight} $\wgt_V(P)$ of
377: $P$ in $V$ is the order of vanishing of $w_V$ at $P$.
378: We call $P$ a {\it ramification point} of $V$ if
379: $\wgt_V(P)>0$; otherwise we call $P$ {\it ordinary}.
380: A local analysis yields the formula
381:        $$\wgt_V(P)=\sum_{j=0}^r
382:        (\epsilon_j(V,P)-j).$$
383: We call $P$ a {\it simple} ramification point if
384: $\wgt_V(P)=1$; otherwise we call $P$ {\it special}.
385: The point $P$ is special if and only if the section
386:        $$Dw_V\in H^0\Big(C,{\cal J}^1_C
387:        ({\cal L}^{\otimes r+1}\otimes
388:        \w_C^{\otimes r(r+1)/2})\Big),$$
389: locally obtained from $w_V$ by differentiating, 
390: vanishes at $P$.
391: 
392: The {\it total weight} of the ramification points of
393: $V$ is the (finite) sum
394:       $$\wgt_V:=\sum_{P\in C}\wgt_V(P).$$
395: It is equal to the degree of the line bundle of which
396: $w_V$ is a section, that is,
397:       $$\wgt_V=(r+1)(d+(g-1)r),$$
398: a formula usually referred to as the
399: {\it Brill--Segre} or {\it Pl\"ucker formula}.
400: 
401: The {\it canonical system} is the complete linear
402: system of sections of $\w_C$. Its rank is $g-1$, and its
403: degree is $2(g-1)$. For each point $P$ of $C$, its
404: {\it Weierstrass weight} $\wgt(P)$
405: is its weight in the canonical system, and the
406: {\it Weierstrass order sequence} at
407: $P$ is the increasing sequence of orders at $P$ of
408: the canonical system.
409: 
410: For each integer $i\geq -1$ and each $P\in C$,
411: let $V_C(i,P)$ denote the complete linear system of
412: sections of $\w_C((i+1)P)$.}
413: \end{claim}
414: 
415: \begin{claim}\label{prop5}{\bf Proposition.} 
416: Let $C$ be a smooth
417: curve of genus $g$. For each integer
418: $i\geq 0$ and each $P\in C$, the following two
419: statements hold for $V:=V_C(i,P)$:
420: \begin{enumerate}
421: \item\label{it5}
422: The weight $\wgt_V(P)$ of $P$ as a ramification point of
423: $V$ satisfies
424:       \be
425:       \wgt_V(P)=g+\wgt(P),\label{eq:15}
426:       \ee
427: where $\wgt(P)$ is the Weierstrass weight of $P$.
428: \item \label{it4} The total weight $\wgt_V$ of
429: the ramification points of $V$ satisfies
430:       \be
431:       \wgt_V=g(g+i)^2.\label{eq:14}
432:       \ee
433: \end{enumerate}
434: \end{claim}
435: 
436: \proof From the Riemann--Roch theorem, for
437: each $j=0,\dots,i$,
438:       $$\dim V(-jP)=g+i-j.$$
439: In particular, comparing dimensions, we get
440:       $$V(-iP)=V(-(i+1)P)=H^0(C,\w_C).$$
441: Hence, the order sequence of $V$ at $P$ is
442:       $$0,\, 1,\, \dots,\, i-1,\,
443:       (i+1)+\epsilon_0,\, (i+1)+\epsilon_1,\,
444:       \dots,\, (i+1)+\epsilon_{g-1},$$
445: where $\epsilon_0,\,\epsilon_1,\,\dots,\,\epsilon_{g-1}$ 
446: is the Weierstrass order sequence at $P$. Thus
447:       $$\wgt_V(P)=\sum_{k=0}^{g-1}(i+1+\epsilon_k-i-k)
448:       =g+\wgt(P).$$
449: 
450: The second statement is a direct application of
451: the Brill--Segre formula, using that the rank of $V$
452: is $g+i-1$ and its degree is $2g-2+(i+1)$.
453: \qed
454: 
455: \section{The general curve}\label{sec3}
456: 
457: \begin{claim}\label{WiPgen}{\bf Proposition.}
458: Fix an integer $i_0\geq 0$.
459: Let $C$ be a general smooth curve of genus
460: $g\geq 1$. Then the following two statements hold for
461: each nonnegative integer $i\leq i_0$:
462: \begin{enumerate}
463: \item For a general point $P$ of $C$, the linear system
464: $V_C(i,P)$ ramifies at $P$ with weight $g$, and has
465: otherwise at most simple ramification points.
466: \item For any two points $P$ and $R$ of $C$,
467:     \begin{equation}\label{WP-R1}
468:       h^0(C,\w_C((i+1)P-(g+i)R))\leq 1
469:     \end{equation}
470: \vskip-0.8cm
471:     \begin{equation}\label{WP-R2}
472:       h^0(C,\w_C((i+1)P-(g+i+2)R)=0.
473:     \end{equation}
474: \end{enumerate}
475: \end{claim}
476: 
477: \proof Let us first observe that the property required
478: of $C$ is open. Indeed, let $f\: X\to S$ be any family of
479: smooth curves, that is, a projective, smooth map with
480: connected fibers of dimension 1. Consider the fibered
481: product $X^{(2)}:=X\x_S X$ of two copies of $f$, and
482: denote by $p_1$ and $p_2$ the projection maps. Denote by
483: $\Delta$ the diagonal subscheme of $X^{(2)}$. Let
484: $\w_f$ denote the relative canonical bundle of $f$.
485: Then $\w:=p_2^*\w_f$ is the relative canonical bundle
486: of $p_1$. Let
487:         $${\cal V}:=p_{1*}(\w((i+1)\Delta)).$$
488: A fiberwise analysis shows that 
489: $\cal V$ is a bundle of rank $g+i$
490: with formation commuting with base change. 
491: For each integer
492: $\ell\geq 0$, denote by $\cal J^\ell$ the 
493: bundle 
494: of rank $\ell+1$
495: of $p_1$-relative
496: jets of $\w((i+1)\Delta)$ truncated in 
497: order $\ell$, and denote by
498: $\psi_\ell\: p_1^*{\cal V}\to{\cal J}^\ell$ 
499: the map locally
500: obtained by differentiating the sections 
501: of $\w((i+1)\Delta)$ up
502: to order $\ell$ along the fibers of $p_1$. 
503: Let $W_{i,1}$ (resp.
504: $W_{i,2}$) be the closed subset of $X^{(2)}$ 
505: where $\psi_{g+i-1}$
506: (resp. $\psi_{g+i+1}$) has rank at most 
507: $g+i-2$ (resp. $g+i-1$).
508: Also, let $W_i$ be the closed subset 
509: of $X^{(2)}$ where
510: $\psi_{g+i-1}$ has rank at most 
511: $g+i-1$.
512: By
513: Proposition~\ref{prop5}, $W_i$ 
514: contains $\Delta$ with multiplicity
515: $g$. Let $W'_i:=W_i-g\Delta$ and 
516: $Z_i:=\Delta\cap W'_i$. Let
517: $W''_i\subset W'_i$ be the ramification scheme of the map
518: $p_1|_{W'_i}$. Let $U_i\subseteq S$ be the intersection of
519: $S-f(p_1(W_{i,1}\cup W_{i,2}))$ with $f(X-p_1(W''_i\cup Z_i))$.
520: Since $p_1$ is proper, and $f$ is both proper and open, $U_i$ is
521: an open subscheme of $S$. Let $U:=U_0\cap\cdots\cap U_{i_0}$. The
522: formation of $U$ commutes with base change. Thus a fiberwise
523: analysis reveals that $U$ consists of the set of points $s\in S$
524: such that the proposition holds for $C:=X(s)$.
525: 
526: Now, keeping in mind the existence of a versal family of
527: smooth curves, it is enough to exhibit a single curve $C$
528: for which the statement holds.
529: We will actually show a somewhat 
530: stronger existence result:
531: 
532: \begin{claim}\label{WCQ}{\bf Lemma.}
533: Fix nonnegative integers $i_0$ and $j_0$.
534: Let $g$ be a positive integer.
535: Then there is a smooth
536: pointed curve $(C,Q)$ of genus $g$ for which
537: the following three statements hold for
538: each nonnegative integers $i\leq i_0$ and $j\leq j_0$:
539: \begin{enumerate}
540: \item The linear system $V_C(j,Q)$ ramifies at
541: $Q$ with weight g, and has otherwise at most simple
542: ramification points.
543: \item For each $P\in C$ distinct from $Q$, either $Q$
544: is an ordinary point or a simple ramification point of
545: $V_C(i,P)$.
546: \item For each $P\in C$ distinct from $Q$, the linear
547: system $V$ of sections of $\w_C((i+1)P+(j+1)Q)$ given by
548:         $$V:=H^0(\w_C((i+1)P))+H^0(\w_C((j+1)Q))$$
549: satisfies
550:         $$\dim V(-(g+i+j)R)\leq 1\quad{\rm and}\quad
551:         V(-(g+i+j+2)R)=0$$
552: for each $R\in C$ distinct from $P$ and $Q$.
553: \end{enumerate}
554: \end{claim}
555: 
556: We will first see how the lemma implies the
557: proposition. Set $j_0=i_0$, and consider the pointed curve
558: $(C,Q)$ given by the lemma. Then the two statements of
559: Proposition~\ref{WiPgen} hold for $C$. Indeed,
560: the first statement holds for $P=Q$, whence
561: for $P$ in a neighborhood of $Q$, that is,
562: for a general $P$.
563: 
564: As for the second statement, first notice that 
565: (\ref{WP-R1}) and (\ref{WP-R2}) 
566: hold for $P=Q$ and every $R\in C$, a consequence of
567: the first statement of the lemma for $j:=i$. They hold
568: as well for $R=Q$ and every $P\in C$ distinct from $Q$, a
569: consequence of the second statement of the lemma. 
570: Furthermore, they hold for $R=P$ and any $P\in C$. 
571: Indeed, the first statement of the lemma for $j:=0$ 
572: implies that the canonical linear system
573: has at most simple ramification points. Thus
574: $h^0(\w_C((1-g)P))\leq 1$ and $h^0(\w_C(-(g+1)P))=0$.
575: 
576: Finally, fix a point $P\in C$ distinct from $Q$ and a 
577: point $R\in C$ distinct from $P$ and $Q$. For $j:=0$, 
578: the linear system $V$ defined in the lemma is the 
579: system of sections of $\w_C((i+1)P+Q)$ that are zero 
580: on $Q$. Since $R\neq Q$, the third statement of the
581: lemma yields (\ref{WP-R1}) and (\ref{WP-R2}).
582: 
583: It is thus enough to prove the lemma, 
584: what we do below.\qed
585: 
586: \vskip0.2cm
587: 
588: \proof (Lemma \ref{WCQ})
589: We will do induction on $g$.
590: The initial step is taken care of below.
591: 
592: Let $C$ be any elliptic curve and 
593: $Q\in C$ any point. Then the ramification points
594: of the complete linear system of sections of 
595: $\w_C((j+1)Q)$
596: are simple. (These are the $(j+1)^2$ points $R$
597: for which $Q-R$ is $(j+1)$-torsion, what includes $Q$.)
598: In fact, it follows from the Riemann--Roch
599: theorem that every complete linear system has only simple
600: ramification points. Thus Statements~1 and 2 of the lemma
601: hold. Now, given $P\in C$ distinct from $Q$, since the
602: vector subspace $V$ of $H^0(\w_C((i+1)P+(j+1)Q))$ defined
603: in  Statement~3 has codimension 1, the order sequence of
604: $V$ at a point $R$ is obtained either from
605:         $$0,\,1,\,\dots,\,g+i+j-1,\,g+i+j\quad{\rm or}\quad
606:         0,\,1,\,\dots,\,g+i+j-1,\,g+i+j+1$$
607: by removing an order. In any case, there is at most one order of
608: $V$ at $R$ above $g+i+j-1$, that is $\dim V(-(g+i+j)R)\leq 1$, and
609: all orders are at most $g+i+j+1$, that is $V(-(g+i+j+2)R)=0$.
