1: % Izzet:
2: %
3: % definability and degree
4: % tau-invertible sheaves and degrees. how many up to isom?
5: %
6: %% EDIT; FIX
7: %% Fix \PO
8: %% After ++++++ is scraps.
9: %% See %% ??
10:
11: %\documentclass[11pt]{article}
12: \documentclass{amsart}
13:
14: \raggedbottom
15:
16: \usepackage{fullpage,eucal,amsmath,amsthm,amsfonts,verbatim,amsopn,amssymb}
17:
18: \usepackage[all]{xy}
19:
20: \newcommand\dmo{\DeclareMathOperator}
21:
22: \renewcommand\sec{\section}
23: \newcommand\subs{\subsection*}
24: %\newcommand\ssubs{\subsubsection*}
25: \newcommand\para{\paragraph*}
26: \newcommand\bea{\begin{array}{r@{,\ \ \ \ \ \ \ }l}}
27: \newcommand\ena{\end{array}}
28: \newcommand\barr{\begin{array}}
29: \newcommand\earr{\end{array}}
30: \newcommand\ben{\begin{enumerate}}
31: \newcommand\een{\end{enumerate}}
32: \newcommand\beq{\begin{equation}}
33: \newcommand\eeq{\end{equation}}
34: \newcommand\bqa{\begin{eqnarray*}}
35: \newcommand\eqa{\end{eqnarray*}}
36: \newcommand\bqan{\begin{eqnarray}}
37: \newcommand\eqan{\end{eqnarray}}
38: \newcommand\bit{\begin{itemize}}
39: \newcommand\eit{\end{itemize}}
40: \newcommand\bfi{\begin{figure}[htbp]}
41: \newcommand\efi{\end{figure}}
42: \newcommand\bce{\begin{center}}
43: \newcommand\ece{\end{center}}
44: \newcommand\note{\marginpar}
45: \newcommand\bpr{\begin{proof}}
46: \newcommand\epr{\end{proof}}
47:
48: \newcommand{\ignore}[1]{}
49:
50: %% theorems
51:
52: \newtheorem{theorem}{Theorem}[section]
53: \newtheorem{conjecture}[theorem]{Conjecture}
54: \newtheorem{lemma}[theorem]{Lemma}
55: \newtheorem{corollary}[theorem]{Corollary}
56: \newtheorem{proposition}[theorem]{Proposition}
57:
58: \theoremstyle{definition}
59:
60: \newtheorem{definition}[theorem]{Definition}
61: \newtheorem{example}[theorem]{Example}
62: \newtheorem{remark}[theorem]{Remark}
63: \newtheorem{remarks}[theorem]{Remarks}
64: \newtheorem{examples}[theorem]{Examples}
65: \newtheorem{question}[theorem]{Question}
66:
67: \newcommand\bthm{\begin{theorem}}
68: \newcommand\ethm{\end{theorem}}
69: \newcommand\bcn{\begin{conjecture}}
70: \newcommand\ecn{\end{conjecture}}
71: \newcommand\bla{\begin{lemma}}
72: \newcommand\ela{\end{lemma}}
73: \newcommand\bco{\begin{corollary}}
74: \newcommand\eco{\end{corollary}}
75: \newcommand\bpro{\begin{proposition}}
76: \newcommand\epro{\end{proposition}}
77: \newcommand\bdf{\begin{definition}}
78: \newcommand\edf{\end{definition}}
79: \newcommand\bex{\begin{example}}
80: \newcommand\eex{\end{example}}
81: \newcommand\brm{\begin{remark}}
82: \newcommand\erm{\end{remark}}
83: \newcommand\brms{\begin{remarks}}
84: \newcommand\erms{\end{remarks}}
85: \newcommand\bqu{\begin{question}}
86: \newcommand\equ{\end{question}}
87:
88: %% symbol abbrevatiations
89:
90: \newcommand\ind{\indent}
91: \newcommand\no{\noindent}
92: \newcommand\bsn{\bigskip \noindent}
93: \newcommand\bsk{\bigskip}
94: \newcommand\msn{\medskip \noindent}
95: \newcommand\msk{\medskip}
96: \newcommand\ssn{\smallskip \noindent}
97: \newcommand\ssk{\smallskip}
98:
99: \newcommand\tu{\textup}
100:
101: %% math and logical symbols
102:
103: \newcommand\Prf{\textsc{Proof}}
104: \newcommand\prf{\textsc{Proof. \quad}}
105: \newcommand\prfprop{\textsc{Proof of Proposition. \quad}}
106: \newcommand\sprf{\textsc{Sketch of Proof. \quad}}
107:
108: \newcommand\sr{\stackrel}
109: \newcommand\<{\langle}
110: \renewcommand\>{\rangle}
111: \newcommand\ot{\otimes}
112: \newcommand\op{\oplus}
113:
114: \newcommand\sub{\subseteq}
115: \renewcommand\sup{\supseteq}
116: \newcommand\subn{\subset}
117: \newcommand\supn{\supset}
118: \newcommand\all{\forall}
119: \newcommand\ex{\exists}
120: \newcommand\surj{\twoheadrightarrow}
121: \newcommand\inj{\hookrightarrow}
122: \newcommand\im{\Rightarrow}
123: \newcommand\rim{\Leftarrow}
124: \newcommand\ra{\rightarrow}
125: \newcommand\la{\leftarrow}
126: \newcommand\lra{\longrightarrow}
127: \newcommand\lla{\longleftarrow}
128: \newcommand\hra{\hookrightarrow}
129: \newcommand\map{\mapsto}
130: \newcommand\lmap{\longmapsto}
131: \newcommand\seq{\sim}
132: \newcommand\teq{\approx}
133: \newcommand\teqq{\parallel}
134: \newcommand\ch{\choose}
135: \newcommand\eq{\equiv}
136: \newcommand\dv{\uplus}
137: \newcommand\rst{|}
138:
139: %% letters and fonts
140:
141: \newcommand\al{{\alpha}}
142: \newcommand\be{{\beta}}
143: \newcommand\ga{{\gamma}}
144: \newcommand\dl{\delta}
145: \newcommand\dltl{\tilde{\delta}}
146: \newcommand\lm{\lambda}
147: \newcommand\Lm{\Lambda}
148: \newcommand\cm{\otimes}
149: \newcommand\om{\omega}
150: \newcommand\omt{\omega^\tau}
151: \newcommand\Om{\Omega}
152: \newcommand\Omt{\Omega^\tau}
153: \newcommand\Omtp{\Omega^{\tau,p}}
154: \newcommand\Omti{\Omega^{\tau,i}}
155: \newcommand\Omtcx{\Omega^{\tau\bullet}}
156:
157: \newcommand\pd{\partial}
158: \newcommand\xtl{\tilde{X}}
159:
160: \newcommand\xbar{\overline{x}}
161: \newcommand\ybar{\overline{y}}
162: \newcommand\zbar{\overline{z}}
163: \newcommand\abar{\overline{a}}
164: \newcommand\bbar{\overline{b}}
165: \newcommand\cbar{\overline{c}}
166: \newcommand\albar{{\overline{\alpha}}}
167: \newcommand\bebar{{\overline{\beta}}}
168: \newcommand\gabar{{\overline{\gamma}}}
169:
170: \newcommand\A{{\mathbb{A}}}
171: \newcommand\M{{\mathbb{M}}}
172: \newcommand\N{{\mathbb{N}}}
173: \newcommand\Z{{\mathbb{Z}}}
174: \newcommand\K{{\mathbb{K}}}
175: \renewcommand\L{{\mathbb{L}}}
176: \newcommand\F{{\mathbb{F}}}
177: \newcommand\Q{{\mathbb{Q}}}
178: \newcommand\R{{\mathbb{R}}}
179: \newcommand\C{{\mathbb{C}}}
180: \renewcommand\H{{\mathbb{H}}}
181: \renewcommand\P{{\mathbb{P}}}
182: \renewcommand\S{{\mathbb{S}}}
183: \newcommand\T{{\mathbb{T}}}
184:
185: \renewcommand\a{{\mathfrak{a}}}
186: \newcommand\g{{\mathfrak{g}}}
187: \newcommand\h{{\mathfrak{h}}}
188: \renewcommand\i{{\mathfrak{i}}}
189: \renewcommand\j{{\mathfrak{j}}}
190: \renewcommand\k{{\mathfrak{k}}}
191: \renewcommand\l{{\mathfrak{l}}}
192: \newcommand\m{{\mathfrak{m}}}
193: \newcommand\n{{\mathfrak{n}}}
194: \newcommand\p{{\mathfrak{p}}}
195: \newcommand\s{{\mathfrak{s}}}
196: \renewcommand\u{{\mathfrak{u}}}
197: \renewcommand\v{{\mathfrak{v}}}
198: \newcommand\x{{\mathfrak{x}}}
199: \newcommand\z{{\mathfrak{z}}}
200: \newcommand\gl{{\mathfrak{gl}}}
201: \newcommand\so{{\mathfrak{so}}}
202: \newcommand\U{{\mathfrak{U}}}
203: \newcommand\V{{\mathfrak{V}}}
204: \newcommand\X{{\mathfrak{X}}}
205: \newcommand\Y{{\mathfrak{Y}}}
206:
207: \newcommand\DD{{\mathfrak{D}}}
208: \newcommand\calp{{\mathfrak{p}}}
209: \newcommand\calO{{\mathcal{O}}}
210: \renewcommand\O{{\mathcal{O}}}
211:
212: \newcommand\sym{{\mathbb{S}}}
213:
214: \newcommand\Dun{{\underline{D}}}
215: \newcommand\dun{{\underline{d}}}
216:
217: %% specific abbreviations
218:
219: \newcommand\aff{\textup{Aff}}
220: \newcommand\Hom{\textup{Hom}}
221: \newcommand\Alg{\textup{Alg}}
222: \newcommand\Ext{\textup{Ext}}
223: \renewcommand\dim{\textup{dim}}
224: \newcommand\HH{\textup{H}}
225: \newcommand\hh{h}
226: \newcommand\Spec{\textup{Spec}}
227: \newcommand\BSpec{\textup{\textbf{Spec}}}
228: \newcommand\DSpec{\textup{DSpec}}
229: \newcommand\Der{\textup{Der}}
230: \newcommand\Ker{\textup{Ker}}
231: \newcommand\Mor{\textup{Mor}}
232: \newcommand\HS{\textup{HS}}
233: \newcommand\acl{\textup{acl}}
234: \newcommand\cl{\textup{cl}}
235: \dmo{\spec}{Spec}
236:
237:
238: \newcommand\sm{strongly minimal\,}
239: \newcommand\dring{$\DD$-ring}
240: \newcommand\drings{$\DD$-rings}
241: \newcommand\dfield{$\DD$-field}
242: \newcommand\dfields{$\DD$-fields}
243: \newcommand\dsch{$\DD$-\textup{scheme}}
244: \newcommand\prl{^{(1)}}
245:
246: \newcommand{\GL}{\textup{GL}}
247: \newcommand{\Diff}{\textup{Diff}}
248: \newcommand{\Aff}{\textup{Aff}}
249: \newcommand{\Tr}{\textup{Tr}}
250: \newcommand\MA{{\mathfrak{A}}}
251: \newcommand\MO{{\mathcal{O}}}
252: \newcommand\ME{{\mathcal{E}}}
253: \newcommand\MF{{\mathcal{F}}}
254: \newcommand\MG{{\mathcal{G}}}
255: \newcommand\MH{{\mathcal{H}}}
256: \newcommand\MFT{{\mathcal{F}^\tau}}
257: \newcommand\MGT{{\mathcal{G}^\tau}}
258: \newcommand\tis{$\tau$-invertible sheaf}
259: \newcommand\tiss{$\tau$-invertible sheaves}
260: \newcommand\tauf{$\tau$-form}
261: \newcommand\taufs{$\tau$-forms}
262: \newcommand\PO{PO}
263:
264: %% end abbreviations
265:
266: \begin{document}
267:
268: \title{Order 1 strongly minimal sets in differentially closed fields}
269: \author{Eric Rosen}
270: \address{Department of Mathematics\\
271: Massachusetts Institute of Technology\\
272: Cambridge, MA 02139}
273: \email{rosen@math.mit.edu}
274: \urladdr{http://math.mit.edu/~rosen/}
275: \thanks{This paper constitutes a revised version of part of my dissertation, at the
276: University of Illinois at Chicago (2005). I am grateful to Lawrence Ein,
277: Henri Gillet, David Marker, and Thomas Scanlon for helpful discussions on this material.}
278:
279: \date{\today}
280:
281: \begin{abstract}
282: We give a classification of non-orthogonality classes of trivial order
283: 1 strongly minimal sets in differentially closed fields. A central idea
284: is the introduction of $\tau$-forms, functions on the prolongation of
285: a variety which are analogous to 1-forms. Order 1 strongly minimal sets then
286: correspond to smooth projective curves with $\tau$-forms. We also
287: formulate our results scheme-theoretically, in terms of $\tau$-differentials
288: and $\tau$-invertible sheaves on curves, thereby obtaining additional
289: information about the strongly minimal sets.