610: 
611: Assume from now on that $g>1$, and that the claim holds for
612: smaller genera and any integers $i_0$ and $j_0$. We will
613: employ a degeneration technique in order to apply the
614: induction hypothesis.
615: 
616: Let $(Y,A)$ and $(Z,B)$ be nonrational smooth pointed curves
617: of genera $g_Y$ and $g_Z$, with $g_Y+g_Z=g$. From the
618: induction hypothesis, we may assume that the statements
619: of the lemma hold for $(C,Q)$ replaced by $(Y,A)$ and
620: all nonnegative integers $i\leq i_0$ and
621: $j\leq g_Z+i_0+j_0+1$, and for
622: $(C,Q)$ replaced by $(Z,B)$ and all
623: nonnegative integers $i\leq i_0$ and $j\leq g_Y+i_0+j_0+1$.
624: 
625: Let $C_0$ be the curve of compact type 
626: that is the union of
627: $Y$, of $Z$, and of a chain of rational curves
628: $E_1,\dots,E_{n-1}$ connecting $A$ to 
629: $B$, where $n\geq 2$.
630: Our convention is that 
631: $E_1$ contains $A$ and $E_{n-1}$ contains $B$.
632: Let $v$ be any integer such that $0<v<n$, and let
633: $Q_0$ be any point of $E_v$ that is not a node of $C_0$.
634: 
635: Let $S:={\rm Spec}(\CC[[t]])$, and denote its special point
636: by 0 and generic point by $\eta$. Since there are no
637: obstructions to deforming pointed nodal curves, there
638: are a projective, flat map $f\: X\to S$ and a section
639: $\lambda\: S\to X$ of $f$ such that
640: $(X(0),\lambda(0))=(C_0,Q_0)$ and
641: $(X(\eta),\lambda(\eta))$ is a smooth pointed curve over
642: the field of formal Laurent series $\CC[[t]][1/t]$.
643: 
644: Let $C$ be the base extension of $X(\eta)$ to the algebraic
645: closure of $\CC[[t]][1/t]$. Set $Q:=\lambda(\eta)$. It is
646: enough to see that the statements of the lemma hold for
647: $(C,Q)$. Indeed, the argument is quite standard, and is
648: summarized below. Though the pointed curve $(C,Q)$ is not
649: defined over $\CC$, it is defined over a finitely generated
650: extension $L$ of $\QQ$. If the statements of the lemma hold
651: for $(C,Q)$, they also hold for the base extension of
652: $(C,Q)$ over any algebraically closed field containing $L$.
653: But, since $\CC$ has many transcendentals
654: over $\QQ$, there is an algebraically closed field
655: containing $L$ which is isomorphic to $\CC$. 
656: So, if the statements of the lemma hold for $(C,Q)$, 
657: they hold as well for some pointed curve over $\CC$.
658: 
659: Now, any finite set of points of $C$ is 
660: defined over a finite field extension of $\CC[[t]][1/t]$.
661: Replacing $S$ by its normalization in this field extension,
662: we may assume that these are rational points of $X(\eta)$,
663: and thus that there are sections of $f$ intersecting
664: $X(\eta)$ at them. By making a further base extension, if
665: necessary, and a sequence of blowups at the singular
666: points of the special fiber, we may assume that the total
667: space $X$ is regular, and that these sections factor through
668: the smooth locus of $f$. The compensation for this
669: is a change of the special fiber. However, the special fiber
670: will have the same specification as the $C_0$ we described
671: above. Thus, no confusion will ensue if we keep calling
672: by $C_0$ this new fiber. Also, the section $\lambda$ can be
673: extended to a section of this new family.
674: 
675: Now, let $P$ and $R$ be points of $C$ with $P$ distinct from
676: $Q$ and $R$ distinct from $P$ and $Q$. As we mentioned
677: above, we may assume there are sections $\gamma\:S\to X$ and
678: $\rho\:S\to X$ through the smooth locus of $f$ such that
679: $\gamma(\eta)=P$ and $\rho(\eta)=R$. Set $P_0:=\gamma(0)$
680: and $R_0:=\rho(0)$. Let $\Gamma$ and $\Lambda$ be the images
681: of $\gamma$ and $\lambda$, respectively.
682: 
683: Fix nonnegative integers $i\leq i_0$ and $j\leq j_0$.
684: Let $\w$ be the relative dualizing bundle of $f\: X\to S$.
685: Let $V_\eta$ be the linear system of sections of the line
686: bundle $\w(\eta)((i+1)P+(j+1)Q)$ given by
687:        $$V_\eta:=H^0(\w(\eta)((i+1)P))+
688:        H^0(\w(\eta)((j+1)Q)).$$
689: Assume that $R$ is a ramification point of $V_\eta$. To
690: prove the statements of the lemma hold for $(C,Q)$, it is
691: enough to prove the following three statements:
692: \begin{enumerate}
693:     \item For $i=0$, the system $V_\eta$ ramifies at $Q$ with weight
694:     $g$, and $R$ is a simple ramification point of $V_\eta$.
695:     \item For $j=0$, the point $Q$ is a ramification point of
696:     $V_\eta$ of weight $g+i$ or $g+i+1$.
697:     \item $\dim V_\eta(-(g+i+j)R)\leq 1$ and
698:     $V_\eta(-(g+i+j+2)R)=0$.
699: \end{enumerate}
700: 
701: We will employ techniques of limit linear series, 
702: from
703: \cite{MathCont}, to show the
704: above three statements. 
705: There are two cases to consider:
706: 
707: \vskip0.2cm
708: 
709: {\noindent\emph{Case 1:} Assume that $P_0\in E_u$
710: for some $u$.}
711: 
712: \vskip0.1cm
713: 
714: Since $C_0$ is of compact type, there is an effective
715: divisor $D$ of $X$ supported on $C_0$ such that, letting
716:        $${\cal L}:=\w((i+1)\Gamma+(j+1)\Lambda+D),$$
717: we have ${\cal L}|_{E_m}\cong{\cal O}_{E_m}$ for each
718: $m=1,\dots,n-1$,
719:        $${\cal L}|_Z\cong\w_Z((g_Y+i+j+3)B)\quad
720:        {\rm and}\quad
721:        {\cal L}|_Y\cong\w_Y((1-g_Y)A).$$
722: Since, from the induction hypothesis, $A$ is a ramification
723: point of weight $g_Y$ of the complete linear system of
724: sections of $\w_Y(A)$, the point $A$ is not a Weierstrass
725: point of $Y$. Then $V:=H^0(X,{\cal L})\cap V_\eta$ restricts
726: to a linear system $V_Z$ of dimension $g+i+j$ of sections of
727: $\w_Z((g_Y+i+j+3)B)$. Also from the induction hypothesis,
728: $B$ is not a Weierstrass point of $Z$. So the order
729: sequence of $B$ in the complete linear system of sections of
730: $\w_Z((g_Y+i+j+3)B)$ is
731:        $$0,\,1,\,\dots,\,g_Y+i+j+1,\,g_Y+i+j+3,\,\dots,\,
732:        g+i+j+2.$$
733: As a consequence, the weight $w_B$ of $B$ as a ramification
734: point of the linear system $V_Z$ satisfies
735:        \be\label{wB1}
736:        w_B\leq 2(g_Y+i+j)+3g_Z,
737:        \ee
738: with equality if and only if $V_Z=H^0(\w_Z((g_Y+i+j+1)B))$.
739: 
740: Analogously, choosing an appropriate $D$, we obtain a linear
741: system $V_Y$ of dimension $g+i+j$ of sections of
742: $\w_Y((g_Z+i+j+3)A)$, and the weight $w_A$ of $A$
743: as a ramification point of $V_Y$ satisfies
744:        \be\label{wA1}
745:        w_A\leq 2(g_Z+i+j)+3g_Y,
746:        \ee
747: with equality if and only if $V_Y=H^0(\w_Y((g_Z+i+j+1)A))$.
748: 
749: Let $r:=g+i+j-1$. Using the Pl\"ucker formula, the number
750: $N$ of ramification points of $V_Y$ and $V_Z$ on
751: $(Y-A)\cup(Z-B)$, counted with their respective weights,
752: satisfies
753:        \begin{eqnarray*}
754:        N&=&(r+1)\Big((2g_Z+g_Y+i+j+1)+r(g_Z-1)\Big)-w_B\\
755:        &+&(r+1)\Big((2g_Y+g_Z+i+j+1)+r(g_Y-1)\Big)-w_A\\
756:        &=&N'+5g+4(i+j)-w_A-w_B,
757:        \end{eqnarray*}
758: where
759:        $$N':=(r+1)\Big((2g+i+j)+r(g-1)\Big)-2g-i-j.$$
760: 
761: Now, from the theory of limit linear series, each one of the
762: ramification points of $V_Y$ or $V_Z$ on $(Y-A)\cup(Z-B)$
763: is a limit of ramification points of $V_\eta$, and its
764: weight as a ramification point is the sum of the weights of
765: the ramification points of $V_\eta$ converging to it.
766: Besides those, since $P$ and $Q$ are ramification points
767: of $V_\eta$ with weights at least $g+j$ and $g+i$,
768: respectively, the points $P_0$ and $Q_0$ appear
769: as limits of ramification points of $V_\eta$ with weights
770: summing up to at least $2g+i+j$. Thus, from the Pl\"ucker
771: formula, at most $N'$ ramification points of $V_\eta$,
772: counted with their weights, converge to $(Y-A)\cup(Z-B)$. So
773:        $$5g+4(i+j)-w_A-w_B\leq 0.$$
774: 
775: However, Inequalities (\ref{wB1}) and (\ref{wA1}) for $w_B$
776: and $w_A$ yield the opposite inequality:
777:        $$5g+4(i+j)-w_A-w_B\geq 0.$$
778: Thus, equalities hold, and hence
779:        $$V_Y=H^0(\w_Y((g_Z+i+j+1)A))\quad{\rm and}\quad
780:        V_Z=H^0(\w_Z((g_Y+i+j+1)B)).$$
781: In addition, $P$ and $Q$ are ramification points of
782: $V_\eta$ of weights $g+j$ and $g+i$, respectively, and
783: all the other ramification points of $V_\eta$ converge
784: to $(Y-A)\cup(Z-B)$. In particular, Statement 2 
785: and the first part of Statement~1 are shown.
786: 
787: Now, since $R$ is a ramification point of $V_\eta$, and
788: $R$ is distinct from $P$ and $Q$, we have
789: $R_0\in(Y-A)\cup(Z-B)$. So $R_0$ is a
790: ramification point of either $V_Y$ or $V_Z$. From the
791: induction hypothesis, the complete linear systems of
792: sections of $\w_Y((g_Z+i+j+1)A)$ and $\w_Z((g_Y+i+j+1)B)$
793: have at most simple ramification points, other than $A$ or
794: $B$. Thus $R$ is the unique ramification point of $V_\eta$
795: converging to $R_0$ and its weight is 1.
796: So the remainder of Statement 1 is shown.
797: 
798: As for Statement 3, assume, 
799: without loss of generality, that
800: $R_0\in Z$. Set
801: $n:=\dim V_\eta(-(g+i+j)R)$, and let
802: $\sigma_1,\dots,\sigma_n$ form a $\CC[[t]]$-basis of
803: $V\cap V_\eta(-(g+i+j)R)$. Their restrictions to $Z$ are
804: sections of $V_Z$ vanishing with multiplicity at least
805: $g+i+j$ on $R_0$. Assume, by contradiction, that
806: $n\geq 2$. Since $R_0$ is a simple ramification point of
807: $V_Z$, the sections $\sigma_1|_Z,\dots,\sigma_n|_Z$
808: are linearly dependent. Thus, there is
809: a nonzero $n$-tuple $(c_1,\dots,c_n)\in\CC^n$ such that
810: $c_1\sigma_1+\cdots+c_n\sigma_n$ vanishes on $Z$, and hence
811: on the whole $C_0$. Thus
812:        \be\label{sigma}
813:        c_1\sigma_1+\cdots+c_n\sigma_n=t\sigma
814:        \ee
815: for some $\sigma\in H^0(X,{\cal L})$. Also $\sigma\in
816: V_\eta(-(g+i+j)R)$, and hence $\sigma$ is a $\CC[[t]]$-linear
817: combination of $\sigma_1,\dots,\sigma_n$. Plugging this linear
818: combination in (\ref{sigma}) we obtain a nontrivial
819: $\CC[[t]]$-linear relation among the sections $\sigma_i$, a
820: contradiction. Thus $n\leq 1$. A similar analysis, using that
821: $V_Z(-(g+i+j+2)R_0)=0$, shows that $V_\eta(-(g+i+2)R)=0$,
822: finishing the proof of Statement 3.
823: 
824: \vskip0.2cm
825: 
826: {\noindent\emph{Case 2:} Assume $P_0$ belongs to either $Y$
827: or $Z$.}
828: 
829: \vskip0.1cm
830: 
831: Without loss of generality, we may assume that $P_0\in Z$.
832: Again, since $C_0$ is of compact type, there is an effective
833: divisor $D$ of $X$ supported on $C_0$ such that, letting
834:        $${\cal L}:=\w((i+1)\Gamma+(j+1)\Lambda+D),$$
835: we have ${\cal L}|_{E_m}\cong{\cal O}_{E_m}$ for each
836: $m=1,\dots,n-1$,
837:        $${\cal L}|_Z\cong\w_Z((i+1)P_0+(g_Y+j+2)B)
838:        \quad{\rm and}\quad
839:        {\cal L}|_Y\cong\w_Y((1-g_Y)A).$$
840: As before, $V:=H^0(X,{\cal L})\cap V_\eta$ restricts to a
841: linear system $V_Z$ of dimension $g+i+j$ of sections of
842: $\w_Z((i+1)P_0+(g_Y+j+2)B)$.
843: 
844: Now,
845:        $$V\supseteq H^0(X,\w((i+1)\Gamma))+
846:        H^0(X,\w((j+1)\Lambda+D)).$$
847: Reasoning as in Case 1, we can show that
848: $H^0(X,\w((j+1)\Lambda+D))$ restricts to
849: $H^0(\w_Z((g_Y+j+1)B))$.
850: On the other hand, the exact sequence
851:        $$0\to H^0(\w_Z((i+1)P_0))\to
852:        H^0(\w((i+1)\Gamma)|_{C_0})
853:        \to H^0(\w_Y(A))$$
854: shows that $h^0(\w((i+1)\Gamma)|_{C_0})=g+i$, and hence
855: that $H^0(X,\w((i+1)\Gamma))$ restricts to a vector
856: subspace of $H^0(\w_Z((i+1)P_0+B))$ containing
857: the subspace $H^0(\w_Z((i+1)P_0))$. Thus
858:        $$V_Z\supseteq H^0(\w_Z((i+1)P_0))+
859:        H^0(\w_Z((g_Y+j+1)B)),$$
860: and a dimension count shows that equality holds.
861: 
862: The weight $w_B$ of $B$ as a ramification point of $V_Z$
863: depends on its weight as a ramification point of
864: $V_Z(i,P_0)$. Now, from the induction hypothesis,
865: $B$ is either an ordinary point or a simple ramification
866: point of $V_Z(i,P_0)$.
867: Hence, the order sequence at $B$ of the linear system
868: $V_Z$ is either
869:        $$1,\,2,\,\dots,\,g_Y+j,\,g_Y+j+2,\,g_Y+j+3,\,
870:        \dots,\,g+i+j,\,g+i+j+1$$
871: or
872:        $$1,\,2,\,\dots,\,g_Y+j,\,g_Y+j+2,\,g_Y+j+3,\,
873:        \dots,\,g+i+j,\,g+i+j+2.$$
874: At any rate,
875:        \begin{equation}\label{wB2}
876:        w_B\leq g_Y+j+2(g_Z+i)+1.
877:        \end{equation}
878: Notice that, if $i=0$, then $B$ is an ordinary point of
879: $V_Z(0,P_0)$, as it is an ordinary point of $Z$,
880: and thus Inequality (\ref{wB2}) is strict.
881: 
882: On the other hand, let $D'$ be an effective divisor of $X$
883: supported on $C_0$ such that, letting
884:        $${\cal M}:=\w((i+1)\Gamma+(j+1)\Lambda+D'),$$
885: we have ${\cal M}|_{E_m}\cong{\cal O}_{E_m}$ for each
886: $m=1,\dots,n-1$,
887:        $${\cal M}|_Y\cong\w_Y((g_Z+i+j+3)A)
888:        \quad{\rm and}\quad
889:        {\cal M}|_Z\cong\w_Z((i+1)P_0-(g_Z+i)B).$$
890: Since, as mentioned above, $B$ is either an ordinary point
891: or a simple ramification point of $V_Z(i,P_0)$, we have that
892: $H^0(X,{\cal M})\cap V_\eta$ restricts to a linear
893: system $V_Y$ of dimension $g+i+j$ of sections of
894: $\w_Y((g_Z+i+j+3)A)$.
895: 
896: Since $A$ is not a Weierstrass point of $Y$, the sequence of
897: orders at $A$ of the complete linear system
898: of sections of $\w_Y((g_Z+i+j+3)A)$ is
899:        $$0,\,1,\,\dots,\,g_Z+i+j+1,\,g_Z+i+j+3,\,
900:        g_Z+i+j+4,\,\dots,\,g+i+j+2.$$
901: Since $V_Y$ has codimension 2 in $H^0(\w_Y((g_Z+i+j+3)A))$,
902: the weight $w_A$ of $V_Y$ at $A$ satisfies
903:        \begin{equation}\label{wA2}
904:        w_A\leq 2(g_Z+i+j)+3g_Y,
905:        \end{equation}
906: with equality if and only if $V_Y=H^0(\w_Y((g_Z+i+j+1)A))$.
907: 
908: As in Case 1, using the Pl\"ucker formula, the number $N$ of
909: ramification points of $V_Y$ and $V_Z$ on $(Y-A)\cup(Z-B)$,
910: counted with their respective weights, satisfies
911:        $$N=N'+4g+4i+3j-w_A-w_B,$$
912: where
913:        $$N':=(g+i+j)(2g+i+j)+(g+i+j)(g+i+j-1)(g-1)-g-i.$$
914: As in Case 1, since $Q$ is a ramification point of $V_\eta$
915: with weight at least $g+i$, there are at most $N'$
916: ramification points of $V_\eta$, counted with their
917: respective weights, converging to $(Y-A)\cup(Z-B)$. So
918:        $$4g+4i+3j-w_A-w_B\leq 0.$$
919: On the other hand, Inequalities (\ref{wB2}) and (\ref{wA2})
920: yield
921:        $$w_A+w_B\leq 4g+4i+3j+1.$$
922: In particular, $w_A\geq 2(g_Z+i+j)+3g_Y-1$, whence
923:        $$V_Y\subset H^0(\w_Y((g_Z+i+j+2)A)).$$
924: Also, $Q$ has weight $g+i$ or $g+i+1$ in $V_\eta$.
925: Thus Statement 2 is shown.
926: Furthermore, 
927: if $i=0$ we have $w_A+w_B=4g+4i+3j$. In this case,
928: $V_Y=H^0(\w_Y((g_Z+i+j+1)A))$ and 
929: $Q$ has weight $g+i$ in $V_\eta$, showing the first 
930: part of Statement 1.
931: 
932: If $Q$ has weight $g+i+1$ in $V_\eta$, all other ramification
933: points converge to $(Y-A)\cup(Z-B)$. If $Q$ has weight $g+i$,
934: there is at most one ramification point of 
935: $V_\eta$, other than $Q$,
936: converging outside $(Y-A)\cup(Z-B)$, and that point is 
937: simple. If $R$ is that point, then
938: $\dim V_\eta(-(g+i+j)R)\leq 1$ and 
939: $V_\eta(-(g+i+j+2)R)=0$ because
940: of the simplicity of $R$.
941: 
942: Assume now that $R_0\in (Y-A)\cup(Z-B)$. Let us first
943: consider the case $R_0\in Y-A$. In this case, since,
944: from the induction hypothesis,
945: the complete linear system of sections of
946: $\w_Y((g_Z+i+j+2)A)$ has at most simple
947: ramification points, other than $A$, we have
948:        \begin{eqnarray*}
949:        h^0(\w_Y((g_Z+i+j+2)A-(g+i+j)R_0)&\leq& 1,\\
950:        h^0(\w_Y((g_Z+i+j+2)A-(g+i+j+2)R_0)&=& 0.
951:        \end{eqnarray*}
952: Thus $\dim V_Y(-(g+i+j)R_0)\leq 1$ and $V_Y(-(g+i+j+2)R_0)=0$ as
953: well. It follows, as in Case 1, that 
954: $$
955: \dim V_\eta(-(g+i+j)R)\leq 1\quad {\rm and}\quad 
956: V_\eta(-(g+i+j+2)R)=0.
957: $$ 
958: Furthermore,
959: if $i=0$, since in this case $V_Y=H^0(\w_Y((g_Z+i+j+1)A))$, all
960: the ramification points of $V_Y$ distinct from $A$ are simple.
961: Thus $R_0$ is simple in $V_Y$, and hence $R$ is simple in
962: $V_\eta$.
963: 
964: Assume now that $R_0\in Z-B$. There are two cases to consider.
965: First, assume $R_0=P_0$. Since, by induction hypothesis, the
966: complete linear system of sections of $\w_Z((g_Y+j+1)B)$ has at
967: most simple ramification points other than $B$, the weight of
968: $P_0$ as a ramification point of $V_Z$ is either $g+j$ or $g+j+1$.
969: Since $P$ has at least weight $g+j$ in $V_\eta$, and $R\neq P$,
970: the latter must hold, and $R$ must be a simple ramification point
971: of $V_\eta$. In particular, $\dim V_\eta(-(g+i+j)R)\leq 1$ and
972: $V_\eta(-(g+i+j+2)R)=0$.
973: 
974: Finally, assume $R_0\neq P_0$. Then 
975: $$
976: \dim V_Z(-(g+i+j)R_0)\leq 1\quad {\rm and} \quad 
977: V_Z(-(g+i+j+2)R_0)=0
978: $$
979: from the induction hypothesis, and
980: hence $\dim V_\eta(-(g+i+j)R)\leq 1$ and $V_\eta(-(g+i+j+2)R)=0$.
981: Thus Statement~3 is shown. Also, if $i=0$, then
982: $V_Z=H^0(\w_Z((g_Y+j+1)B))$, and, since $R_0\neq P_0$, the weight
983: of $R_0$ in $V_Z$ is equal to its weight in the complete linear
984: system of sections of $\w_Z((g_Y+j+1)B)$. By induction hypothesis,
985: this weight is one, and thus $R$ is a simple ramification point of
986: $V_\eta$. So Statement~1 is shown.\qed
987: 
988: \begin{claim}\label{ws}{\bf Corollary.} 
989: If $C$ is a general
990: smooth curve of genus $g\geq 1$, then all its
991: Weierstrass points are simple.
992: \end{claim}
993: 
994: \proof 
995: Apply Statement 1 of
996: Proposition~\ref{WiPgen} for $i_0:=0$ and $i:=0$. 