290: This work partially generalizes and extends results of Hrushovski and Itai.
291: \end{abstract}
292:
293: \maketitle
294:
295: \sec*{Introduction}
296:
297: This paper addresses the problem of classifying the strongly minimal sets
298: definable in differentially closed fields. A nice description of
299: the background to the question and known results can be found in the
300: introduction of \cite{HI},
301: so we only give a brief introduction here. For more on the model theory of
302: differential fields, see also \cite{MMP}, \cite{Pil02},
303: \cite{HS}, and \cite{HICM}.
304: The precise problem is to classify strongly minimal\ sets up to non-orthogonality,
305: a natural logical equivalence relation. Associated to each strongly minimal\ set
306: is a combinatorial geometry, which is either non-locally modular, locally
307: modular non-trivial, or trivial, thus dividing strongly minimal\ sets into three
308: classes. Hrushovski and Sokolovi\'c \cite{HS} classified
309: both the non-locally modular and the locally modular non-trivial strongly minimal\ sets.
310: The former are non-orthogonal to the field of constants, which is strongly minimal,
311: and the latter correspond to (isogeny classes of) simple abelian varieties
312: that do not descend to the constants. What remains is to classify the
313: trivial ones.
314:
315: Let $K$ be a universal domain for DCF, the theory of differentially closed
316: fields, and let $k$ denote the field of constants.
317: A strongly minimal\ set, or formula, determines a unique strongly minimal\ type, and sets/formulas
318: determining the same type are non-orthogonal, so one can equally well consider
319: strongly minimal\ types. Recall that if $q$ is a strongly minimal\ type over a differential field $L$,
320: then the {\em order} of $q$ is $\textup{td}((\abar, \abar', \ldots)/L)$,
321: for $\abar$ realizing $q$, which is always finite.
322: It seems to be unknown whether there are (trivial)
323: strongly minimal\ types of arbitrarily large order, though there are examples of order 2.
324: One consequence of \cite{HI} is that there are many types of order 1.
325:
326: Hrushovski \cite{HJ} proved that every order 1 strongly minimal\ type that is orthogonal to $k$
327: is $\om$-{\em categorical} and thus trivial, since the classification of
328: the non-trivial locally modular strongly minimal\ types by Manin kernels implies that
329: they are never $\om$-categorical.
330: (Alternatively, one can derive the triviality of
331: order 1 locally modular strongly minimal sets
332: as a consequence of the fact that any non-trivial locally modular
333: strongly minimal set is non-orthogonal to the Manin kernel of a simple abelian
334: variety~\cite{HS} and Buium's proof that any such Manin kernel has transcendence
335: degree at least 2). With Itai, Hrushovski then gives a quite precise
336: description of the trivial strongly minimal\ types over $k$, which we summarize below.
337: In this paper, we consider all order 1 strongly minimal\ types,
338: partially generalizing and extending some results of \cite{HI}.
339:
340: \subs{Background}
341: It is well-known, and easy to see, that (complete, stationary) $n$-types in
342: pure algebraically closed fields correspond to irreducible Zariski closed
343: subsets of $\A^n$. Call such a set an {\em embedded affine variety},
344: as it is an affine variety together with a closed embedding into affine space.
345: We recall a similar geometric description of complete strongly minimal\ types in DCF,
346: which also holds more generally for types of finite rank.
347: (See \cite{HI} or \cite{Pil02} for more details.)
348:
349: Say that two types $p$ and $q$ are {\em interdefinable}
350: if, perhaps passing to non-forking extensions over a common base field,
351: there is a definable bijection between their sets of realizations. Then
352: any strongly minimal\ type $p$ in DCF is interdefinable with a type $q$ such that for
353: any realization $\abar$ of $q$, $\dl(\abar) = s(\abar)$, where $s(\xbar)$ is a
354: tuple of polynomials. Since interdefinable types are non-orthogonal, we
355: may assume that all (strongly minimal) types are of this form.
356:
357: Given an affine variety $V \sub \A^n$, the map $\dl$ above, acting
358: componentwise, is a section of the first prolongation $\tau V \sub
359: \A^{2n}$, a (possibly reducible) affine variety defined below,
360: also called the {\em shifted tangent bundle}.
361: By the above reduction, a strongly minimal\ type $p$ corresponds to a pair
362: $(V,s)$, where $V \sub \A^n$ is an embedded affine variety,
363: and $s: V \ra \tau V\sub \A^{2n}$
364: is a section of the {\em first prolongation}. In this case,
365: one says that $p$ {\em lives on} the variety $V$. It is easy to check
366: that the order of $p$ equals the dimension of $V$, so
367: the order 1 types are exactly those that live on curves.
368:
369: In general, a pair $(V,s)$ as above determines a Kolchin closed set
370: of finite Morley rank, $\Xi(V,s):= \{v \in V | \dl(v)=s(v) \}$. When $V$
371: is a curve, this set is necessarily strongly minimal, so there's actually a bijection
372: between (Kolchin closed) order 1 strongly minimal sets and pairs $(C,s)$, $C$ an embedded
373: affine curve and $s$ a section of the prolongation, $s:C\ra \tau C$.
374: (Moreover, this bijection preserves fields of definition.)
375: This gives a precise, purely geometric characterization of order 1
376: strongly minimal sets.
377: What remains, though, is to determine whether or not such a
378: strongly minimal\ set $\Xi(C,s)$ is trivial, and to understand non-orthogonality.
379:
380: If $X$ is a strongly minimal\ set living on a curve $C$, and $C$ is birational to
381: some $C'$, then $X$ is interdefinable with some $X'$ living on $C'$.
382: Since every curve is birational to a unique smooth projective curve,
383: it suffices to consider such sets living on (embedded) smooth projective
384: curves. Strictly speaking, then, we want to classify strongly minimal
385: sets of the form $\Xi(C,s)$, $C$ an embedded smooth projective
386: curve, $s: C \ra \tau C$. But we will consider $C$ as an
387: abstract curve, since the particular embedding in $\P^n$ is unimportant,
388: as different embeddings yield interdefinable strongly minimal sets.
389:
390: Given a curve $C$ (or any variety) defined over $k$, the prolongation
391: $\tau C$ equals the tangent variety $TC$. In this case, a section
392: of the prolongation $s: C \ra \tau C$ is just a vector field. For curves,
393: there is also a natural bijection between vector fields and 1-forms.
394: Given a curve $C$ and a vector field $s$, let $\om$ be the 1-form such
395: that $\om (s) = 1$ almost everywhere. Thus, order 1 strongly minimal\ sets defined over
396: the constants are also represented by pairs $(C,\om)$, defined over
397: the constants.
398:
399: Using this representation, Hrushovski and Itai obtain a rather complete
400: description of non-orthogonality classes of trivial order 1 strongly minimal\ sets
401: defined over the constants. To any such class they associate a unique
402: (smooth projective) curve, the main idea being to pick a canonical
403: {\em strictly minimal} set in any such class. Second, they prove that for
404: any curve of genus $\geq 2$, there are many classes associated to that curve,
405: thereby proving that there are indeed `very many' trivial order 1 strongly minimal\ sets
406: defined over the constants.
407:
408: \subs{Results}
409:
410: The original motivation for this work was to generalize results from
411: \cite{HI} to all order 1 strongly minimal\ sets. One of the main new ideas is the
412: introduction of $\tau$-forms, which are functions on the prolongation of
413: a curve, analogous to 1-forms, which are functions on the tangent variety.
414: This enables one to develop a `geometric approach' to order one strongly minimal\ sets
415: as in \cite{HI}, with a pair $(C,\omt)$, $C$ a curve, $\omt$ a $\tau$-form,
416: representing a strongly minimal\ set. We can then characterize non-orthogonality
417: classes of trivial strongly minimal\ sets in terms of `essential $\tau$-forms',
418: appropriately defined, generalizing a central result of \cite{HI}.
419:
420: In a separate paper~\cite{Ros}, we introduce and develop the theory of $\tau$-differentials,
421: the algebraic analog of $\tau$-forms. We use this material here to give an alternative
422: characterization of order one strongly minimal sets in terms of curves with
423: $\tau$-invertible sheaves.
424: This makes it possible to determine the dimension of the space of global
425: $\tau$-forms on a curve, in terms of its Kodaira-Spencer class.
426: Further, we show that the set of global $\tau$-forms on a curve is (uniformly) definable,
427: as is the dimension of this set.
428: It also leads to the introduction of the prolongation cone of a curve, a vector bundle
429: into which both the tangent variety and prolongation naturally embed.
430:
431: \subs{Open questions}
432:
433: The main question left open by this work is whether for every curve $C$
434: of genus $>1$ there are $\sim$-essential \taufs. For curves defined over
435: $k$, a positive answer follows from \cite{HI}. A positive
436: answer for all curves would show that there are order 1 trivial strongly minimal\ sets
437: that are orthogonal to any such set defined over $k$.
438: %It also remains to
439: %look more carefully at strongly minimal\ sets on curves of genus 0 or 1. When everything
440: %is defined over $k$, then the situation is well-understood, but the general
441: %case may be more complex, especially for elliptic curves.
442:
443: Two main problems about trivial strongly minimal\ sets in DCFs, asked by
444: Hrushovski, remain completely open. Are they all $\om$-categorical?
445: And can they be classified? Extending the classification beyond order 1
446: certainly seems difficult. For example, presumably one would need to
447: be able to determine whether a finite rank type given as a pair $(V,s)$
448: is strongly minimal. But perhaps there are particular cases of order 2 types,
449: thus living on surfaces, that are more accessible.
450: There are also related questions, which might be approached
451: by existing methods. Hrushovski suggests that one could try to show
452: that the solutions to a generic order 2 equation are strongly minimal,
453: to answer a question of Poizat.
454: Another well-known question is whether every U-Rank 1 type is strongly minimal.
455: This is related to Hrushovski and Scanlon's result~\cite{HrSc}
456: that U-Rank and Morley rank differ in DCF.
457:
458: \sec{Prolongations}
459: \label{prolong}
460: We recall the construction of the prolongation of a variety over a differentially
461: closed field. Buium's original definition uses the language of schemes,
462: though we give a description in local coordinates used more often by
463: model-theorists. For a variety $V$ defined over the constants, its prolongation
464: is just the tangent variety $TV$. In general, though, it is a $TV$-{\em torsor}.
465: Thus it is also called the {\em shifted tangent bundle}.
466: In this case, the fibers of the canonical projection to $V$ are no longer
467: vector spaces, but affine spaces, defined below.
468:
469: \para{Affine Spaces}
470: An affine space is essentially a vector space without a
471: distinguished point as origin. Affine spaces arise in model theory as combinatorial
472: geometries that are locally modular but not modular. Just as a locally
473: modular geometry becomes modular when one fixes any point, choosing a point in
474: an affine space naturally gives a vector space. The presentation here
475: differs from than that in, for example, Hodges' textbook~\cite{Hod},
476: but is basically
477: equivalent. (The word ``affine'' will also be used in a
478: completely different sense, in connection with affine varieties.)
479:
480: \bdf
481: \label{aff}
482: Let $K$ be a field. A {\em K-affine space} is a triple $(A, V, \alpha)$,
483: where $A$ is a set, $V$ is a $K$-vector space, and $\alpha$ is a regular
484: action of $V$ on $A$. The {\em dimension} of $A$, $\dim(A)$, is just $\dim(V)$.
485: \edf
486:
487: We will write $\alpha(v,a)$ as $v \cdot a$, and often omit the action $\alpha$
488: when it is understood.
489:
490: \brm
491: \label{affbasics}
492: \ben
493: \item For any $K$-vector space $V$, $(V, V)$ is an affine space in a natural way.