997: \qed
998: 
999: \begin{claim}\label{arbD}{\bf Proposition.} 
1000: Fix an integer $i_0\geq 0$. 
1001: Let $C$ be a general smooth curve of genus $g\geq 1$. 
1002: Then for any two distinct points $P$ and $R$ of $C$, 
1003: and any nonnegative integer $i\leq i_0$,
1004:        $$h^0(C,\omega_C((i+1)P-(g+i-2)R))=2.$$
1005: \end{claim}
1006: 
1007: \proof A line bundle of degree 2 on an elliptic curve has 
1008: (at most) 2 linearly independent sections. Thus we 
1009: may assume $g\geq 2$. Also, for $i=0$, 
1010:        $$h^0(C,\omega_C((i+1)P-(g+i-2)R)=
1011:          h^0(C,\omega_C(-(g-2)R))=2,$$
1012: since $R$ is at most a simple Weierstrass point of 
1013: $C$, a consequence of 
1014: Corollary~\ref{ws}. 
1015: So we need 
1016: only show the stated equality for integers $i>0$.
1017: 
1018: For each integer $j\geq 2$ (resp. $j\geq 1$), let 
1019: $M_j$ be the moduli space of smooth curves (resp. 
1020: let $M_{j,1}$ be the moduli space of smooth pointed 
1021: curves) of genus $j$. 
1022: Let $\ol M_j$ and $\ol M_{j,1}$ denote their 
1023: respective compactifications by stable (resp. stable, pointed) curves. 
1024: For each 
1025: positive integer $i\leq i_0$, let 
1026: $D^{(i)}\subseteq M_{g+i}$ be the subset 
1027: parameterizing curves admitting a covering 
1028: of degree at most $g+i-2$ of the projective line 
1029: totally ramified at a point. 
1030: By \cite{A}, Thm. 3.11, p. 333, the subvariety 
1031: $D^{(i)}$ is irreducible of codimension 2. Let 
1032: $\ol D^{(i)}\subset\ol M_{g+i}$ be the closure of 
1033: $D^{(i)}$. 
1034: 
1035: Let $\mu_i\:M_{g,1}\times M_{i,1}\to 
1036: \ol M_{g+i}$ be the natural map, associating to a 
1037: pair of smooth pointed curves the stable uninodal curve 
1038: which is the union of these curves identified at the 
1039: marked points. Let $E^{(i)}:=\mu_i^{-1}(\ol D^{(i)})$. 
1040: Let 
1041: $\rho_i\:E^{(i)}\to M_g$ be the natural map, 
1042: forgetting the second 
1043: pointed curve and the marked point on the first curve. 
1044: Since $C$ is general, we may assume that, 
1045: for each $i=1,\dots,i_0$, the 
1046: curve $C$ is parameterized 
1047: by a point of $M_g$ over which the fiber of 
1048: $\rho_i$ has minimum dimension. We claim this 
1049: dimension is at most $3i-3$, whence less than 
1050: $\dim M_{i,1}$. Indeed, if the dimension were 
1051: larger, then $E^{(i)}$ would have codimension at 
1052: most 1 
1053: in $M_{g,1}\times M_{i,1}$, and hence would dominate 
1054: $\ol D^{(i)}$ under $\mu_i$. So $\ol D^{(i)}$ would 
1055: be contained in the boundary $\ol M_{g+i}-M_{g+i}$, 
1056: an absurd. From the claim, for each $i=1,\dots,i_0$, 
1057: the general smooth pointed curve $(Y_i,B_i)$ of genus 
1058: $i$ is such that,
1059: for any $P\in C$, the 
1060: pair of pointed curves $((C,P),(Y_i,B_i))$ is 
1061: not parameterized by $E^{(i)}$. Consequently, the 
1062: stable uninodal curve $X_i$, union of $C$ and $Y_i$ 
1063: with $P$ and $B_i$ identified, is 
1064: parameterized by a point of 
1065: $\ol M_{g+i}$
1066: off $\ol D^{(i)}$, 
1067: for each $i=1,\dots,i_0$. 
1068: 
1069: Suppose, by contradiction, that for certain 
1070: distinct points $P$ and $Q$ of $C$, and a certain 
1071: positive integer $i\leq i_0$, we have
1072:        $$h^0(C,\w_C((i+1)P-(g+i-2)Q)\geq 3.$$
1073: Put $g':=g+i$. Since, by Riemann--Roch, 
1074: $h^0(C,\w_C((i+1)P-iQ))=g$, 
1075: there is an integer $j$ with $2\leq j<g$ 
1076: such that
1077:        $$h^0(C,\w_C((i+1)P-(g'-j)Q)=
1078:          h^0(C,\w_C((i+1)P-(g'-j-1)Q)=j+1.$$
1079: Again by Riemann--Roch, 
1080:        \begin{equation}
1081: 	 h^0(C,{\cal O}_C((g'-j)Q-(i+1)P))>
1082: 	 h^0(C,{\cal O}_C((g'-j-1)Q-(i+1)P)).
1083:        \end{equation}
1084: Thus, there is a map 
1085: $\phi\:C\lra \PP^1$ of degree $g'-j$ such that 
1086: $\phi^*(0)=(g'-j)Q$ and $\phi^*(\infty)\geq(i+1)P$. 
1087: Let $i'$ be the integer such that $i'+1$ is the 
1088: multiplicity of $P$ in 
1089: $\phi^*(\infty)$. 
1090: Then $i'\geq i$.
1091: 
1092: Set $Y:=Y_i$ and $B:=B_i$. 
1093: Since $B$ is general, $B$ is not a Weierstrass 
1094: point of $Y$. Thus, since $i'\geq i$, we have 
1095: $h^0(Y,{\cal O}_Y(i'B))<h^0(Y,{\cal O}_Y((i'+1)B))$.  
1096: So, there is a map 
1097: $\psi\:Y\lra\PP^1$ of degree $i'+1$ such that 
1098: $\psi^*(\infty)=(i'+1)B$. 
1099: 
1100: Putting together the maps 
1101: $\phi$ and $\psi$, we may construct the 
1102: covering with source $X_i$ depicted  
1103: %<below,
1104: %>
1105: in Figure~1 below,
1106: %
1107: \begin{figure}[ht]
1108: \begin{center}
1109: \includegraphics[height=5cm, angle=0]{admm.eps}
1110: \caption{The covering.}
1111: \end{center}
1112: \end{figure}
1113: %<\begin{center}
1114: %<\includegraphics[height=5cm]{admm.pdf}
1115: %<\end{center}
1116: which can be represented by a point $[X_i]$ of the 
1117: (compactification of the) Hurwitz scheme 
1118: parameterizing (pseudo)admissible 
1119: coverings of the projective line of 
1120: degree $(g'-j)$ totally ramified at a 
1121: point; see Remark \ref{rmk1}.
1122: Since 
1123: coverings of $\PP^1$ form a dense open subscheme of 
1124: this compactification, the curve 
1125: $X_i$ is limit of smooth curves equipped with a 
1126: degree-$(g'-j)$ map to the projective line totally 
1127: ramified at a 
1128: point. Since $j\geq 2$, it follows that $[X_i]$ 
1129: lies on the boundary of $D^{(i)}$,
1130: a contradiction.\qed
1131: 
1132: \begin{claim}\label{rmk1}{\bf Remark.} {\rm The 
1133: Hurwitz scheme we used in the proof of 
1134: Proposition \ref{arbD} is 
1135: mentioned in \cite{Diaz/Exc}, Section 5. It can be 
1136: constructed following the same reasoning used in the 
1137: construction of the Hurwitz scheme of (simple) 
1138: admissible coverings, given in the proof 
1139: of \cite{HaMu}, Thm. 4, p. 58. Also, the local 
1140: descriptions of both schemes are the same, given on 
1141: \cite{HaMu}, p. 62. From this description we see that 
1142: the Hurwitz scheme is equidimensional. Now, there 
1143: is a natural forgetful map from the Hurwitz scheme 
1144: to a corresponding moduli space of pointed genus-0 
1145: curves, taking a covering to its target. 
1146: This map is finite and surjective, also by 
1147: \cite{HaMu}, Thm. 4, p.~58. Since
1148: the moduli spaces of pointed genus-$0$ curves 
1149: are irreducible (see \cite{Kn} or \cite{Ke}),
1150: it follows that each irreducible 
1151: component of the Hurwitz scheme covers the target. So 
1152: coverings of $\PP^1$ form a dense open subscheme of 
1153: the Hurwitz scheme, a fact used in the proof 
1154: of Proposition \ref{arbD}.}
1155: \end{claim}
1156: 
1157: \begin{claim}\label{rmk2}{\bf Remark.} {\rm We tried 
1158: to prove Proposition \ref{arbD} using the same induction 
1159: argument used in the proof of Lemma \ref{WCQ}. 
1160: However, we could not prove the 
1161: initial step, that is, the following statement:}
1162: {\it Let $C$ be a general elliptic 
1163: curve, $Q\in C$ a general point, and $P\in C-\{Q\}$ any 
1164: point. Let $i$ and $j$ be nonnegative integers. Then the 
1165: linear system $V$ of sections of the line bundle 
1166: $\w_C((i+1)P+(j+1)Q)$ 
1167: generated by $H^0(\w_C((i+1)P))$ and $H^0(\w_C((j+1)Q))$ 
1168: satisfies $\dim V(-(i+j-1)R)=2$ for 
1169: each $R\in C-\{P,Q\}$.}
1170: \end{claim}
1171: 
1172: \section{Weierstrass divisors}\label{sec4}
1173: 
1174: \begin{claim}\label{wronski} {\rm ({\it Wronski maps})
1175: Let $C$ be a smooth curve of genus $g$. 
1176: For each integer $j\geq 0$, consider the family of
1177: linear systems $V_C(j,P)$ for $P$ varying on $C$.
1178: More precisely, let
1179: $p_1$ and $p_2$ denote the projections of $C\times C$ onto
1180: the first and second factors, and 
1181: $\Delta\subset C\times C$ the diagonal.
1182: The relative canonical bundle of $p_1$
1183: is simply the pullback $p_2^*\w_C$ of the canonical bundle
1184: $\w_C$ of $C$. For each integer $j\geq 0$, let
1185:        $${\cal L}_j:=p_2^*\w_C((j+1)\Delta),\quad
1186:        {\cal E}_j:=p_{1*}{\cal L}_j.$$
1187: Notice that, for each point $P$ of $C$,
1188: identifying $\{P\}\times C$ with $C$
1189: in the natural way,
1190: ${\cal L}_j|_{\{P\}\times C}=\w_C((j+1)P)$.
1191: Also, as $h^0(\w_C((j+1)P))=g+j$ for
1192: every $P\in C$, the sheaf ${\cal E}_j$ is a
1193: bundle of rank
1194: $g+j$ and ${\cal E}_j|_P=H^0(\w_C((j+1)P))$.
1195: 
1196: For each integer $\ell\geq 0$ and each line bundle
1197: ${\cal M}$ on $C\times C$, let
1198: ${\cal J}^\ell_{p_1}({\cal M})$
1199: be the bundle 
1200: of rank $\ell+1$ 
1201: of $p_1$-relative jets of ${\cal M}$
1202: truncated in order $\ell$. Let
1203:        $$\rho_{j,\ell}\:p_1^*{\cal E}_j\to
1204:        {\cal J}^\ell_{p_1}({\cal L}_j)$$
1205: be the map of bundles locally obtained by 
1206: differentiating 
1207: up to order $\ell$ along the fibers of $p_1$
1208: the sections of ${\cal L}_j$. We call $\rho_{j,\ell}$ 
1209: a {\it Wronski map}.
1210: 
1211: The map $\rho_{j,g+j-1}$ is a map of bundles of the
1212: same rank. Taking determinants, we get a section $z_j$
1213: of the line bundle
1214:         $$\bigwedge^{g+j}{\cal J}^{g+j-1}_{p_1}({\cal L}_j)
1215:         \otimes\bigwedge^{g+j}p_1^*{\cal E}_j^{\vee},$$
1216: which is naturally isomorphic, using the truncation
1217: sequence of the bundles of jets, to
1218:         $$p_2^*\w_C((j+1)\Delta)^{\otimes g+j}\otimes
1219:         p_2^*\w_C^{\otimes (g+j)(g+j-1)/2}\otimes
1220:         \bigwedge^{g+j}p_1^*{\cal E}_j^{\vee},$$
1221: or more simply to
1222:         $$p_2^*\w_C^{\otimes (g+j)(g+j+1)/2}
1223:         \Big((g+j)(j+1)\Delta\Big)\otimes\bigwedge^{g+j}
1224:         p_1^*{\cal E}_j^{\vee}.$$}
1225: \end{claim}
1226: 
1227: \begin{claim}\label{jWd} 
1228: {\rm ({\it Weierstrass divisors.}) 
1229: Keep the notation used in Subsection \ref{wronski}.
1230: Let $Z_j\subseteq C\times C$ denote the zero scheme 
1231: of $z_j$. The section $z_j$ is a relative wronskian.
1232: More precisely, for each $P\in C$, on $\{P\}\times C$,
1233: identified with $C$ in the natural way,
1234: the section $z_j$ restricts to the
1235: wronskian of the linear system $V_C(j,P)$. Hence, $Z_j$
1236: consists of the pairs $(P,Q)\in C\times C$ such that
1237: $V_C(j,P)$ ramifies at $Q$. Now, since $z_j$ is nonzero,
1238: being so on each fiber, $Z_j$ is a Cartier divisor.
1239: By Proposition~\ref{prop5}, the divisor $Z_j$ intersects
1240: each fiber $\{P\}\times C$ at $(P,P)$ with multiplicity
1241: $g+\wgt(P)$, where $\wgt(P)$ is the weight of $P$ in the
1242: canonical system of $C$. Thus $Z_j$ contains $\Delta$ with
1243: multiplicity exactly $g$. Let
1244:         $$W_j:=Z_j-g\Delta.$$
1245: Then $W_j$ is, set-theoretically, the locus of pairs
1246: $(P,Q)\in C\times C$ such that either $P=Q$ and $P$ is a
1247: Weierstrass point of $C$, or $P\neq Q$ and $Q$ is a
1248: ramification point of $V_C(j,P)$. We call $W_j$ the 
1249: {\it $j$-th Weierstrass divisor of $C$.}}
1250: \end{claim}
1251: 
1252: \begin{claim}\label{prop7}{\bf Proposition.} 
1253: Let $C$ be a smooth curve of
1254: genus $g\geq 1$ and $j$ a nonnegative integer. Let 
1255: $\Delta$ be the diagonal of $C\times C$, and 
1256: $p_1$ and $p_2$ the projections of $C\times C$ 
1257: onto the indicated factors. Let 
1258: $\w_C$ be the canonical bundle of $C$, and set 
1259: $K_\ell:=c_1(p_\ell^*\w_C)$ for $\ell=1,2$. 
1260: Let $W_j\subseteq C\times C$ be 
1261: the $j$-th Weierstrass divisor of $C$. 
1262: Then its class $[W_j]$ in the Chow group of
1263: $C\times C$ satisfies
1264:         \be
1265:         [W_j]=\frac{1}{2}(g+j)(g+j+1)K_2+j(g+j+1)[\Delta]
1266:         +\frac{1}{2}j(j+1)K_1.\label{eq:17}
1267:         \ee
1268: \end{claim}
1269: 
1270: \proof Use 
1271: the notation in Subsections \ref{wronski} 
1272: and \ref{jWd}. Since $W_j=Z_j-g\Delta$, and $Z_j$ is the
1273: zero scheme of a section of the line bundle
1274:         $$p_2^*\w_C^{\otimes (g+j)(g+j+1)/2}
1275:         \Big((g+j)(j+1)\Delta\Big)\otimes\bigwedge^{g+j}
1276:         p_1^*{\cal E}_j^{\vee},$$
1277: we get
1278:         \be
1279:     [W_j]=\frac{1}{2}(g+j)(g+j+1)K_2+j(g+j+1)[\Delta]-
1280:         p_1^*c_1({\cal E}_j).\label{eq:67}
1281:         \ee
1282: To finish, we need only show that
1283:         \be
1284:         c_1({\cal E}_j)=-\frac{1}{2}j(j+1)c_1(\w_C).
1285:     \label{eq:c1Ei}
1286:     \ee
1287: 
1288: We show (\ref{eq:c1Ei}) by induction on $j$.
1289: First of all,
1290:         $${\cal E}_0=p_{1*}p_2^*\w_C=
1291:         H^0(\w_C)\otimes{\cal O}_C.$$
1292: Since ${\cal E}_0$ is free, $c_1({\cal E}_0)=0$.
1293: 
1294: Assume now that $j>0$ and
1295: $c_1({\cal E}_{j-1})=-(j(j-1)/2)c_1(\w_C)$.
1296: Consider the natural short exact sequence
1297:         $$0\to p_2^*\w_C(j\Delta)\to p_2^*\w_C((j+1)\Delta)
1298:         \to p_2^*\w_C((j+1)\Delta)|_{\Delta}\to 0.$$
1299: Since $H^1(\w_C(jP))=0$ for each $P\in C$,
1300: applying $p_{1*}$ to the sequence above,
1301: we get the exact sequence
1302:         $$0\to{\cal E}_{j-1}\to{\cal E}_j\to
1303:         p_{1*}p_2^*\w_C((j+1)\Delta)|_{\Delta}\to 0.$$
1304: Now, $p_\ell|_\Delta$ is an isomorphism for $\ell=1,2$. So
1305: $p_{1*}p_2^*\w_C|_\Delta=\w_C$. In addition,
1306: $p_{1*}{\cal O}_{C\times C}(-\Delta)|_\Delta=\w_C$. Thus
1307:         \begin{eqnarray*}
1308:     c_1({\cal E}_j)&=&c_1({\cal E}_{j-1})+
1309:         c_1(p_{1*}p_2^*\w_C((j+1)\Delta)|_{\Delta})\\
1310:     &=&-(j(j-1)/2)c_1(\w_C)+(1-(j+1))c_1(\w_C)\\
1311:     &=&-(j(j+1)/2)c_1(\w_C),
1312:     \end{eqnarray*}
1313: as claimed.\qed
1314: 
1315: \begin{claim}\label{lem}{\bf Proposition.} 
1316: Let $C$ be a general
1317: smooth curve of genus $g\geq 1$ and $j$ a nonnegative
1318: integer. Let $W_j\subseteq C\times C$ be the 
1319: $j$-th Weierstrass divisor of $C$. Then $W_j$ 
1320: is nonsingular and intersects 
1321: the diagonal $\Delta$ transversally, at the pairs 
1322: $(P,P)$ such that $P$ is a
1323: Weierstrass point of $C$.
1324: \end{claim}
1325: 
1326: \proof Let us show first that $W_j$ intersects 
1327: $\Delta$ transversally. 
1328: As pointed out in Subsection \ref{jWd}, the intersection 
1329: $W_j\cap\Delta$ is,
1330: set-theoretically, the set of pairs $(P,P)$ such that
1331: $P$ is a Weierstrass point of $C$. As $C$ is general,
1332: by Corollary~\ref{ws},
1333: all its Weierstrass points are simple, and number
1334: $g^3-g$ by Pl\"ucker Formula.
1335: Now, since the intersection $W_j\cap\Delta$ is
1336: finite, the number of points of intersection, weighted
1337: by their intersection multiplicities, is equal to 
1338: the degree
1339: of the product $[W_j][\Delta]$. Using the notation 
1340: and Formula (\ref{eq:17}) of Proposition \ref{prop7}, 
1341: and using the Formulas
1342:        \be
1343:        \int_{C\times C}K_2[\Delta]=
1344:        \int_{C\times C}K_1[\Delta]=
1345:        -\int_{C\times C}[\Delta]^2=2g-2
1346:        \label{eq:kd}
1347:        \ee
1348: and
1349:        \be
1350:        \int_{C\times C}K_1K_2=4(g-1)^2,
1351:        \label{eq:kk}
1352:        \ee
1353: we get
1354:        $$\int_{C\times C}[W_j][\Delta]=
1355:        (g+j)(g+j+1)(g-1)-2j(g+j+1)(g-1)
1356:         +j(j+1)(g-1),$$
1357: which is exactly $g^3-g$. Thus the intersection
1358: multiplicities are all one.
1359: 
1360: As a corollary of the transversal intersection, 
1361: $W_j$ is nonsingular at its points on $\Delta$. So, 
1362: let now $(P,Q)\in W_j$ for $P$ and $Q$ distinct, and 
1363: let us show that $W_j$ is nonsingular at 
1364: $(P,Q)$ as well. 
1365: 
1366: Let $J:={\rm Pic}^{g-1}(C)$, the component of the 
1367: Picard scheme of $C$ parameterizing line bundles of 
1368: degree $g-1$. Let $\Theta\subset J$ be the theta divisor, 
1369: parameterizing line bundles with nontrivial global 
1370: sections. Let
1371:          $$\mu\: C\times C \to J$$
1372: be the map taking a pair $(R,S)$ to the point of 
1373: $J$ representing the bundle $\w_C((j+1)R-(g+j)S)$. 
1374: 
1375: We claim that $\mu(W_j)\subseteq\Theta$. Indeed, let 
1376: $C^{(3)}:=C\times C\times C$, and denote by $p_{1,2}$ 
1377: and $p_3$ the projection maps of 
1378: $C^{(3)}$ onto the indicated factors. Let 
1379: $\Delta_{1,3}$ and $\Delta_{2,3}$ be the indicated 
1380: diagonals of $C^{(3)}$. Set
1381:        $${\cal F}:=p_3^*\w_C((j+1)\Delta_{1,3}
1382:          -(g+j)\Delta_{2,3}).$$
1383: Recall the notation of Subsection \ref{wronski}. From 
1384: the construction of $\Theta$, to show that 
1385: $\mu(W_j)\subseteq\Theta$, it is enough to show that 
1386: the Wronski map $\rho_{j,g+j-1}$ represents 
1387: universally the cohomology of ${\cal F}$ or, put more 
1388: simply, that $\rho_{j,g+j-1}$ can be viewed as a 
1389: presentation of the right derived image 
1390: $R^1(p_{1,2})_*{\cal F}$. 
1391: 
1392: Let ${\cal G}:=p_3^*\w_C((j+1)\Delta_{1,3})$. Then 
1393: ${\cal F}\subseteq{\cal G}$. From the definition of 
1394: the Wronski map $\rho_{j,g+j-1}$, we get that 
1395: $\rho_{j,g+j-1}$ is the image under $(p_{1,2})_*$ of 
1396: the quotient map ${\cal G}\to{\cal G}/{\cal F}$. Thus, 
1397: the map $\rho_{j,g+j-1}$ 
1398: is the first map
1399: in the following 
1400: piece of the long derived sequence of 
1401: $0\to{\cal F}\to{\cal G}\to{\cal G}/{\cal F}\to 0$ 
1402: under $(p_{1,2})_*$:
1403:         $$(p_{1,2})_*{\cal G}\to
1404:           (p_{1,2})_*({\cal G}/{\cal F})\to
1405:           R^1(p_{1,2})_*{\cal F}\to 
1406: 	  R^1(p_{1,2})_*{\cal G}.$$
1407: Now, a fiberwise analysis shows that 
1408: $R^1(p_{1,2})_*{\cal G}=0$. Thus $\rho_{j,g+j-1}$ 
1409: is a presentation for $R^1(p_{1,2})_*{\cal F}$, finishing 
1410: the proof that $\mu(W_j)\subseteq\Theta$. 