494: \item Given $(A, V)$ and $a \in A$, there is a natural bijection $i_a:V\lra A$
495: given by $i_a(v) = v\cdot a$.
496: \een
497: \erm
498:
499: \bdf
500: An {\em affine map} between $K$-affine spaces $(A, V, \alpha)$ and $(B, W, \beta)$
501: is a map $f:A \lra B$ such that there is a linear map $\lambda f: V \lra W$ so that
502: for all $a\in A, v \in V$, $f(v\cdot a) = (\lambda f (v))\cdot f(a)$.
503: Let $\Aff(A,B)$ denote the set of all affine maps from $A$ to $B$.
504: (By putting coordinates on $A$ and $B$, this set can be endowed with a vector
505: space structure.)
506:
507: \edf
508:
509: \brm
510: \label{affrems}
511: \ben
512: \item Given an affine map $f$, the linear map $\lambda f$ is uniquely determined.
513: In fact, $\lambda$ is a functor from the category of $K$-affine spaces to the
514: category of $K$-vector spaces.
515: \item Given affine maps $f, g$, from $(A, V)$ to $(B, W)$, $\lambda f = \lambda g$
516: if and only if there is a $w \in W$ such that for all $a \in A$, $f(a) = w \cdot g(a)$.
517: \item There is a short exact sequence
518: %\note{[?!]}
519: $$
520: 0 \lra B \stackrel{\mu}{\lra}\aff(A, B)\stackrel{\lambda}{\lra}\Hom(V, W) \lra 0
521: $$
522: where for all $b_0 \in B, \mu(b_0)$ is the constant affine map
523: $\mu(b_0): A \lra B$ mapping each $a \in A$ to $b_0$. Thus $\dim(\aff(A, B)) =
524: \dim(B) + \dim(A)\dim(B)$.
525: \een
526: \erm
527:
528: We are particularly interested in the case, $W = K$, i.e., when $B$ is
529: 1-dimensional.
530: Then $\aff(A, B)$ is something like a dual space to $(A, V)$, though
531: $\dim(\aff(A,B))=\dim(A) + 1$.
532:
533: \brm
534: Let $(A,V)$ be an affine space. Recall that the {\em affine group} of $V$,
535: denoted $\Aff(V)$, is the group generated by the translation group $\Tr(V)$
536: and $\GL(V)$. In fact, $\Aff(V)=\Tr(V)\rtimes \GL(V)$.
537: %%change \bowtie to \rtimes
538: The automorphism group of $(A,V)$ is naturally isomorphic to $\Aff(V)$.
539: \erm
540:
541: \para{Varieties, tangent spaces, and prolongations}
542: A prevariety is a topological space with an open cover,
543: $W = W_1 \cup \ldots \cup W_m$ and compatible coordinate charts, $f_i:W_i \lra V_i$,
544: $V_i$ an affine variety. For all $i, j \leq m$, let $U_{i, j} = f_i(W_i \cap
545: W_j) \sub V_i$ and $f_{i, j}: U_{i, j} \lra U_{j, i}$ be $f_j\circ f_i^{-1}$.
546: For our purposes, a variety will be an irreducible, smooth separated prevariety.
547:
548: Recall that, given a polynomial $p(X)$,
549: $p^\delta(X)$ denotes the polynomial obtained by applying $\delta$ to
550: each of the coefficients. For more information, see
551: Marker~\cite{Mar03} or Pillay~\cite{Pil02}. The following easy observation
552: will be useful.
553:
554: \bla
555: For any $n$, let $\epsilon: K[x_1, \ldots , x_n] \lra K[x_1, \ldots , x_n]$
556: be the map that takes any polynomial $f$ to $f^\dl$, which is obtained
557: from $f$ by taking the derivative of each coefficient. Then $\epsilon$
558: is a derivation on $K[x_1, \ldots , x_n]$. Further, $\epsilon$ commutes
559: with each derivation $\frac{d}{dx_i}$.
560: \ela
561:
562: \bpr
563: Straightforward.
564: \epr
565:
566: \bdf
567: Let $V \subseteq K^n$ be an irreducible smooth affine variety,
568: $I(V) = \<p_1,\ldots , p_m\>$. The tangent space of $V$ is
569: $$
570: TV = \{(a, u) \in K^{2n} : \sum_{i=1}^n\frac{\partial p_j}{\partial x_i}(a)
571: \cdot
572: u_i = 0, j = 1, \ldots , m\}
573: $$
574: The first prolongation of $V$ is
575: $$
576: \tau V = \{(a, u) \in K^{2n} : \sum_{i=1}^n\frac{\partial p_j}{\partial x_i}(a)
577: \cdot u_i + p_j^\delta(a)= 0, j = 1, \ldots , m\}
578: $$
579:
580: Since everything is functorial, one can also define $TV$ and $\tau V$
581: for general varieties, in a coordinate free manner.
582: \edf
583:
584: \brm
585: There are natural projection maps $\pi_T: TV \lra V$ and $\pi_\tau: \tau V \lra
586: V$. For smooth $V$, and $v \in V$, $\pi_T^{-1}(v)=TV_a$ is a $\dim(V)$-dimensional
587: vector space, and $\pi_\tau^{-1}(v)=\tau (V)_a$ is a $\dim(V)$-dimensional
588: affine space naturally acted on by $\pi_T^{-1}(v)$.
589:
590: Further, let $TV \times_V\tau V = \{(a, u, w) \in K^{3n} : (a, u) \in TV, (a, w) \in
591: \tau V\}$. The map $p: TV \times_V\tau V \lra \tau V$ given by
592: $p(a, u, w) = p(a, u+w)$ makes $\tau V$ a {\em torsor} under $TV$.
593: \erm
594:
595: We now introduce a new construction, the prolongation cone of a variety,
596: into which both the tangent variety and prolongation naturally embed. For more
597: details, see~\cite{Ros}.
598:
599: \bdf
600: Let $V \subseteq K^n$ be an irreducible smooth affine variety,
601: $I(V) = \<p_1,\ldots , p_m\>$. The {\em prolongation cone} of $V$ is
602: $$
603: \tilde{V} = \{(a, u, u') \in K^{2n+1} : \sum_{i=1}^n\frac{\partial p_j}{\partial x_i}(a)
604: \cdot
605: u_i + p_j^\delta(a)\cdot u'= 0, j = 1, \ldots , m\}
606: $$
607:
608: As in the case of the tangent variety, there is a natural projection map
609: $\tilde{\pi}:\tilde{V} \lra V$, whose fibres are vector spaces, now of dimension
610: $\dim (V) + 1$. More generally, the prolongation cone is a $\dim(V) + 1$ vector
611: bundle over $V$. Again, the construction also globalizes to general varieties.
612:
613: Observe that the intersection of the prolongation cone with the hyperplane $u' = 0$
614: is the tangent variety. Likewise, the intersection with the hyperplane $u' = 1$
615: is the prolongation.
616: \edf
617:
618: \brm
619: The tangent space $TV$ is a variety with `additional linear structure' on the fibers
620: $TV_v$. We recall
621: how to make this notion precise and indicate how to make rigorous the
622: notion that the fibers $\tau V_v$ of the prolongation are affine spaces.
623:
624: One can define a vector bundle as follows (see \cite{GHL}, p.\ 15).
625: Let $E$ and $V$ be smooth varieties over a field $K$, $\pi:E\lra V$ a regular map.
626: We say that $(\pi , E, V)$ is a {\em vector bundle of rank n} if the following
627: conditions hold.
628: \ben
629: \item $\pi$ is surjective.
630: \item There exists a finite open cover $(U_i)_{i\in I}$ of $V$, and
631: isomorphisms $h_i:\pi^{-1}(U_i)\lra U_i\times K^n$ such that for any
632: $x\in U_i$, $h_i(\pi^{-1}(x))=\{x\}\times K^n$.
633: \item For any $i,j \in I$, there is a regular map $g_{ij}:U_i\cap U_j \lra
634: \GL_n(K)$ such that the map
635: $$
636: h_i \circ h_j^{-1}:(U_i\cap U_j) \times K^n \lra (U_i\cap U_j) \times K^n
637: $$
638: is of the form $h_i\circ h_j^{-1}(x,v)=(x,g_{ij}(x)\cdot v)$.
639: \een
640:
641: $V\times K^n$ is the {\em trivial vector bundle} over $V$, and the definition
642: of a vector bundle says that it is {\em locally trivial}. Likewise, the
643: prolongation $\tau V$ is a locally trivial affine bundle, in a similar sense.
644: That is, let $V$
645: be a variety and $(K^n,K^n)$ $n$-dimensional affine space. Then
646: $(\pi , E, V)$ as above is a {\em trivial affine bundle of rank n} if it is
647: isomorphic to $V\times K^n$, where we consider $K^n$ as an affine, rather than
648: vector, space. More generally, $(\pi , E, V)$ as above is
649: an {\em affine bundle of rank n} if we have the following.
650: \ben
651: \item $\pi$ is surjective.
652: \item There exists a finite open cover $(U_i)_{i\in I}$ of $V$, and
653: isomorphisms $h_i:\pi^{-1}(U_i)\lra U_i\times K^n$ such that for any
654: $x\in U_i$, $h_i(\pi^{-1}(x))=\{x\}\times K^n$.
655: \item For any $i,j \in I$, there is a map $g_{ij}:U_i\cap U_j \lra
656: \Aff(K^n)$ such that the map
657: $$
658: h_i \circ h_j^{-1}:(U_i\cap U_j) \times K^n \lra (U_i\cap U_j) \times K^n
659: $$
660: is of the form $h_i\circ h_j^{-1}(x,v)=(x,g_{ij}(x)\cdot v)$.
661: \een
662:
663: \erm
664:
665: \brm
666: Another way to formalize the notion of an affine bundle is suggested by
667: the notion of a fiber bundle associated to a principal bundle
668: (See \cite{KN96}, p.\ 50--55. Compare also section III.3 on
669: Affine Connections, p.\ 125.)
670: Roughly, given a manifold $M$ and a Lie group $G$,
671: a principal $G$-bundle $P(G,M)$ is a fiber bundle
672: $\pi:E\lra M$, with a free $G$-action
673: on each fiber $\pi^{-1}(x), x\in M$. If $F$ is some other manifold, with
674: a $G$-action, then one can construct a fiber bundle $E(G,M,F,P)$ which is
675: the fiber bundle over $M$, with standard fiber $F$, and structure group
676: $G$, associated to the principal fiber bundle $P$. In our setting, $M$ is our
677: variety $V$, $G$ is $\Aff(K^n)$, and $F$ is our affine space.
678: \erm
679:
680: The following lemma is due to Buium \cite{Bui93}.
681: \bla
682: Let $V$ be a smooth variety of dimension $n$. Then $\tau V$ is an affine bundle
683: of rank $n$.
684: \ela
685:
686: \para{Tangent and lifting maps}
687: We now consider maps between varieties. Let $V,W$ be varieties,
688: and $\phi:V \ra W$ a regular map. The map $\phi$ determines a map
689: from $TV$ to $TW$, written $T\phi$, the {\em tangent map} of $\phi$,
690: which restricts, for each $a \in V$,
691: to a linear map on fibers, $T\phi_a:TV_a \ra TW_{f(a)}$.
692:
693: An important special case occurs when $f$ is a regular function on $V$,
694: viewed as a map from $V$ to $\A^1$. We consider the {\em differential},
695: $df$, which is the composition $df = \pi\circ Tf$, where
696: $\pi$ is the projection from $T\A^1 \ra \A^1$ onto the {\em tangent vector
697: component}. Thus, $df$ is a regular function on $TV$, and we have a
698: $K$-linear derivation $d:K[V]\ra K[TV]$. Alternatively, $d$ is a
699: derivation from $K[V]$ to $\Om[V]$, the regular differential forms
700: on $V$, which is a $K[V]$-module with a natural embedding in $K[TV]$.
701:
702:
703: For prolongations, there is also a
704: {\em lifting map} from $\tau V$ to $\tau W$,
705: which we write $\phi^{(1)}$, which restricts, for each $a\in V$, to an
706: affine map on fibers, $\phi^{(1)}_a:\tau V_a \ra \tau W_{f(a)}$.
707: For affine varieties, $V \sub K^n, W \sub K^m$, the map $\phi^{(1)}$
708: is given by $\phi^{(1)}_a((a,u)) = (\phi(a),d\phi_a(u)+\phi^\dl(a))$.