1411: 
1412: Let ${\cal L}:=\w_C((j+1)P-(g+j)Q)$, and denote 
1413: by $[{\cal L}]$ the point of $J$ representing ${\cal L}$. 
1414: Since $(P,Q)\in W_j$, we have $[{\cal L}]\in\Theta$. 
1415: By Proposition \ref{WiPgen}, $h^0(C,{\cal L})=1$. Thus, 
1416: it follows from \cite{acgh}, Prop. (4.2), p. 189, that 
1417: $[{\cal L}]$ is a nonsingular point of $\Theta$. 
1418: Furthermore, identifying the cotangent space of 
1419: $J$ at $[{\cal L}]$ with $H^0(C,\w_C)$, the cotangent 
1420: space of $\Theta$ at $[{\cal L}]$ is the quotient by 
1421: the subspace $H^0(C,\w_C(-F))$, where $F$ is the 
1422: unique effective divisor of $C$ such that 
1423: ${\cal L}={\cal O}_C(F)$. 
1424: 
1425: Identifying the cotangent space of $C\times C$ at 
1426: $(P,Q)$ with $\w_C|_P\oplus\w_C|_Q$, the induced map 
1427: of cotangent spaces 
1428: $d\mu^*\: T^*_{J,[{\cal L}]}\to T^*_{C\times C,(P,Q)}$ 
1429: is equivalent 
1430: to the evaluation map,
1431:        $$\epsilon\: H^0(C,\w_C)\to \w_C|_P\oplus\w_C|_Q.$$
1432: We claim that $\epsilon(H^0(C,\w_C(-F)))\neq 0$. 
1433: Indeed, if that were not the case, we would have 
1434: $H^0(C,\w_C(-F))=H^0(C,\w_C(-F-P-Q))$, that is,
1435:        $$h^0(C,{\cal O}_C((g+j-1)Q-(j+2)P))
1436:          =h^0(C,{\cal O}_C((g+j)Q-(j+1)P)).$$
1437: By the Riemann--Roch theorem,
1438:        $$h^0(C,{\cal O}_C((g+j)Q-(j+1)P))=
1439:          h^0(C,{\cal L})=1,$$
1440: and thus, also by the Riemann--Roch theorem,
1441:        $$h^0(C,\w_C((j+2)P-(g+j-1)Q))=3.$$
1442: However, this contradicts Proposition \ref{arbD}. 
1443: 
1444: Since $\mu(W_j)\subseteq\Theta$, the image of 
1445: $\epsilon(H^0(C,\w_C(-F)))$ in the cotangent space of 
1446: $W_j$ at $(P,Q)$ is zero. Since 
1447: $\epsilon(H^0(C,\w_C(-F)))\neq 0$, that cotangent space 
1448: is a proper quotient of the cotangent space of 
1449: $C\times C$ at $(P,Q)$, and thus has dimension at most 1. 
1450: Since $W_j$ is a divisor, it follows that $W_j$ is 
1451: nonsingular at $(P,Q)$.\qed
1452: 
1453: \section{Special ramification classes}\label{sec5}
1454: 
1455: \begin{claim}\label{srl} 
1456: {\rm ({\it Special ramification loci.})
1457: Let $C$ be a smooth curve of genus $g\geq 1$. 
1458: For each nonnegative integer $i$, consider the 
1459: following loci in $C\times C$:
1460: \begin{enumerate}
1461: \item The locus $D^+_i$ of pairs $(P,Q)\in C\times C$
1462: such that
1463:        $$(g+i-1)Q-(i+1)P$$
1464: is linearly equivalent to an effective divisor.
1465: \item The locus $E^+_i$ of pairs $(P,Q)\in C\times C$ such
1466: that
1467:        $$(g+i+1)Q-(i+1)P$$
1468: is linearly equivalent to a moving effective divisor.
1469: \item The locus $SW^+_i$ of pairs
1470: $(P,Q)\in C\times C$ such 
1471: that
1472: $Q$ is a special ramification
1473: point of $V_C(i,P)$.
1474: \end{enumerate}
1475: We claim that, set-theoretically,
1476:        \be
1477:        SW^+_i=D^+_i\cup E^+_i.\label{eq:swde}
1478:        \ee
1479: 
1480: Indeed, by the Riemann--Roch Theorem, for a pair
1481: $(P,Q)\in C\times C$, the divisor
1482: $(g+i-1)Q-(i+1)P$ is linearly equivalent to an
1483: effective one if and only if
1484:         \be
1485:     h^0(\w_C((i+1)P-(g+i-1)Q))\geq 2,\label{eq:6b3}
1486:     \ee
1487: while $(g+i+1)Q-(i+1)P$ is linearly equivalent to a
1488: moving effective divisor if and only if
1489:         \be
1490:     h^0(\w_C((i+1)P-(g+i+1)Q))\geq 1.\label{eq:6b4}
1491:     \ee
1492: At any rate, if $(P,Q)\in D^+_i\cup E^+_i$, then $Q$ is a
1493: special ramification point of $V_C(i,P)$, that is,
1494: $(P,Q)\in SW^+_i$
1495: 
1496: On the other hand, let
1497: $(P,Q)\in C\times C-(D^+_i\cup E^+_i)$. Then
1498:         \begin{eqnarray*}
1499:     &&h^0(\w_C((i+1)P-(g+i-1)Q))=1,\\
1500:     &&h^0(\w_C((i+1)P-(g+i+1)Q))=0.
1501:     \end{eqnarray*}
1502: So, either $Q$ is an ordinary or a simple ramification
1503: point of $V_C(i,P)$, that is, $(P,Q)\not\in SW^+_i$.
1504: 
1505: Let $\Delta$ be the diagonal subscheme of $C\times C$.
1506: Notice that $E_i^+\cap\Delta$ consists of the pairs
1507: $(P,P)$ such that $P$ is a Weierstrass point of $C$.
1508: However, if $g>1$, both $D_i^+$ and $SW_i^+$ contain
1509: $\Delta$. (If $g=1$, then 
1510: $D_i^+=E_i^+=SW_i^+=\emptyset$.)
1511: 
1512: Let $D_i$, $E_i$ and $SW_i$ be the loci of points in
1513: $D_i^+$, $E_i^+$ and $SW_i^+$ that lie off $\Delta$. Of
1514: course, Expression (\ref{eq:swde}) implies
1515: $SW_i=D_i\cup E_i$. Our Proposition~\ref{A2CC} claims
1516: that, if $C$ is general, then
1517: $SW_i=D_i\cup E_i$ holds in a more
1518: refined way, in the cycle group of $C\times C$. Before
1519: stating it, we need to endow $D_i$, $E_i$ and $SW_i$
1520: with natural subscheme structures.}
1521: \end{claim}
1522: 
1523: \begin{claim}\label{srs} 
1524: {\rm ({\it Special ramification schemes}) 
1525: Keep the notation of Subsection \ref{srl}, and recall 
1526: that of Subsections \ref{wronski} and \ref{jWd}. 
1527: Notice that the subsets 
1528: $D_i^+$ and $E_i^+$ are the supports
1529: of the degeneracy schemes of $\rho_{i,g+i-2}$ and
1530: $\rho_{i,g+i}$, respectively. So we may give 
1531: $D_i^+$ and $E_i^+$ the
1532: corresponding scheme structures.
1533: Give $D_i$ and $E_i$ the
1534: corresponding open subscheme structures. We say that 
1535: $D_i$ and $E_i$ are the {\it $i$-th special ramification 
1536: schemes of type Diaz and Cukierman}, respectively. Call 
1537: $E_i^+$ the {\it $i$-th expanded special ramification 
1538: scheme of type Cukierman}.
1539: 
1540: In addition, differentiating along the fibers of $p_1$ a 
1541: section of ${\cal O}_{C\times C}(Z_i)$ defining $Z_i$, 
1542: we obtain a section of 
1543: ${\cal J}^1_{p_1}({\cal O}_{C\times C}(Z_i))$, 
1544: well-defined modulo $\CC^*$. By functoriality, its zero 
1545: scheme contains a pair $(P,Q)$ if and only if $Q$ is a 
1546: special Weierstrass point of $V_C(i,P)$. Thus the zero 
1547: scheme gives a scheme structure for
1548: $SW_i^+$. Give $SW_i$ the induced open subscheme 
1549: structure. We say that $SW_i$ is the {\it $i$-th special 
1550: ramification scheme of $C$}.
1551: 
1552: Now, $Z_i=W_i+g\Delta$. As done 
1553: for $Z_i$, we can differentiate along
1554: the fibers of $p_1$ a section of ${\cal O}_{C\times C}(W_i)$
1555: defining $W_i$ to obtain a section of ${\cal J}^1_{p_1}({\cal
1556: O}_{C\times C}(W_i))$. Its zero scheme $S$ coincides with the
1557: scheme $SW_i$ off $\Delta$, because $Z_i$ coincides with $W_i$
1558: there. Moreover, if $C$ is general, then $S$ does not intersect
1559: $\Delta$, and hence $S=SW_i$ scheme-theoretically. Indeed, let $P$
1560: be a point of $C$. If $(P,P)\in W_i$, then $P$ is a Weierstrass
1561: point of $C$. Moreover, as $C$ is general, by Corollary~\ref{ws},
1562: the point $P$ is a simple Weierstrass point. So, it follows from
1563: Proposition \ref{prop5} that $W_i$ intersects the fiber
1564: $\{P\}\times C$ transversally at $(P,P)$. Thus the derivative
1565: along $\{P\}\times C$ of a section defining $W_i$ does not vanish
1566: at $(P,P)$. So $S\cap\Delta=\emptyset$.}
1567: \end{claim}
1568: 
1569: \begin{claim}\label{matrices}{\bf Lemma.}
1570: Let ${\cal O}$ be a local 
1571: ring, and $r$ a nonnegative integer. Let $M$ be a matrix 
1572: with $r+2$ rows and $r+1$ columns and entries in 
1573: ${\cal O}$. Let $M_1$ and $M_2$ be the submatrices 
1574: obtained from $M$ by removing the last row, and
1575: the last two rows, respectively. Assume that the 
1576: matrix obtained from $M_1$ by taking residues 
1577: has rank at least $r$. Let $z$
1578: denote the determinant of $M_1$. Then there are 
1579: $u,v\in{\cal O}$ such that
1580: \begin{enumerate}
1581:     \item $(z,u)$ is the ideal of all maximal minors 
1582:       of $M_2$,
1583:     \item $(z,v)$ is the ideal of all maximal minors 
1584:       of $M$,
1585:     \item $(z,uv)$ is the ideal generated by the two 
1586:       maximal minors of $M$ obtained by removing each 
1587:       of the last two rows.
1588: \end{enumerate}
1589: \end{claim}
1590: 
1591: \proof We may write $M$ in the form
1592:         $$M=\left[\matrix{A&a&b\cr c&f_1&f_2
1593:         \cr d&g_1&g_2\cr e&h_1&h_2}\right],$$
1594: where $A$ is a square matrix of size $r-1$, where $a$ 
1595: and $b$ are column vectors of 
1596: size
1597: $r-1$, where 
1598: $c$, $d$ and $e$ are row vectors of 
1599: size
1600: $r-1$, 
1601: and where $f_1$, $f_2$, $g_1$, $g_2$,
1602: $h_1$ and $h_2$ are elements of ${\cal O}$.
1603: 
1604: Let $I$ and $J$ be the ideals of ${\cal O}$ generated,
1605: respectively, by all maximal minors of the submatrices
1606:         $$M_2=\left[\matrix{A&a&b\cr c&f_1&f_2}\right]
1607:         \quad{\rm and }\quad M=\left[
1608:         \matrix{A&a&b\cr c&f_1&f_2\cr d&g_1&g_2\cr
1609:     e&h_1&h_2}\right].$$
1610: Also, let $K\subseteq{\cal O}$ be the ideal generated by 
1611: the determinants of the square submatrices
1612:         $$M_1=\left[
1613:         \matrix{A&a&b\cr c&f_1&f_2\cr d&g_1&g_2}
1614:         \right]\quad{\rm and }\quad
1615:         M'_1:=\left[\matrix
1616: 	{A&a&b\cr c&f_1&f_2\cr e&h_1&h_2}
1617:         \right].$$
1618: Notice that the determinant of the first matrix is $z$.
1619: 
1620: From the hypothesis, the matrix obtained from $M_2$ 
1621: by taking residues has rank at least $r-1$. Thus, 
1622: performing row and column operations on $M$, including 
1623: column and row exchanges, we may, without changing the 
1624: ideals $I$, $J$ and $K$, assume that $A$ is the 
1625: identity matrix, $a=b=0$ and $c=d=e=0$. Then
1626: $z=f_1g_2-f_2g_1$ and
1627:     \begin{eqnarray*}
1628:     I&=&(f_1,\, f_2),\\
1629:     J&=&(f_1g_2-f_2g_1,\,f_1h_2-f_2h_1,\,g_1h_2-g_2h_1),\\
1630:     K&=&(f_1g_2-f_2g_1,\, f_1h_2-f_2h_1).
1631:     \end{eqnarray*}
1632: 
1633: Now, since the matrix obtained from $M_1$ 
1634: by taking residues has rank at least $r$, at least one 
1635: among $f_1,\, f_2,\, g_1,\, g_2$ is invertible.
1636: 
1637: If $f_1$ is invertible, then
1638:         $$g_1h_2-g_2h_1=(g_1/f_1)(f_1h_2-f_2h_1)-
1639:         (h_1/f_1)(f_1g_2-f_2g_1).$$
1640: Thus, the lemma holds for $u=1$ and $v=f_1h_2-f_2h_1$. 
1641: The case where $f_2$ is invertible is similar.
1642: 
1643: If $g_1$ is invertible, then
1644:         \begin{eqnarray*}
1645:         (f_1h_2-f_2h_1)&=&(f_1/g_1)(g_1h_2-g_2h_1)+
1646:         (h_1/g_1)(f_1g_2-f_2g_1),\\
1647:     f_2&=&(g_2/g_1)f_1-(1/g_1)(f_1g_2-f_2g_1).
1648:         \end{eqnarray*}
1649: Thus the lemma holds for $u=f_1$ and $v=g_1h_2-g_2h_1$. 
1650: A similar analysis holds if $g_2$ is invertible.\qed
1651: 
1652: \begin{claim}\label{A2CC}{\bf Proposition.}
1653: Let $C$ be a general
1654: smooth curve of genus $g\geq 1$ and $i$ a nonnegative
1655: integer. Let $\Delta$ be the diagonal of $C\times C$ and 
1656: $W_i$ the $i$-th Weierstrass divisor. 
1657: Let $SW_i$ be the $i$-th special ramification 
1658: scheme, and $D_i$ and $E_i$ the $i$-th special 
1659: ramification schemes of type Diaz and Cukierman, 
1660: respectively. Let $E_i^+$ be the $i$-th expanded 
1661: special ramification scheme of type Cukierman. 
1662: Then these ramification schemes are finite and satisfy, 
1663: in the cycle group of $C\times C$:
1664:         $$[SW_i]=[D_i]+[E_i]\quad{\rm and }\quad
1665:         [E_i^+]=[E_i]+(g+1)[W_i\cap\Delta].$$
1666: \end{claim}
1667: 
1668: \proof Since $C$ is general, by Statement 1 of
1669: Proposition~\ref{WiPgen}, the set $SW_i$ is finite
1670: for each $i\geq 0$. Thus, so are $D_i$ and $E_i$
1671: by Expression~(\ref{eq:swde}). It follows that
1672: $E_i^+$ is finite, because $E_i^+\cap\Delta$ is the
1673: set of points $(P,P)$ such that $P$ is Weierstrass, whence
1674: is finite.
1675: 
1676: Recall the notation of Subsections \ref{wronski}, 
1677: \ref{jWd}, \ref{srl} and \ref{srs}. 
1678: Set $r:=g+i-1$. Both equalities can be proved locally. 
1679: Thus, let $(P,Q)\in C\times C$ and ${\cal O}$ be the 
1680: local ring of $C\times C$ at $(P,Q)$. 
1681: As a map of ${\cal O}$-modules, 
1682: $\rho_{i,r+1}$ is given by a matrix $M$ of the form 
1683: described in the proof of Lemma \ref{matrices}. Let 
1684: us use the notation described in the statement of 
1685: that lemma.
1686: 
1687: Let $K\subseteq{\cal O}$ define $SW_i^+$. Then 
1688: $K=(z,z')$, where $z$ (resp. $z'$) 
1689: is the maximal minor obtained 
1690: from $M$ by removing the last (resp. last but one) row. 
1691: Notice that, 
1692: from the nature of $M$ as a ``wronskian matrix'', $z'$ is 
1693: also the derivative of $z$ along $p_1$. Let 
1694: $I$ and $J$ be the ideals of ${\cal O}$ defining $D_i^+$ 
1695: and $E_i^+$, respectively. Then $I$ and $J$ are the 
1696: ideals of all the
1697: maximal minors of $M_2$ and $M$, respectively.
1698: 
1699: Now, since $C$ is a general curve, by Statement 2 of 
1700: Proposition \ref{WiPgen},
1701:         $$h^0(\w_C((i+1)P-(g+i)Q))\leq 1.$$
1702: This translates in the matrix obtained from $M_1$ by 
1703: evaluating at $(P,Q)$ having rank at least $r$. 
1704: Applying Lemma \ref{matrices},
1705: there are $u,v\in{\cal O}$ such
1706:     $$I=(z,u),\quad J=(z,v),\quad K=(z,uv).$$
1707: Now, since $E_i^+$ is finite-dimensional and 
1708: $C\times C$ is smooth, the sequence $z,v$ is regular. 
1709: The same holds for the sequence $z,u$ 
1710: if $P\neq Q$. It follows 
1711: that $[SW_i]=[D_i]+[E_i]$.
1712: 
1713: The second equality in the statement of the
1714: proposition is obvious off $\Delta$. Thus, assume 
1715: $Q=P$. Since $E_i^+\cap\Delta=W_i\cap\Delta$, we may 
1716: also assume that $P$ is a Weierstrass point of $C$. 
1717: 
1718: Let $t$ be a local parameter of $C$ at $P$, and 
1719: $t_1,t_2\in{\cal O}$ be its pullbacks with respect to 
1720: the projections $p_1$ and $p_2$. Then $t:=t_2-t_1$ is a 
1721: local equation for $\Delta$. As we saw in Subsection 
1722: \ref{srs}, we have $z=t^gw$, where
1723: $w\in{\cal O}$ defines $W_i$, and is not divisible by 
1724: $t$. Letting $\partial$ denote the 
1725: derivative with respect to $t_2$, we have
1726:        $$z'=\partial z=\partial(t^gw)=
1727:        gt^{g-1}w+t^g\partial w.$$
1728: Thus $t^{g-1}$ divides $z$ and $z'$, and hence each 
1729: element of $K$, in particular $uv$. Since $E_i^+$ is 
1730: finite, $t$ does not divide $v$, and hence $t^{g-1}|u$. 
1731: Let $L:=t^{1-g}K$. Then there are two expressions for 
1732: $L$:
1733:        \be
1734:        L=(tw,uv/t^{g-1})\quad{\rm and }\quad
1735:        L=(tw,gw+t\partial w).
1736:        \label{I}
1737:        \ee
1738: Since $W_i\cap\Delta$ is finite, the 
1739: sequences $gw+t\partial w, t$ and $w, t$ are 
1740: regular. Thus, from the second expression for $L$ above, 
1741: we get
1742:        $$\ell({\cal O}/L)=2\ell({\cal O}/(t,w))+
1743:        \ell({\cal O}/(w,\partial w)).$$
1744: Now, $\ell({\cal O}/(w,\partial w))=0$ because $w$ and 
1745: $\partial w$ cut out $SW_i$, and $SW_i$ does not meet 
1746: $\Delta$. Also, by Lemma \ref{lem}, $W_i$ intersects 
1747: $\Delta$ transversally. Thus
1748: $\ell({\cal O}/(t,w))=1$, and hence $\ell({\cal O}/L)=2$.
1749: 
1750: Now, since the sequence $z, v$ is regular, and 
1751: $z=t^gw$, also the sequence $tw,v$ is regular. Thus, 
1752: from the first expression for $L$ in (\ref{I}), 
1753: we get
1754:        $$\ell({\cal O}/L)=\ell({\cal O}/(tw,u/t^{g-1}))+
1755:          \ell({\cal O}/(tw,v)),$$
1756: and whence $\ell({\cal O}/(tw,v))\leq 2$. Since 
1757: ${\cal O}$ is regular, and the sequence $tw,v$ is 
1758: regular, so is the sequence $v,w$. Thus
1759:        $$\ell({\cal O}/(tw,v))=\ell({\cal O}/(t,v))+
1760:          \ell({\cal O}/(w,v)).$$
1761: Since $E_i^+$ contains $(P,P)$, the function $v$ is zero 
1762: on $(P,P)$. Thus, since also $t$ and $w$ vanish on 
1763: $(P,P)$, we get 
1764: $\ell({\cal O}/(t,v))=\ell({\cal O}/(w,v))=1$. So, the 
1765: multiplicity of $E_i^+$ at $(P,P)$ is
1766:        $$\ell({\cal O}/(z,v))=
1767:        g\ell({\cal O}/(t,v))+\ell({\cal O}/(w,v))
1768:        =(g+1).$$
1769: Since, by Lemma \ref{lem}, the multiplicity of 
1770: $W_i\cap\Delta$ at $(P,P)$ is 1, we are done.\qed
1771: 
1772: \begin{claim}\label{prop07}{\bf Proposition.}
1773: Let $C$ be a general
1774: smooth curve of genus $g\geq 1$ and $i$ a nonnegative
1775: integer. Let $SW_i$ be the $i$-th special ramification 
1776: scheme of $C$. Then
1777:         \be
1778:     \int_{C\times C}[SW_i]=2ig(g-1)\Big((i+2)(g+i)^2
1779:     +2(g+i)+2\Big).
1780:     \label{eq:me1}
1781:     \ee
1782: \end{claim}
1783: 
1784: \proof Recall the notation of Subsections 
1785: \ref{wronski}, \ref{jWd}, \ref{srl} and \ref{srs}. 
1786: Since $C$ is general, $SW_i$ is finite. Also,
1787: $SW_i$ is the zero scheme of a section of the
1788: rank-2 bundle
1789: ${\cal J}^1_{p_1}({\cal O}_{C\times C}(W_i))$. Thus
1790: its class in
1791: the Chow group of $C\times C$ satisfies
1792:         $$[SW_i]=c_2({\cal J}^1_{p_1}
1793:         ({\cal O}_{C\times C}(W_i))).$$
1794: Using the truncation sequence for bundles of jets, we get
1795:         $$[SW_i]=[W_i](c_1(p_2^*\w_C)+[W_i]).$$
1796: Now, $c_1(p_2^*\w_C)=K_2$. Using Expression
1797: (\ref{eq:17}) for $j=i$, and taking into account
1798: that $K_\ell^2=0$ for $\ell=1,2$, we get
1799:         \begin{eqnarray*}
1800:     [SW_i]&=&i(g+i+1)\Big((g+i)^2+g+i+1\Big)K_2[\Delta]\\
1801:     &+&\frac{1}{2}i(i+1)\Big((g+i)^2+g+i+1\Big)K_1K_2\\
1802:     &+&i^2(g+i+1)^2[\Delta]^2+i^2(g+i+1)(i+1)K_1[\Delta].