709:
710: When $f$ is a regular function on $V$, composing $f^{(1)}$ with
711: $\pi$, as above, one gets a map $\tau f = \pi\circ f^{(1)}$,
712: which we call a {\em $\tau$-differential}.
713: Below, Lemma~\ref{twder},
714: we will see that $\tau:K[V]\ra K[\tau V]$ is a derivation (in fact,
715: a $\tau$-derivation, as defined below.
716:
717: \brms
718: \ben
719: \item When the derivation $\dl$ on $K$ is trivial, then for any variety
720: $V$, $\tau V = TV$. Likewise, given any regular map between varieties,
721: $f:V \ra W$, $f\prl= Tf$. More generally, this holds true over an
722: arbitrary differential field if everything is defined over the field of
723: constants.
724:
725: \item In \cite{Ros}, Proposition 3.15, it is shown that that the map
726: $\tau$ coincides with something introduced by Buium,
727: in a different context.
728: \een
729: \erms
730:
731: The next lemma follows from the description, above, of the
732: lifting map in local coordinates.
733:
734: \bla
735: \label{taudiff}
736: Let $V$ be a variety, $f$ a regular function on $V$. For each
737: $a\in V$, $\lm(\tau f_a)=df_a$.
738: \ela
739:
740: %One can define the tangent space and prolongation of a variety by patching.
741: %Also, what about other charts? \ldots
742:
743: \sec{$\tau$-forms}
744: \label{tauforms}
745: Our main idea is to look at an analog of 1-forms on $\tau V$, which will be functions
746: $f:\tau V \lra K$ such that on each fiber of $\pi_\tau : \tau V \lra V$, $f$ is
747: an affine map, as defined above.
748: We recall the definition of 1-forms, following the presentation of
749: Shafarevich~\cite{Sha}. We then use this formalism
750: to introduce $\tau$-forms.
751: For an alternative, more general treatment of this material,
752: in the language of commutative algebra and scheme theory, see \cite{Ros}.
753:
754: Let $V$ be a variety, and let $\Phi [V]$ be the set of all functions
755: $\phi$ mapping each point $v \in V$ to a linear map
756: $\phi (v): TV_v\lra K$.
757: Note that $\Phi [V]$ naturally forms a (very large) $K[V]$-module.
758: Given $f \in K[V]$, the differential $df$ is a function in $\Phi [V]$.
759: One could look at the submodule of $\Phi[V]$ generated by
760: $\{df: f \in K[V]\}$, but this is somewhat too small.
761: Instead, we say that an element $\phi \in \Phi [V]$ is a regular differential
762: 1-form on $V$ if every $v\in V$ has a neighborhood $U$ such that the
763: restriction of $\phi$ to $U$ belongs to the $K[U]$-submodule of
764: $\Phi[U]$ generated by the elements $df, f \in K[U]$.
765:
766: The regular differential 1-forms on $V$ form a $K[V]$-module,
767: denoted $\Om [V]$.
768:
769: \bla
770: The map $d: K[V]\lra\Om[V]$ satisfies
771: \ben
772: \item $dc=0$ for all $c \in K$.
773: \item $d(f + g) = df+dg$
774: \item $d(fg)= fdg + gdf$.
775: \een
776: \ela
777:
778: \bpro
779: Every $v\in V$ has an affine neighborhood $U$ such that
780: $\Om [U]$ is a free $K[U]$-module of rank $\dim (V)$.
781: \epro
782:
783: We now introduce $\tau$-forms, imitating this construction.
784: Let $\Psi[V]$ be the set of all functions $\psi$ mapping each
785: $v \in V$ to an affine map $\psi(v): \tau V_v\lra K$.
786: As above, $\Psi [V]$ forms a $K[V]$-module and,
787: given $f \in K[V]$, the $\tau$-differential $\tau f$ is in $\Psi [V]$.
788:
789: \bdf
790: We say that $\psi\in \Psi[V]$ is a {\em regular $\tau$-form} if every $v\in V$
791: has a neighborhood $U$ such that the restriction of $\psi$ to $U$ belongs
792: to the $K[U]$-submodule of $\Psi[U]$ generated by the elements
793: $\tau f,f \in K[U]$.
794:
795: The regular $\tau$-forms on $V$ form a $K[V]$-module,
796: denoted $\Omt [V]$.
797: \edf
798:
799: Note that for any variety $V$, \tauf\ $\omt \in \Omt [V]$,
800: and $v \in V$, there is an open neighborhood $U$ of $v$ such
801: that locally, on $U$,
802: $$
803: \omt = (\sum_{i=1}^ng_i\tau f_i)
804: $$
805: for $g_i,f_i\in K[U]$.
806:
807: \brm
808: Given a variety $V$, note that a differential form on $V$ is a
809: regular function on $TV$. Thus the module $\Om[V]$ embeds
810: naturally in $K[TV]$, and the differential map is a $K$-linear derivation,
811: $d: K[V]\ra K[TV]$. Likewise, a $\tau$-form on $V$ is a regular
812: function on $\tau V$. By Lemma~\ref{twder}, below, $\tau$ is actually
813: a derivation.
814: \erm
815:
816: \brm
817: \label{iotaiota}
818: From the definition, there is a map $\tau: K[V]\lra\Omt[V]$, much
819: like the derivation map $d: K[V]\lra\Om[V]$ to 1-forms.
820: But for \taufs\ there is also a natural embedding of $K[V]$ into
821: $\Omt[V]$, which we now describe.
822: Because $K$ is differentially closed, there is a $c\in K$
823: such that $\dl(c)=1$. So $c\in K[V]$ and $\tau c\in \Omt[V]$ is
824: the constant function on $\tau V$ with value 1. By the definition
825: of $\Omt[V]$, for each $f\in K[V], f\tau c \in \Omt[V]$, where
826: $f\tau c$ is constant on each fiber $\tau V_v$, with value $f(v)$.
827: One sees immediately that the map $\iota:K[V]\lra \Omt [V]$, given by
828: $\iota(f)=f\tau c$, is an embedding (and is independent of the choice
829: of $c$). Occasionally, given $f \in K[V]$, we will also write $f \in \Omt[V]$,
830: where to be more precise we mean that $\iota f\in \Omt[V]$.
831:
832: Call a \tauf\ $\omt$ {\em trivial} if it equals $\iota f$, for some $f\in K[V]$.
833: %because it is a constant map on each fiber $\tau V_v, v \in V$.
834: (Clearly, a \tauf\ is trivial if and only if it is trivial
835: on some non-empty open subset.) Further, given $e\in K \sub K[V]$,
836: call $\iota e$ a {\em constant trivial} $\tau$-form.
837: \erm
838:
839: \bpro
840: \label{freemod}
841: Every $v\in V$ has an (affine) neighborhood $U$ such that
842: $\Omt [U]$ is a free $K[U]$-module of rank $\dim (V) + 1$.
843: \epro
844:
845: \prf
846: %\textsf{(There must also be a more direct argument. One should be
847: %able to imitate the proof that $\Om[V]$ is a locally free module,
848: %rather than using this fact in the proof.)}
849: %
850: We know that $v\in V$ has a neighborhood $U_1$ such that
851: $\Om[U_1]$ is a free $K[U_1]$-module of rank $\dim (V)$.
852: Clearly, there is a neighborhood $U_2$ of $v$ such that there is a
853: regular section $E_2$ of $\tau U_2$. In other words, there is
854: a map $\phi_2: U_2 \lra \tau U_2$ that is an isomorphism from
855: $U_2$ to $E_2$, whose inverse is the projection map
856: $\pi_\tau|_{U_2}$. Let $U=U_1\cap U_2$, $E=E_2\cap\tau U$,
857: and $\phi =\phi_2|_U$, so that $E$ is a section of $U$.
858:
859: $\tau U$ is a $TU$-torsor, and the section $E$ gives us an
860: isomorphism $\phi'$ from $TU$ to $\tau U$, which maps the
861: 0-section of $TU$ to $E$. Given $u\in TU_a$,
862: $$
863: \phi'(u)=u\cdot\phi(\pi_T(u))
864: $$
865: where $u\cdot\phi(\pi_T(u))$ denotes the action of $TU_a$
866: on $\tau U_a$.
867: (On each affine fiber, $(\tau V_a, TV_a)$, this is the
868: bijection from $TV_a$ to $\tau V_a$ that one gets by
869: fixing the point $\phi(a) \in \tau V$, as in Remark~\ref{affbasics}.2.)
870:
871: The isomorphism $\phi'$ determines a bijection $\Phi'$ between
872: $\Om[U]$ and the set $\Psi\sub\Omt[U]$ of \taufs\ on $U$
873: that take the value 0 on all of $E$. Given $\om\in\Om[U]$, let
874: $\Phi'(\om)$ be the \tauf\ on $U$ such that for $u \in \tau U$,
875: $$
876: \Phi'(\om)(u) = \om(\phi'^{-1}(u)).
877: $$
878: In fact, this bijection is an isomorphism of $K[U]$-modules.
879:
880: We want to show that $\Omt[U]=\Psi \oplus \iota K[U]$.
881: Let $\omt\in\Omt[U]$, and let $g\in K[U]$ be the function
882: $g(a)=\omt(\phi(a))$. Then $\omt_0=\omt -\iota g$ is in
883: $\Psi$, which implies that $\Omt[U]=\Psi + \iota K[U]$.
884:
885: Finally, observe that $\Psi + \iota K[U]$ is a free direct
886: sum $\Psi \oplus \iota K[U]$, making $\Omt[U]$ into a free $K[U]$-module of dimension
887: $\dim(V)+1$. For suppose that there is an $\omt \in \Psi,
888: \omt=\Phi'(\om)$, $\om\in\Om[U]$, and a $g\in K[U]$ such that
889: $\omt=\iota g$. Then $\omt$ is constant on each affine $\tau$-fiber,
890: so $\om$ must be the trivial 1-form, and $\omt$ must be identically
891: zero. Thus, $g$ is also identically zero, as desired.
892: \qed
893:
894: \msk
895:
896: The following lemma is suggestive.
897:
898: \bla
899: \label{twder}
900: The map $\tau: K[V]\lra\Omt[V]$ satisfies
901: \ben
902: \item $\tau c=\dl c$ for all $c \in K$.
903: \item $\tau(f + g) = \tau f + \tau g$
904: \item $\tau(fg)= f\tau g + g\tau f$.
905: \een
906: \ela
907:
908: \prf Conditions 1.\ and 2.\ are immediate from the definitions.
909: To prove 3., we show first that
910: $(fg)^\delta = f^\delta g + g^\delta f$. To simplify notation,
911: we assume that $f,g$ are polynomials in one variable. Let
912: $f(x)=\sum_{i=0}^ma_ix^i$ and $g(x) = \sum_{j=0}^nb_jx^j$. Then
913: $$
914: \begin{array}{lll}
915: (fg)^\delta&=&\sum_{k=0}^{m+n}\sum_{i=0}^k(a_i^\delta b_{k-i} +
916: a_ib^\delta_{k-i})x^k\\\\
917: &=&\sum_{k=0}^{m+n}\sum_{i=0}^ka_i^\delta b_{k-i}x^k +
918: \sum_{k=0}^{m+n}\sum_{i=0}^ka_ib^\delta_{k-i}x^k\\\\
919: &=&f^\delta g + fg^\delta
920: \end{array}
921: $$
922: as desired.
923:
924: Then for $(a, u) \in \tau (V)$,
925: $$
926: \begin{array}{lll}
927: \tau(fg)(a,u) &=& d(fg)_a\cdot u + (fg)^\delta(a)\\\\
928: &=& f(a)dg_a\cdot u + g(a)df_a\cdot u + f(a)g^\delta(a)+g(a)f^\delta(a)
929: \\\\
930: &=&f(a)\tau g(a,u) + g(a)\tau f(a,u)
931: \end{array}
932: $$
933: which completes the proof.
934: \qed
935:
936: \brm
937: This lemma says that the map $\tau: K[V]\lra\Omt[V]$ is a {\em $\tau$-derivation},
938: that is, a derivation such that for all $a,b \in K$, $\dl(a)\tau(b) = \dl(b)\tau(a)$.
939: In fact, this map is the fundamental example of such a derivation.
940: For more information, see~\cite{Ros}.
941: \erm
942:
943: \bco
944: Let $V$ be a variety. For any polynomial $F\in K[T_1,\ldots,T_m]$,
945: and functions $f_1,\ldots,f_m\in K[V]$,
946: $$
947: \begin{array}{lll}
948: \tau(F(f_1,\ldots,f_m))&=&\sum_{i=1}^m\frac{\partial F}{\partial T_i}
949: (f_1,\ldots,f_m)\tau f_i + F^\delta(f_1,\ldots,f_m)
950: \end{array}
951: $$
952: \eco
953: \prf
954: It suffices to consider $F(T_1,\ldots,T_m)=
955: cT_1^{a_1}\cdots T_m^{a_m}$ a monomial. Then
956: $$
957: \begin{array}{lll}
958: \tau (cf_1^{a_1}\cdots f_m^{a_m}) &=&
959: \sum_{i=1}^m (a_i f_i^{a_i-1}\tau f_i(\prod_{j\neq i}f_j^{a_j}))+
960: (\tau c)(f_1^{a_1}\cdots f_m^{a_m})
961: \\\\
962: &=&\sum_{i=1}^m\frac{\partial F}{\partial T_i}
963: (f_1,\ldots,f_m) \tau f_i+ F^\delta(f_1,\ldots,f_m)
964: \end{array}
965: $$
966: as desired.
967: \qed
968:
969: \brm
970: We now define rational 1-forms and $\tau$-forms on $V$.
971: Notice first that the above definition actually gives a sheaf
972: of modules of 1-forms and $\tau$-forms. Given any
973: open set $U \sub V$, one can define $\Om [U]$ and $\Omt [U]$
974: as above, and for open subsets $U \sub W \sub V$,
975: there are natural restriction maps $\rho_{W,U}:\Om [W] \lra \Om [U]$
976: and $\rho_{W,U}^\tau:\Omt [W] \lra \Omt [U]$.
977: Define an equivalence relation on 1-forms,
978: where $\om_1 \in \Om [U_1]$ and $\om_2 \in \Om [U_2]$, $U_i$ non-empty,
979: are equivalent if they agree on $U_1 \cap U_2$ (or on any open set).
980: A rational 1-form is an equivalence class under this relation,
981: and $\Om (V)$ denotes the set of rational 1-forms.
982: One can then easily define the domain of a rational 1-form.
983: Recall that $\Om (V)$ is a $\dim(V)$-dimensional vector space
984: over $K(V)$, the field of rational function on $V$.
985:
986: The rational $\tau$-forms $\Omt (V)$ are defined in exactly the same way
987: and also form a $K(V)$-vector space.
988: As with regular \taufs, there is a natural embedding
989: $\iota: K(V)\lra\Omt(V)$.
990: \erm
991:
992: The next result follows immediately from Proposition~\ref{freemod}.
993:
994: \bpro
995: \label{taurat}
996: Given an $n$-dimensional variety $V$, $\Omt(V)$ is an
997: $(n+1)$-dimensional $K(V)$-vector space.
998: \epro
999:
1000: From now on, by 1-form or $\tau$-form we will mean rational
1001: 1-form or $\tau$-form. Following Hrushovski-Itai~\cite{HI},
1002: regular forms will be referred to as {\em global forms}.
1003:
1004: \brm
1005: In the language of schemes, we can express Proposition~\ref{freemod} by
1006: saying that, given a smooth variety $V$, $\Omt_V$ is a locally free
1007: sheaf of dimension $\dim(V) + 1$. In fact, elements of $\Omt_V$ are
1008: exactly the rational functions on the prolongation cone $\tilde{V}$ that
1009: are linear on each fiber. Equivalently, they are sections of the dual
1010: bundle of $\tilde{V}$. This situation is analogous to the fact that
1011: 1-forms are functions on the tangent bundle and sections of the cotangent
1012: bundle.
1013: \erm
1014:
1015: \subs{The $\Lm$ map}
1016: We describe a functor $\Lambda$ from $\tau$-forms to 1-forms, that
1017: is a precise analog of the functor $\lambda$ from affine spaces to
1018: vector spaces, defined above.
1019:
1020: \bla
1021: \label{lambda}
1022: Let $V$ be an irreducible smooth variety. There is a natural map
1023: $\Lambda_V: \Omega^\tau (V) \lra \Omega (V)$, such that for each
1024: $\omega^\tau \in \Omega^\tau(V)$, and $v\in V$,
1025:
1026: $$
1027: \Lambda_V(\omega^\tau)_v = \lambda(\omega^\tau_v)
1028: $$
1029: where $\lambda(\omega^\tau_v): TV_v\lra K$ is the linear map
1030: associated to the affine map $\omega^\tau_v: \tau V_v \lra K$.
1031:
1032: Given $f \in K(V)$ and $\tau f\in\Omt(V)$,
1033: then $\Lm_V(\tau )=df\in\Om(V)$.
1034: \ela
1035:
1036: \prf
1037: Fix $\omt\in\Omt(V)$, and let $\om =\Lm_V(\omt)$. By the definition of $\Lm_V$,
1038: for each $v\in V$, $\om_v$ is a linear map on $TV_v$, so one must show
1039: that these linear maps vary smoothly on $V$. It suffices to check
1040: this locally.
1041:
1042: Let $v\in V$, and choose an open neighborhood $U\sub V$ of $v$, such
1043: that on $U$, $\omt$ is given by
1044: $\sum_{i=1}^ng_idf_i + \iota h$, $f_i,g_i,h \in K[U]$.
1045: Then $\om|_U=\sum_{i=1}^ng_idf_i$, as desired.
1046:
1047: The last assertion follows from Lemma~\ref{taudiff}.
1048: \qed
1049:
1050: \msk
1051:
1052: Notice that $\Lambda_V: \Omega^\tau(V) \lra \Omega(V)$ is clearly surjective.
1053: The map $\Lambda_V$ restricts to a map from global $\tau$-forms to global
1054: 1-forms, but this map is in general {\em not} surjective.
1055: Below, the remarks preceding Proposition~\ref{cohom} provide more information
1056: about this restricted map.
1057:
1058: Observe also that the composition of $\iota: K(V) \lra \Omt(V)$ with
1059: $\Lambda_V: \Omt(V)$ is trivial, i.e., $\Lambda_V \circ \iota$ is the zero
1060: map. As $\iota$ is injective, and $\dim(\Omt(V)) = \dim(\Om(V)) +
1061: \dim (K(V)) = n + 1$, as $K(V)$-vector spaces, one has the following
1062: proposition.
1063:
1064: \bpro
1065: For any variety $V$, there is a short exact sequence
1066: $$
1067: 0 \lra K(V) \lra \Omt(V) \lra \Om(V) \lra 0.
1068: $$
1069: \epro
1070:
1071: A description of this sequence,
1072: reformulated in terms of locally free sheaves on $V$
1073: is given in~\cite{Ros}, where it is shown that, as an extension of
1074: $\Om(V)$ by $K(V)$, it corresponds to the Kodaira-Spencer class of $V$.
1075:
1076: We now describe how to pullback \taufs. Recall that, given a
1077: rational map $\phi:V\lra W$ between varieties, there is a pullback map
1078: $\phi^*:\Om(W)\lra\Om(V)$ that can be defined as follows.
1079: Let $\om\in\Om(W)$, $v\in V$ and $y=\phi(v)$. Then for all
1080: $a\in TV_v$, $\phi^*\om_v(a)=\om_y(T\phi(a))$.
1081:
1082: \bdf
1083: Let $\phi:V\lra W$ be a rational map between varieties. For each
1084: $\omt\in\Omt(W)$, we define the {\em pullback} of $\omt$ by $\phi$, written
1085: $\phi^{\tau*}\omt$, to be the \tauf\ on $V$, given as follows.
1086: For $v\in V, y=\phi(v)$, and $a\in\tau V_v$, then
1087: $\phi^{\tau*}\omt(a)= \omt_y(\phi\prl(a))$.
1088: \edf
1089:
1090: \bla
1091: \label{comm}
1092: Let $V, W$ be varieties, $\phi: V \lra W$ a morphism. The following diagram is
1093: commutative.
1094:
1095: $$
1096: \xymatrix{
1097: \Omt (V) \ar[d]^{\Lm_V} &\Omt (W) \ar[l]_{\phi^{\tau *}} \ar[d]^{\Lm_W}\\
1098: \Om (V) &\Om (W)\ar[l]_{\phi^*}
1099: }
1100: $$
1101: \ela
1102:
1103: \prf It suffices to check this on fibers $\tau V_a$ and $TV_a$, $a \in V$,
1104: $\tau W_{\phi(a)}$ and $TW_{\phi(a)}$, $\phi(a) \in W$.
1105: We have $\phi\prl_a: \tau(V)_a \lra \tau W_{\phi(a)}$ and $T\phi_a:
1106: TV_a \lra TW_{\phi (a)}$; notice that $\lambda(\tau\phi_{a})=d\phi_{(a)}$.
1107: Then $(\phi^{\tau *}\omega^\tau)_a = \omega^\tau\circ\phi\prl_a$ and
1108: $\lambda(\phi^{\tau *}\omega^\tau)_a)=\lambda(\omega^\tau)\circ\lambda(\tau\phi_a)$.
1109: Going around the square in the other direction,
1110: $\phi^*(\lambda(\omega^\tau_{\phi (a)})) =
1111: \lambda(\omega^\tau)\circ\lambda(\tau\phi_a)$,
1112: as desired.
1113: \qed
1114:
1115: \brm
1116: By the previous lemma, one can make $\Lm$ into a functor from the
1117: category of \taufs\ to the category of 1-forms. Given $\phi:V\lra W$,
1118: we have $\Lm(\phi^{\tau*})=\phi^*$.
1119: (Alternatively, one can define $\Lm$ explicitly and fiberwise, as in
1120: Lemma~\ref{lambda}.)
1121: \erm
1122:
1123: \sec{Algebraic curves}
1124: \label{algcurves}
1125:
1126: %+++
1127:
1128: \subs{Global $\tau$-forms}
1129:
1130: In this section, by curve we mean a smooth projective curve, unless noted otherwise.
1131: Recall that $\Omega [V]$ denotes the set of global 1-forms on a variety $V$.
1132:
1133: \bdf
1134: Given a variety $V$, let $\Omega^\tau [V]$ denote the set of global $\tau$-forms
1135: on $V$, a $K$-subspace of $\Omega^\tau (V)$.
1136: \edf
1137:
1138: Note that for any variety $V$, the map $\Lambda_V:\Omega^\tau(V) \lra \Omega(V)$
1139: restricts to a map from $\Omega^\tau[V]$ to $\Omega[V]$. The next lemma is immediate.
1140:
1141: \bla
1142: \label{kern}
1143: Let $C$ be a curve.
1144: The kernel of $\Lambda_C:\Omt [C] \lra \Om [C]$, $\textup{Ker}(\Lm_C)$, is the set of
1145: $\omt \in \Omt[C]$ such that for all $c \in C$, $\omt_c$ is a constant affine
1146: map on $\tau C_c$.
1147: Thus each $\omt \in \textup{Ker}(\Lm_C)$ determines in a natural way a regular
1148: function on $C$. Since the only regular functions on $C$ are the
1149: constant functions, there is a natural isomorphism $\textup{Ker}(\Lm_C) \cong K$.
1150:
1151: More generally, the same holds for any smooth projective variety.
1152: \ela
1153:
1154: We can now describe the fibers of the map $\Lm_C : \Omt [C] \lra \Om [C]$.
1155: For all $\omt_1, \omt_2 \in \Omt [C]$, $\Lm_C (\omt_1) = \Lm_C (\omt_2)$ if and only if
1156: $\Lm_C (\omt_1 - \omt_2) = 0 \in \Om [C]$ if and only if there is an $e \in K$ such that
1157: for all $a \in \tau C$, $\omt_1(a) = \omt_2(a) + e$. In this
1158: case, we say that $\omt_1$ is a {\em translate} of $\omt_2$, or that they
1159: are {\em parallel}. Clearly, parallelism classes are 1-dimensional $K$-affine
1160: spaces and are the fibers of the map $\Lambda_C$.
1161:
1162: Recall that on a curve $C$ of genus $g$, $\Omega[C]$ is (by definition) a
1163: $g$-dimensional $K$-vector space.
1164: Here, we note that the computation of dimension of $\Omt[C]$ is somewhat
1165: more complicated, depending on the {\em Kodaira-Spencer class} of $C$.
1166:
1167: \bpro
1168: \label{globdim}
1169: Let $C$ be a curve of genus $g$. Then
1170: $1 \leq \dim_K \Omt[C] \leq g+1$.
1171: Further, $\dim_K \Omt[C] = g + 1$ if and only if
1172: $C$ is isomorphic to a curve defined over the field $k$ of constants.
1173: \epro
1174:
1175: \prf
1176: The first statement follows immediately from the above remarks as, given the
1177: map $\Lm_C : \Omt [C] \lra \Om [C]$, it follows that $\dim_K \Omt [C]
1178: \leq \dim_K \Om[C] + \dim_K \Ker (\Lm_C) = g +1$. The other inequality
1179: follows from the natural embedding $\iota$ of $K[C] \cong K$ into $\Omt[C]$.
1180: The proof of the second statement,which is equivalent to Proposition~\ref{cohom},
1181: uses scheme-theoretic machinery developed below.
1182: \qed
1183:
1184: \msk
1185:
1186: For curves of genus 1, this yields a complete description of the global $\tau$-forms.
1187:
1188: \bco
1189: Let $C$ be a curve of genus 1. If $C$ is isomorphic to a curve defined over $k$,
1190: then $\dim_K \Omt[C] =2$. Otherwise, $\dim_K\Omt[C]=1$. Further, in both cases,
1191: we can give an explicit description of the global $\tau$-forms.
1192: \eco
1193:
1194: \prf
1195: Suppose first that $C$ is defined over $k$. Then, $\tau C \cong TC$, the tangent
1196: variety. In this case, it is clear that any \tauf\ can be written (uniquely) as a sum of
1197: a 1-form and of a \tauf\ of the form $\iota f, f \in K(C)$ (as defined in
1198: Remark~\ref{iotaiota}), and likewise for global forms.
1199: Thus $\dim_K\Omt[C] = \dim_K\Om[C] + \dim_K K[C] = 1+1 = 2$.
1200: Otherwise, if $C$ is not defined over $k$, then $\dim_K \Omt[C] = 1$, by the above
1201: proposition. Here, $\Omt[C] = \{\iota e | e\in K, \iota e$ a constant trivial
1202: $\tau$-form$\}$.
1203: \qed
1204:
1205: \msk
1206:
1207: Recall that Hrushovski and Itai call a 1-form on a curve {\em essential} if it is
1208: not the pullback of any other 1-form, and show that there are many essential
1209: global 1-forms on any curve of genus $\geq 2$.
1210:
1211: \bpro
1212: \label{hrit}
1213: \textup{(Hrushovski-Itai)} Let $C$ be a curve of genus $\geq 2$,
1214: defined over $k$. There exists a finite or countable union
1215: ${\mathcal{E}} = \cup_lS_l$ of proper subspaces of $\Om [C]$, such that any 1-form in
1216: $\Om [C] \backslash {\mathcal{E}}$ is essential.
1217: There exists an essential global 1-form, defined over $k$.
1218: \epro
1219:
1220: Here, we are unable to determine whether an analgous statement holds for global $\tau$-forms.
1221: Another natural question is whether the global \taufs\ on a curve of positive genus
1222: are exactly the pullbacks of invariant global \taufs\ on its Jacobian, as is true
1223: of 1-forms.
1224:
1225: \subs{Rational $\tau$-forms}
1226:
1227: In this section, we introduce an equivalence relation on
1228: $\tau$-forms that will be useful for the study of strongly minimal
1229: sets in the next section. Here, there is no need to assume that
1230: curves are projective.
1231:
1232: \bdf
1233: Let $C$ be a curve, $\pi_\tau: \tau C \ra C$ the canonical projection.
1234: Let $\omt$ be a \tauf\, on a curve $C$. For each $c\in C$, $\omt_c$
1235: is either
1236: (i) everywhere defined on $\pi_\tau^{-1}(c)$, and a bijective map to $K$,
1237: (ii) everywhere defined on $\pi_\tau^{-1}(c)$, but a constant map to $K$,
1238: (iii) everywhere undefined on the affine fiber $\pi_\tau^{-1}(c)$.
1239: In case (ii), say that $c$ is a {\em zero} of $\omt$ (where zero here means that
1240: given the affine map $\omt_c:\pi_\tau^{-1}(c)\lra K$, the corresponding
1241: linear map $\lm(\omt_c): K \lra K$ is the zero map).
1242: In case (iii), say that $c$ is a {\em pole} of $\omt$.
1243:
1244: Let $Z_{\omt}$ denote the set of zeros of $\omt$, and
1245: $P_{\omt}$ its set of poles.
1246: \edf
1247:
1248: Observe that for any $\omt \in \Omt (C)$, either $Z_{\omt}$ is finite
1249: or $\omt$ is trivial, in which case $Z_{\omt}$ is an open subset of $C$.
1250:
1251: \bdf
1252: Let $C$ be a curve, $\omt \in \Omt (C)$ non-trivial.
1253: The {\em null set} of $\omt$ is
1254: $$
1255: N_{\omt} = \{a \in \tau C | \omt(a) = 0\}.
1256: $$
1257: This is a rational section of the algebraic variety $\tau C$,
1258: and thus birational to $C$.
1259:
1260:
1261: Say that $\omt_1, \omt_2 \in \Omt (C)$ are {\em $\seq$-equivalent},
1262: written $\omt_1\seq\omt_2$, if
1263: $N_{\omt_1}$ and $N_{\omt_2}$ are almost equal.
1264:
1265: Given a \tauf\ $\omt$, let $\omt / \sim$ denote its $\seq$-class.
1266: \edf
1267:
1268: \bla
1269: \label{normform}
1270: Let $C$ be a curve, and $\omt \in\Omt (C)$ a non-trivial $\tau$-form.
1271: For every $\omt_1 \in \Omt(C)$, there are unique $f,g\in K(C)$ such
1272: that $\omt_1=f\omt+\iota g$.
1273: \ela
1274:
1275: \prf
1276: (The idea is similar to the proof of Proposition~\ref{freemod}.)
1277: If $\omt_1$ is trivial, i.e., there is a $g \in K(C)$ such that
1278: $\omt_1 = \iota g$, then $\omt_1 = 0\omt+\iota g$. So we may
1279: suppose that $\omt_1$ is non-trivial.
1280:
1281: Assume first that $\omt$ and $\omt_1$ are $\seq$-equivalent.
1282: Choose some small enough quasi-affine neighborhood of $U \sub C$
1283: such that
1284: \ben
1285: \item $U$ is disjoint from all the zeros and poles of $\omt$ and $\omt_1$;
1286: \item $N_{\omt}\cap U = N_{\omt_1 }\cap U$;
1287: \item on $U$, $\omt$ and $\omt_1$ are of the form
1288: $$
1289: \omt = \sum_{i=1}^mg_i\tau f_i
1290: $$
1291: %
1292: $$
1293: \omt_1 = \sum_{j=1}^n\gamma_j\tau \phi_j
1294: $$
1295: with $g_i,f_i,\gamma_j,\phi_j, \in K[U]$.
1296: \een
1297: In particular, $\omt$ and $\omt_1$ are regular functions on the quasi-affine
1298: variety $\tau U$. Furthermore, on each fiber $\tau U_u$,
1299: there is a single point $x \in N_{\omt} = N_{\omt_1}$ such that
1300: $\omt(x)=\omt_1(x)=0$. Since $\omt_u$ and $\omt_{1,u}$ are non-constant
1301: affine maps on $\tau U_u$ with the same zero, one must be a
1302: constant multiple of the other. That is, there is an $e_u\in K$,
1303: such that $e_u\omt_u=\omt_{1,u}$. From the definition of $\omt$ and
1304: $\omt_1$ it is clear that the function $e$ on $U$, $e(u) = e_u$,
1305: is a regular function on $U$. So $\omt_1=e\omt$, as desired.
1306:
1307: We now consider the general case. There is a finite set $A \sub N_{\omt}$, such that
1308: $N_{\omt} - A$ is a regular section of $\tau U$, so there is an isomorphism
1309: $\psi:U\lra N_{\omt} - A$. Let $g\in K[U]$ be the regular function
1310: $g=\omt_1\circ \psi$, and define $\omt_2 = \omt_1 - \iota g$.
1311: Then $\omt$ and $\omt_2$ are $\seq$-equivalent, so there is an
1312: $f \in K[U]$ such that $\omt_2=f\omt$. Finally,
1313: $\omt_1 = f\omt + \iota g$, as desired.
1314: \qed
1315:
1316: \brm
1317: Note that this lemma gives another proof that
1318: $\Omt(C)$ is a 2-dimensional $K(C)$-vector space,
1319: which also follows from Proposition~\ref{taurat}.
1320: \erm
1321:
1322: \bco
1323: \label{seqcl}
1324: Let $C$ be a curve, $\omt_1, \omt_2 \in \Omt(C)$, both non-trivial.
1325: Then $\omt_1$ and $\omt_2$ are $\seq$-equivalent if and only if there is a rational
1326: function $f\in K(C)$ such that $\omt_1=f\omt_2$.
1327:
1328: Given a \tauf\ $\omt$, the class $\omt / \sim$ is a 1-dimensional
1329: subspace of $\Omt(C)$.
1330: \eco
1331:
1332: \prf
1333: One direction of the first claim follows immediately from the above proof.
1334: In the other, suppose that $\omt_1=f\omt_2$. Then on any open set $U\sub C$
1335: disjoint from the zeros and poles of $f, \omt_1$, and $\omt_2$, $N_{\omt_1}|_U
1336: =N_{\omt_2}|_U$.
1337:
1338: The second claim follows from the the first and the observation that
1339: $\omt / \sim$ is closed under addition.
1340: \qed
1341:
1342: \bla
1343: \label{sigmanull}
1344: Let $C$ be a curve, $\omt_1, \omt_2 \in \Omt(C)$, both non-trivial.
1345: Suppose that $N_{\omt_1} \cap N_{\omt_2}$ is infinite. Then
1346: $\omt_1$ and $\omt_2$ are $\seq$-equivalent. In other words,
1347: if $N_{\omt_1} \cap N_{\omt_2}$ is infinite, then
1348: $N_{\omt_1}$ and $N_{\omt_2}$ are almost equal.
1349: \ela
1350:
1351: \prf
1352: $N_{\omt_1}$ and $N_{\omt_2}$ are each rational
1353: sections of $\tau C$, so each is a curve on the surface $\tau C$.
1354: If these curves have infinite intersection, they must be equal, up
1355: to a finite set.
1356: \qed
1357:
1358: \msk
1359:
1360: Since we will be concerned with $\sim$-equivalence classes of
1361: \taufs, the following relativized versions of global and essential \taufs\
1362: are important for the main results about strictly minimal sets below.
1363: Equivalent, perhaps more natural, definitions of
1364: these notions, in the language of schemes, are given in Section~\ref{tauinv}.
1365:
1366: \bdf Let $C$ be a smooth projective curve, $\omt$ a rational \tauf.
1367: Say that $\omt$ is $\sim$-{\em global} if it is $\sim$-equivalent to
1368: a global \tauf, in which case we also say that the class $\omt / \sim$ is
1369: {\em global}.
1370: Say that $\omt$ is $\sim$-{\em essential} if every
1371: $\sim$-equivalent \tauf\ is essential.
1372: \edf
1373:
1374: \brm
1375: Let $C, C'$ be curves, and $f:C \lra C'$ a morphism between them.
1376: It is easy to see that for any $\sim$-equivalent \taufs\ $\omt_1$
1377: and $\omt_2$ on $C'$, then $f^*\omt_1$ and $f^*\omt_2$
1378: are $\sim$-equivalent on $C$.
1379: \erm
1380:
1381: \brm
1382: Let $\omt$ be a non-trivial global \tauf\ on a curve $C$ of genus $g \geq 1$.
1383: Then for any $c \in K$, $\omt_c : = c + \omt$ is global and not
1384: $\sim$-equivalent to $\omt$. In particular, there is a family
1385: of global $\sim$-classes of $\tau$-forms indexed by $K$.
1386: \erm
1387:
1388: Recall that we defined two global \taufs\ on a curve $C$ to be parallel
1389: if they are both in the same fiber of the map $\Lm_C$. This notion also
1390: makes sense for rational \taufs.
1391:
1392: \bdf
1393: Let $C$ be a curve. We say that two \taufs\ $\omt_1$ and $\omt_2$ are
1394: {\em parallel}, written $\omt_1\teqq\omt_2$, if there is a $g\in K(C)$ such that
1395: $\omt_2=\omt_1+g$. Equivalently, $\Lm_C(\omt_1) = \Lm_C(\omt_2)$.
1396: \edf
1397:
1398: Parallelism is clearly an equivalence relation.
1399:
1400: \bla Fix a curve $C$.
1401: \ben
1402: \item The trivial \taufs\ on $C$ are a parallelism class.
1403: \item On the non-trivial \taufs, $\seq$-equivalence and parallelism
1404: are cross-cutting equivalence relations. More precisely, for
1405: any non-trivial $\omt_1, \omt_2$,
1406: $|\omt_1/ \seq \cap \omt_2/ \teqq|=1$.
1407: \een
1408: \ela
1409: \prf
1410: Immediate.
1411: \qed
1412:
1413: \para{The $\Lm$ map on curves}
1414: Fix a curve $C$, and a non-trivial \tauf\ $\omt_0\in\Omt(C)$.
1415: By Lemma ~\ref{normform}, every $\omt\in\Omt(C)$
1416: is equal to $f\omt_0+g, f,g\in K(C)$, and, by Corollary~\ref{seqcl}
1417: the $\seq$-equivalence class of $\omt_0$ is
1418: $\omt_0/ \seq=\{f\omt_0:f\in K(C)-\{0\}\}$.
1419:
1420: The following lemma can be easily verified.
1421:
1422: \bla
1423: Let $C$ be a curve, $f\in K(C)$, and $\omt$ a non-trivial $\tau$-form on $C$.
1424: Then $\Lm_C(f\omt)=f\Lm_C(\omt)$.
1425: %Also, $\Lm_C(f(\omt-1)+1)=f\Lm_C(\omt)$
1426: %
1427: Thus the map $\Lm_C$ induces a bijection between the
1428: equivalence class $\omt/ \seq$ and the non-trivial 1-forms.
1429: %In particular, $\Lm_C$ is surjective.
1430: %Also, $\Lm_C$ induces a bijection between the
1431: %equivalence class $\omt/_\teq$ and the non-trivial 1-forms.
1432: \ignore{
1433: \item For any $\omt_1,\omt_2\in\Omt(C)$, $\Lm_C(\omt_1)=\Lm_C(\omt_2)$ if
1434: and only if $\omt_1\teqq\omt_2$. Equivalently, for any $\om\in\Om(C)$, the
1435: fiber $\Lm^{-1}_C(\om)$ is a parallelism class.
1436: }
1437: %end{ignore}
1438: \ela
1439:
1440: \sec{Order one strongly minimal sets}
1441: \label{sm}
1442:
1443: We aim to classify order one strongly minimal sets, up to
1444: non-orthogonality.
1445: Hrushovski proved that every such set is either
1446: trivial or non-orthogonal to the constant field $k$, so we will
1447: be investigating trivial sets.
1448: We first briefly recall the analysis from Hrushovski-Itai of order
1449: one \sm sets over the constants.
1450:
1451: Say that two strongly minimal sets $X$ and $Y$ are {\em birationally
1452: isomorphic}, or {\em birational}, if there is an almost everywhere defined
1453: bijective map from $X$ to $Y$. Any order one strongly minimal set $Y$ is
1454: birational to one of the form
1455: $$
1456: X = \Xi(C, s) = \{ x \in C : \dl x = s(x)\}
1457: $$
1458: where $s: C \lra \tau (C)$ is a rational section of the prolongation.
1459: In this case, one says that $X$ lives on $C$, and the differential
1460: function field of the Kolchin closed set $X$ equals the function
1461: field of $C$.
1462:
1463: Suppose now that $C$ is a curve defined over $k$, so $\tau C = TC$,
1464: the tangent space of $C$. Then the \sm sets $X$ defined over $k$
1465: and living on $C$ are given as $X = \Xi (C, s)$, $s$ a
1466: rational vector field also defined over $k$. There is a natural
1467: 1-1 correspondence between such vector fields $s$ and 1-forms
1468: $\om$, given by the relation $\om(x)s(x) = 1$ for almost all $x$.
1469: So order one \sm sets over the constants also correspond to sets
1470: $$
1471: X = \Xi(C, \om) = \{a \in C : \om(a)\dl a=1\}
1472: $$
1473: where this is taken to include a pole $p$ of $\om$ if $\dl p = 0$.
1474: (This last condition ensures that $\Xi(C, \om)$ is Kolchin closed.)
1475:
1476: In this setting, Hrushovski and Itai analyze \sm sets in terms of
1477: pairs $(C, \om)$, $C$ a smooth projective curve and $\om \in \Om (C)$,
1478: both defined over $k$. We now show how to extend their results
1479: to all curves, using $\tau$-forms.
1480:
1481: \bdf
1482: Let $C$ be a smooth projective curve and $\omt \in \Omt (C)$ a
1483: non-zero $\tau$-form. Let $P \sub C$ be the set of poles $p$ of $\omt$
1484: such that $\dl p=0$, and define
1485: $$
1486: \Xi (C, \omt) = \{a \in C : \omt(a)\dl a = 0\} \cup P
1487: $$
1488: \edf
1489: As above, one can check that $\Xi (C, \omt)$ is always Kolchin closed.
1490:
1491: Observe that $X = (C, \omt) = \{a \in C : \omt(a)\dl a = 0\} $ is a strongly minimal
1492: set. Indeed, let $s$ be a section of $\tau C$ so that $\omt s = 1$ almost
1493: everywhere. Then $\Xi(C, s) = \Xi(C, \omt)$, up to a finite set.
1494: On the other hand,
1495: this is no longer a 1-1 correspondence, as many $\tau$-forms correspond to the
1496: same section and therefore determine the same \sm set.
1497:
1498: \bla
1499: Let $C$ be a curve, $\omt_1, \omt_2 \in \Omt(C)$, both non-trivial.
1500: Then $\omt_1$ and $\omt_2$ are $\seq$-equivalent if and only if
1501: $\Xi(C, \omt_1)$ and $\Xi(C, \omt_2)$ are almost equal.
1502: \ela
1503:
1504: \prf
1505: From the definition, $\Xi(C,\omt)$ only depends on $N_{\omt}$,
1506: so if $\omt_1$ and $\omt_2$ are $\seq$-equivalent, then
1507: $\Xi(C, \omt_1)$ and $\Xi(C, \omt_2)$ are almost equal.
1508:
1509: In the other direction, if
1510: $\Xi(C, \omt_1)$ and $\Xi(C, \omt_2)$ are almost equal,
1511: then $N_{\omt_1}\cap N_{\omt_2}$ is infinite, so
1512: $\omt_1$ and $\omt_2$ are $\seq$-equivalent, by
1513: Lemma~\ref{seqcl}.
1514: \qed
1515:
1516: \bco
1517: \label{formz}
1518: There is a 1-1 correspondence between $\seq$-equivalence
1519: classes of \taufs\, in $\Omt (C)$ and strongly minimal sets living on $C$.
1520: \eco
1521: Notice that we do not have to treat the case $C(k) \sub C$, corresponding
1522: to $s(x)=0$ separately.
1523:
1524: The following series of lemmas provides versions of
1525: Lemmas 2.4 and 2.6 -- 2.9 of \cite{HI}, for \taufs.
1526: (Lemma 2.5 becomes false.) The proofs are the same.
1527:
1528: For the proof of the next lemma, we recall the following easy fact.
1529: Given $\phi:V\lra W$, and $v\in V$, $(\tau \phi)(\dl v)=\dl (\phi(v))$.
1530:
1531: \bla
1532: \label{taupullback}
1533: Let $g: C_1 \lra C_2$ be a dominant regular map between nonsingular curves, and
1534: $\omt_2$ a \tauf\, on $C_2$. Let $\omt_1 = g^{\tau *}\omt_2$ be the pullback of
1535: $\omt_2$ by $g$ to $C_1$. Then
1536: $g^{-1}\Xi(C_2, \omt_2) = \Xi(C_1, \omt_1)$.
1537: \ela
1538:
1539: \prf
1540: Let $c_1\in C_1, c_2=g(c_1)$. Then
1541: $$
1542: \omt_1(c_1)\dl c_1=\omt_2(c_2)(g\prl)(\dl c_1)=\omt_2(c_2)\dl
1543: (g(c_1))=\omt_2(c_2)\dl c_2.
1544: $$
1545: Thus $\omt_1(c_1)\dl c_1=0$ if and only if $\omt_2(c_2)\dl c_2=0$.
1546: \qed
1547:
1548: \bla
1549: \label{infint}
1550: If $\Xi(C, \omt_1)$ and $\Xi(C, \omt_2)$ have infinite intersection,
1551: $\omt_1 $ and $\omt_2$ are $\seq$-equivalent.
1552: \ela
1553:
1554: \prf
1555: For infinitely many points $c$ in the intersection $\Xi(C, \omt_1)
1556: \cap \Xi(C, \omt_2)$, $\omt_1(c)\dl c=0=\omt_2(c)\dl c$. Thus,
1557: $N_{\omt_1}$ and $N_{\omt_2}$ have infinite intersection, and are
1558: therefore $\seq$-equivalent, by Lemma~\ref{sigmanull}.
1559: \qed
1560:
1561: \bla
1562: \label{compat}
1563: Let $g:C_1 \lra C_2$ be a dominant rational map between non-singular curves.
1564: Let $\omt_i$ be a \tauf\, on $C_i, i = 1,2$. Assume
1565: $g\Xi(C_1, \omt_1) \sub \Xi(C_2, \omt_2)$ (perhaps up to a finite set).
1566: Then $\omt_1$ and $g^{\tau *}\omt_2$ are $\seq$-equivalent.
1567: \ela
1568:
1569: \prf
1570: By Lemma~\ref{taupullback}, $g^{-1}\Xi(C_2, \omt_2) = \Xi(C_1, g^{\tau *}\omt_2)$
1571: (a.e.). Thus the intersection of $\Xi(C_1, \omt_1)$ with $\Xi(C_1, g^{\tau *}\omt_2)$
1572: is infinite. By the previous lemma, $\omt_1$ and $g^{\tau*}\omt_2$
1573: are $\seq$-equivalent.
1574: \qed
1575:
1576: \bla
1577: Let $C_1$ be a complete nonsingular curve, $\omt_1$ a global \tauf\, on $C_1$.
1578: Let $g:C_1 \lra C_2$ be a rational function into a complete nonsingular curve,
1579: and $\omt_2$ a \tauf\, on $C_2$ such that $g\Xi(C_1, \omt_1) =
1580: \Xi(C_2, \omt_2)$. Then $\omt_1$ is a $\sim$-global \tauf\ on $C_2$.
1581: \ela
1582:
1583: \prf
1584: As pointed out in \cite{HI}, $g: C_1 \lra C_2$ is a surjective morphism.
1585: It then follows from definitions that $\omt_2$ is a global \tauf\, on $C_2$
1586: if and only if its pullback, $g^{\tau *}\omt_2$ is a global \tauf\, on $C_1$.
1587: By Lemma~\ref{taupullback}, $g^{-1}\Xi(C_2, \omt_2)=\Xi(C_1, g^{\tau *}\omt_1)$,
1588: so $\Xi(C_1, g^{\tau *}\omt_2) \cap \Xi(C_1, \omt_1)$ is infinite.
1589: By Lemma~\ref{infint}, $g^{\tau *}\omt_2 \sim \omt_1$, so by
1590: the above remark, $\omt_1$ is a $\sim$-global form.
1591: \qed
1592:
1593: \bla
1594: \label{noeqrel}
1595: Let $\omt$ be a $\sim$-essential \tauf\ on a curve $C$. Let $E$ be a
1596: definable equivalence relation on $\Xi(C,\omt)$, with finite classes.
1597: Then almost every class of $E$ has one element.
1598: \ela
1599:
1600: \prf
1601: Suppose not. By \cite{HI}, Lemma 1.2, there is a curve $C_0$ and a map
1602: $g: C \ra C_0$ with deg$(g) > 1$ such that $g(\Xi(C,\omt))$ lives on $C_0$.
1603: Thus there is a \tauf\ $\omt_0$ on $C_0$ such that $g(\Xi(C,\omt))=
1604: \Xi(C_0,\omt_0)$, up to a finite set. But by Lemma~\ref{compat},
1605: $\omt$ and $g^{\tau*}(\omt_0)$ are $\sim$-equivalent, contradicting the
1606: assumption that $\omt$ is $\sim$-essential.
1607: \qed
1608:
1609: \bla
1610: \label{nonorthcon}
1611: Let $C$ be a curve of genus $\geq 1$ and $\omt$ a \tauf\ on $C$.
1612: If $\Xi(C,\omt)$ is non-orthogonal to the constants, then there is a
1613: regular map $g:C \ra \P^1$ with $g(\Xi(C,\omt))=k$, up to a finite set.
1614: In particular, there is a \tauf\ $\omt_0$ on $\P^1$ such that
1615: $\Xi(\P^1,\omt_0) = k$ and $g^{\tau*}\omt_0$ is $\sim$-equivalent
1616: to $\omt$ on $C$. Thus, $\omt$ is not $\sim$-essential.
1617: \ela
1618:
1619: \prf
1620: Let $X = \Xi(C,\omt)$ be non-orthogonal to the constants. Then there is
1621: a definable differential rational function $f$ with $f(X) \sub k$, up to a
1622: finite set (e.g., see \cite{HI}, proof of Lemma 2.10). Since for all
1623: $x \in X, \dl(x)=s(x)$, for some polynomial $s$, we can assume that
1624: $f$ is a rational function. Thus $f$ extends to a regular, dominant map
1625: $f:C \ra \P^1$, proving the first claim. The rest of the lemma follows
1626: from Lemma~\ref{compat}.
1627: \qed
1628:
1629: \msk
1630:
1631: We are now able to generalize a central result of \cite{HI}. Recall
1632: that a strongly minimal set $X$ defined over a set $A$ is {\em strictly minimal}
1633: over $A$ if $X \cap \acl(A) = \emptyset$ and, for all $a \in X, \acl(aA) \cap X
1634: = \{a\}$. We say that $X$ is strictly minimal if it is strictly minimal over
1635: some set over which it is defined. A strongly minimal set $X$
1636: is {\em $\omega$-categorical} if for any finite set $B \sub X$, $\acl_X(B)$
1637: is finite. Hrushovski's theorem that all order one strongly minimal\ sets are
1638: $\omega$-categorical~\cite{HJ} plays an important role in the proof of the
1639: following result of \cite{HI}.
1640:
1641: \bpro
1642: \label{strictmin}
1643: Let $C$ be a complete nonsingular curve over $k$ of genus $>1$. Let $\om$
1644: be an essential 1-form on $C$, defined over $k$. Then, after perhaps removing
1645: a finite set, $\Xi(C,\om)$ is strictly minimal, with trivial induced
1646: structure. Two such sets $\Xi(C_1,\om_1)$
1647: and $\Xi(C_2,\om_2)$ are non-orthogonal if and only if there exists an isomorphism
1648: $g:C_1 \ra C_2$ with $\om_1 = g^*\om_2$.
1649: \epro
1650: (In \cite{HI}, they state this result for global 1-forms, which implies
1651: additionally that $\Xi(C,\om)$ has no points in the algebraic closure of
1652: the parameters.)
1653:
1654: We now state one of our main results.
1655:
1656: \bthm
1657: \label{genprop}
1658: Let $C$ be a smooth projective curve of genus $g \geq 1$. Let $\omt$ be
1659: a $\sim$-essential \tauf. Then there is a finite set $A \sub \Xi(C,\omt)$
1660: such that $\Xi(C,\omt) - A$ is trivial strictly minimal. Further, for any
1661: definable equivalence relation with finite classes on $\Xi(C,\omt)$,
1662: over any set of parameters, almost every class has one element.
1663:
1664: Two such sets, $\Xi(C_1,\omt_1)$ and $\Xi(C_2,\omt_2)$ are
1665: not orthogonal if and only if there is an isomorphism $f:C_1 \ra C_2$ such that
1666: $f^{\tau *}\omt_2$ is $\sim$-equivalent to $\omt_1$.
1667: \ethm
1668:
1669: \prf
1670: By Lemma~\ref{nonorthcon}, $X = \Xi(C,\omt)$ is a trivial strongly minimal set and
1671: by Lemma~\ref{noeqrel}, for any definable equivalence relation with finite classes
1672: on $X$, over any set of parameters, almost every class has one element.
1673: Since $X$ is $\omega$-categorical, the set $A$ of elements of $X$ which are
1674: algebraic over the parameters defining $C$ and $\omt$ is finite. Thus
1675: $X - A$ is strictly minimal.
1676:
1677: Suppose $\Xi(C_1,\omt_1)$ and $\Xi(C_2,\omt_2)$ are non-orthogonal.
1678: After removing a finite set of points from each, one is left with two
1679: strictly minimal sets with trivial geometry, so there is a definable
1680: bijection between them. Thus there is a birational map $g: C_1 \ra C_2$,
1681: with $\omt_1 \sim g^{\tau *}\omt_2$, which is an isomorphism.
1682: \qed
1683:
1684: \brm
1685: In a sense, this completely classifies the trivial strictly minimal sets,
1686: in terms of $\sim$-essential forms, but it is not very explicit.
1687: In particular, it leaves open the question whether there are any such forms.
1688: For curves over $k$, one gets an affirmative answer from \cite{HI}, but the general
1689: case remains open.
1690:
1691: A key part of Hrushovski and Itai's proof that there are many trivial strictly
1692: minimal sets living on every curve defined over $k$ of genus $>1$,
1693: which are all orthogonal, is the fact that on such curves generic global
1694: 1-forms are essential. Proving an analogous statement for \taufs\ seems difficult,
1695: since the dimension of global \taufs\ does not depend only on the genus.
1696: Further, in the more general case, one needs to consider $\sim$-essential \taufs,
1697: which are essentially equivalence classes. For these reasons, it is not
1698: clear how one might be able to adapt ideas from \cite{HI} to answer this question.
1699: \erm
1700:
1701: Let us call an equivalence relation on a set $X$ {\em trivial} if it is definable
1702: in the language of equality. Equivalently, there is either a cofinite equivalence class
1703: or all but finitely many elements are in classes of size 1.
1704:
1705: In the statement of Proposition~\ref{strictmin}, there are no non-trivial equivalence
1706: relations on $\Xi(C,\omega)$ defined over $k$. One might ask whether it is possible
1707: to define such a relation using parameters from $K$. If this were the case, then
1708: one would have pairs $(C,\omega)$ and $(C_0,\omt_0)$, $C \not\cong C_0$,
1709: $C$ a curve defined over $k$, $\omega$ an essential
1710: 1-form defined over $k$, $\Xi(C,\omega)$ trivial strictly minimal over $k$, and
1711: $\Xi(C_0,\omt_0)$ strictly minimal over $K$, with a finite-to-one map from
1712: $\Xi(C,\omega)$ to $\Xi(C_0,\omt_0)$. Below, we argue that this can not in fact
1713: happen. The key point is the following lemma.
1714:
1715: \bla
1716: \label{eqprm}
1717: Let $K$ be an $\om$-stable structure, and let $X$ be a trivial strictly minimal set in $K$
1718: defined over a set $A$. Then no non-trivial equivalence relation on $X$ is definable
1719: with parameters from $K$.
1720: \ela
1721:
1722: \prf
1723: As usual, we may assume that the set $A$ over which $X$ is defined is empty.
1724: Suppose for contradiction that there is a tuple $\abar$ and a formula $\phi(\xbar,y,z)$
1725: such that $\phi(\abar,y,z)$ defines a non-trivial equivalence relation $E$ on $X$.
1726: By strong minimality of $X$, it is clear that there is a finite $n$ such that every
1727: class in $E$ is of size $\leq n$ and all but finitely many classes are of the same
1728: size $m > 1$. Let $\theta(\xbar)$ be a formula true of $\abar$ such that for all
1729: $\bbar$, if $K \models \theta(\bbar)$, then $\phi(\bbar,y,z)$ defines an equivalence relation
1730: on $X$ such that all but finitely many elements of $X$ belong to a class of size $m$.
1731:
1732: Let $T$ be the set of equivalence relations $T = \{\phi(\bbar,y,z) | K \models \theta(\bbar)\}$.
1733: Note that, even without elimination of imaginaries in $K$, one can treat $T$ as a definable
1734: set that can be quantified over in a natural way. Observe also that $T$ is infinite
1735: as follows. For if not, choose an $E \in T$ and a pair $(p,q)$ in $E$.
1736: Then $q \in \acl (p)$, contradicting the fact that $X$ is trivial strictly minimal.
1737:
1738: Let $\Sigma$ be the set of finite sequences of 0's and 1's, partially ordered such that,
1739: for $\rho, \sigma \in \Sigma$,
1740: $\rho \leq \sigma$ if and only if $\rho$ is an initial segment of $\sigma$.
1741: For each $\sigma \in \Sigma$, we define a pair $\pi_\sigma = (p_\sigma,q_\sigma)$
1742: of elements in $X$ and a `definable' set $T_\sigma \sub T$ with the following properties.
1743: \ben
1744: \item $T_\sigma$ is infinite and definable over parameters
1745: $\cup_{\rho < \sigma}\{p_\rho,q_\rho\}$.
1746:
1747: \item $T_\sigma$ is the disjoint union of $T_{\sigma,0}$ and $T_{\sigma,1}$.
1748:
1749: \item $T_{\sigma,0}:=\{E \in T_\sigma | E(p_\sigma,q_\sigma)\}$ and
1750: $T_{\sigma,1}:=\{E \in T_\sigma | \neg E(p_\sigma,q_\sigma)\}$ are both non-empty.
1751: \een
1752: Since it is clear that such a set of $T_\sigma$'s contradicts $\omega$-stability,
1753: this will yield the desired contradiction.
1754:
1755: We proceed by induction.
1756: Let $T_\emptyset = T$. Choose a pair $\pi_\emptyset=(p_\emptyset,q_\emptyset)$ such that
1757: there are $E,E' \in T_\emptyset$
1758: with $E(p_\emptyset,q_\emptyset)$ and $\neg E'(p_\emptyset,q_\emptyset)$.
1759: (If this were not possible, then $T$ would only contain one element.)
1760:
1761: Suppose now that for all $\rho \leq \sigma$, $\pi_\rho$ and $T_\rho$ have been defined.
1762: Let $T_{\sigma,0} = \{E \in T_\sigma| E(p_\sigma,q_\sigma)\}$ and
1763: $T_{\sigma,1} = \{E \in T_\sigma| \neg E(p_\sigma,q_\sigma)\}$.
1764: By construction, both $T_{\sigma,0}$ and $T_{\sigma,1}$ are non-empty.
1765: In fact, we claim that both sets must be infinite. For if not, it is easy to see
1766: that $q_\sigma \in \acl(\{p_\rho,q_\rho | \rho < \sigma\}\cup\{p_\sigma\})$,
1767: contradicting the fact that $X$ is trivial strictly minimal.
1768: Now choose $\pi_{\sigma,i}=(p_{\sigma,i},q_{\sigma,i})$, $i = 0,1$, disjoint from all
1769: $\pi_\rho$, $\rho \leq \sigma$, such that
1770: there are $E_i, F_i \in T_{\sigma,i}$ such that $E_i(p_{\sigma,i},q_{\sigma,i})$
1771: and $\neg F_i(p_{\sigma,i},q_{\sigma,i})$.
1772:
1773: It is clear that this construction has the required properties.
1774: \qed
1775:
1776: \bco
1777: Let $C$ be a complete nonsingular curve over $k$ of genus $>1$. Let $\om$
1778: be an essential 1-form on $C$, defined over $k$, so that after perhaps removing
1779: a finite set, $\Xi(C,\om)$ is strictly minimal with trivial induced
1780: structure. Then there is no non-trivial equivalence relation definable on $\Xi(C,\om)$,
1781: even with arbitrary parameters.
1782: \eco
1783:
1784: \prf
1785: By Proposition~\ref{strictmin} and Lemma~\ref{eqprm}.
1786: \qed
1787:
1788: \section{Another description of $\tau$-forms}
1789: \label{tauinv}
1790:
1791: %%% Redo intro
1792:
1793: In this section, we give a new algebraic characterization of the
1794: strongly minimal sets living on a curve $C$, using the theory of $\tau$-differentials
1795: developed in \cite{Ros}. Of course, it is equivalent