1803:     \end{eqnarray*}
1804: Using Formulas (\ref{eq:kd}) and (\ref{eq:kk}), we get
1805:         \begin{eqnarray*}
1806:     \int_{C\times C}[SW_i]&=&i(g+i+1)
1807:     \Big((g+i)^2+g+i+1\Big)(2g-2)\\
1808:     &+&\frac{1}{2}i(i+1)\Big((g+i)^2+g+i+1\Big)4(g-1)^2\\
1809:     &-&i^2(g+i+1)^2(2g-2)+i^2(g+i+1)(i+1)(2g-2).
1810:     \end{eqnarray*}
1811: Simplifying, we get the claimed formula.\qed
1812: 
1813: \begin{claim}\label{thm6DiazCuk}{\bf Theorem.}
1814: Let $C$ be a general smooth curve of genus $g\geq 1$,
1815: and $i$ a nonnegative integer. Let $D_i$ and $E_i$ be the 
1816: $i$-th special ramification schemes of type Diaz and 
1817: Cukierman, respectively. Then $D_i$ and $E_i$ are 
1818: reduced, and
1819:       \be
1820:       \int_{C\times C}[D_i]=g(g-1)
1821:       \Big((g+i-1)^2(i+1)^2-(g-1)^2\Big)
1822:       \label{eq:ex0}
1823:       \ee
1824: and
1825:       \be
1826:       \int_{C\times C}[E_i]=g(g-1)
1827:       \Big((g+i+1)^2(i+1)^2-(g+1)^2\Big).
1828:       \label{eq:ex6b0}
1829:       \ee
1830: \end{claim}
1831: 
1832: \proof Recall the notation of 
1833: Subsections \ref{wronski}, \ref{jWd}, \ref{srl} and 
1834: \ref{srs}. We will 
1835: first compute the degrees of $D_i$ and $E_i$. 
1836: First of all, since $E_i^+$ is finite, and
1837: is the degeneracy scheme of $\rho_{i,g+i}$, applying
1838: Porteous formula (\cite{Fu}, Thm. 14.4, p. 254), 
1839: we get the
1840: following expression for the class $[E_i^+]$ in the
1841: Chow group of $C\times C$:
1842:       $$[E_i^+]=c_2({\cal J}^{g+i}_{p_1}({\cal L}_i)-
1843:       p_1^*{\cal E}_i).$$
1844: Now, $c_2({\cal E}_i)=c_1({\cal E}_i)^2=0$,
1845: since $C$ is one-dimensional. Thus
1846:       $$[E_i^+]=c_2({\cal J}^{g+i}_{p_1}({\cal L}_i))
1847:       -c_1({\cal J}^{g+i}_{p_1}({\cal L}_i))
1848:       c_1(p_1^*{\cal E}_i).$$
1849: Using the truncation sequence of the bundles of jets,
1850: we get
1851:       \begin{eqnarray*}
1852:       c_1({\cal J}^{g+i}_{p_1}({\cal L}_i))&=&
1853:       \sum_{\ell=1}^{g+i+1}(\ell K_2+(i+1)[\Delta]);\\
1854:       c_2({\cal J}^{g+i}_{p_1}({\cal L}_i))&=&
1855:       \sum_{m=2}^{g+i+1}\sum_{\ell=1}^{m-1}
1856:       (\ell K_2+(i+1)[\Delta])(m K_2+(i+1)[\Delta]).
1857:       \end{eqnarray*}
1858: Expanding, and using that $K_2^2=0$, we get
1859:       \begin{eqnarray*}
1860:       c_1({\cal J}^{g+i}_{p_1}({\cal L}_i))&=&
1861:       \frac{1}{2}(g+i+1)(g+i+2)K_2+(i+1)(g+i+1)[\Delta];\\
1862:       c_2({\cal J}^{g+i}_{p_1}({\cal L}_i))&=&
1863:       \frac{1}{2}(i+1)(g+i)(g+i+1)(g+i+2)K_2[\Delta]\\
1864:       &&+\frac{1}{2}(i+1)^2(g+i)(g+i+1)[\Delta]^2.
1865:       \end{eqnarray*}
1866: Finally, using Formula (\ref{eq:c1Ei}) for $j=i$, and
1867: Formulas (\ref{eq:kd}) and (\ref{eq:kk}), we get
1868:       $$\int_{C\times C}[E_i^+]=(i+1)^2g(g-1)(g+i+1)^2.$$
1869: 
1870: Now, it follows from Proposition \ref{lem} that
1871: $W_i$ meets $\Delta$
1872: transversally at $g^3-g$ points. Thus, using
1873: Proposition \ref{A2CC}, we get
1874:       \begin{eqnarray*}
1875:       \int_{C\times C}[E_i]&=&\int_{C\times C}[E_i^+]-
1876:       (g+1)(g^3-g)\\
1877:       &=&g(g-1)\Big((g+i+1)^2(i+1)^2-(g+1)^2\Big),
1878:       \end{eqnarray*}
1879: the stated formula for the degree of $[E_i]$.
1880: 
1881: Now, the expression for the degree of $[D_i]$ follows 
1882: now from the equality $[SW_i]=[D_i]+[E_i]$ proved in
1883: Proposition \ref{A2CC} and Formula (\ref{eq:me1})
1884: for the degree of $[SW_i]$ proved in Proposition
1885: \ref{prop07}.
1886: 
1887: Let us now show that $D_i$ and $E_i$ are reduced. 
1888: Let $(P,Q)\in SW_i$. Let $\wh{\cal O}$ be the completion 
1889: of the local ring of $C\times C$ at $(P,Q)$. 
1890: Let $t_1$ and $t_2$ be local equations in $\wh{\cal O}$ 
1891: for $\{P\}\times C$ and $C\times\{Q\}$, 
1892: respectively. Then $\wh{\cal O}=\CC[[t_1,t_2]]$. 
1893: Let $w\in{\cal O}$ be a local equation for 
1894: $W_i$. Since $(P,Q)\in SW_i$, and since $W_i$ is 
1895: nonsingular by Proposition \ref{lem}, we may assume that 
1896: $w=t_1+u$, where $u\in\CC[[t_2]]$. Now, 
1897: let $w'$ and $u'$ be the derivatives of $w$ and $u$ 
1898: with respect to $t_2$. Then the ideal defining 
1899: $SW_i$ at $(P,Q)$ is $(w,w')$, and the multiplicity of 
1900: the cycle $[SW_i]$ at $(P,Q)$ is 
1901: $\ell(\wh{\cal O}/(w,w'))$. Notice that $w'=u'$, and 
1902:        $$\frac{\wh{\cal O}}{(w,w')}\cong
1903:          \frac{\CC[[t_2]]}{(u')}=
1904:          \frac{\CC[[t_2]]}{(u,u')}\cong
1905: 	 \frac{\CC[[t_1,t_2]]}{(t_1,w,w')}.$$
1906: Thus the multiplicity of the cycle $[SW_i]$ at $(P,Q)$
1907: is the multiplicity $m$ of $SW_i\cap(\{P\}\times C)$ at 
1908: $(P,Q)$.
1909: 
1910: Since the formation of $SW_i$ commutes with base 
1911: change, this multiplicity $m$ satisfies
1912:        $$m=\wgt_V(Q)-1,$$
1913: where $V$ is the complete linear system of sections 
1914: of $\w_C((i+1)P)$. Now, by Propositions \ref{WiPgen} and 
1915: \ref{arbD}, the order sequence of $V$ at $Q$ satisfies
1916:        \begin{eqnarray*}
1917: 	 \epsilon_j(V,Q)&=&j\quad(j=0,1,\dots,g+i-3),\\
1918: 	 \epsilon_{g+i-2}(V,Q)&\leq&g+i-1,\\
1919: 	 \epsilon_{g+i-1}(V,Q)&\leq&g+i+1.
1920:        \end{eqnarray*}
1921: Thus $m\leq 2$, with equality if and only if 
1922:        $$h^0(\w_C((i+1)P-(g+i-1)Q))=2\quad{\rm and}
1923:          \quad h^0(\w_C((i+1)P-(g+i+1)Q))=1,$$
1924: that is, if and only if $(P,Q)\in D_i\cap E_i$. 
1925: Since $[SW_i]=[D_i]+[E_i]$ by Proposition~\ref{A2CC}, 
1926: it follows that $D_i$ and $E_i$ are reduced at 
1927: $(P,Q)$.\qed
1928: 
1929: \begin{thebibliography}{99}
1930: 
1931: \bibitem{A} E.~Arbarello, 
1932: {\it Weierstrass points and moduli of curves}. 
1933: Compo. Math. {\bf 29}, no. 3, (1974), 325--342.
1934: 
1935: \bibitem{acgh} E.~Arbarello, M. Cornalba, 
1936: P. A. Griffiths and J. Harris, 
1937: {\it Geometry of algebraic curves}.
1938: Grundlehren der mathematischen Wissenschaften, 
1939: vol.~267, Springer, New York, 1985.
1940: 
1941: \bibitem{cocoa} CoCoATeam, 
1942: {\it \cocoa: a system for doing Computations in 
1943: Commutative Algebra}.
1944: Available at http://cocoa.dima.unige.it.
1945: 
1946: \bibitem{Cukie} F.~Cukierman,
1947: {\it Families of Weierstrass points}.
1948: Duke Math. J. {\bf 58}, no. 2, (1989), 317--346.
1949: 
1950: \bibitem{CEG} C.~Cumino, E.~Esteves, and L.~Gatto,
1951: {\it Limits of special Weierstrass points}.
1952: To appear.
1953: 
1954: \bibitem{Diaz/Exc} S.~Diaz,
1955: {\it Exceptional Weierstrass points and the divisor
1956: on moduli space that they define}.
1957: Mem. Amer. Math. Soc., vol. 56, no. 327, Amer. Math. 
1958: Soc., Providence, 1985.
1959: 
1960: \bibitem{MathCont} E.~Esteves,
1961: {\it Linear systems and ramification points on
1962: reducible nodal curves}.
1963: In ``Algebra Meeting''  (A. Garcia, E. Esteves and
1964: A. Pacheco, Eds.), 21--35, Mat. Contemp., vol. 14,
1965: Soc. Bras. Mat., Rio de Janeiro, 1998.
1966: 
1967: \bibitem{Fu} W.~Fulton,
1968: {\it Intersection theory}.
1969: Grundlehren der mathematischen Wissenschaften, vol. 2,
1970: Springer, Berlin Heidelberg, 1984.
1971: 
1972: \bibitem{GP} L.~Gatto and F.~Ponza,
1973: {\it Derivatives of wronskians with applications to
1974: families of Weierstrass points}.
1975: Trans. Amer. Math. Soc. {\bf 351}, no. 6, (1999), 2233--2255.
1976: 
1977: \bibitem{HaMu} J.~Harris and D.~Mumford, 
1978: {\it On the Kodaira dimension of the moduli space of 
1979: curves}.
1980: Invent. Math. {\bf 67} (1982), 23--86.
1981: 
1982: \bibitem{Ke} S.~Keel,
1983: {\it Intersection theory of moduli space of stable 
1984: $n$-pointed curves of genus zero}. 
1985: Trans. Amer. Math. Soc. {\bf 330} (1992), 545--574.
1986: 
1987: \bibitem{Kn} F.~Knudsen,
1988: {\it The projectivity of the moduli space of 
1989: stable curves, II}. Math. Scand. {\bf 52} (1983), 
1990: 161--199.
1991: 
1992: \end{thebibliography}
1993: 
1994: \end{document}
1995: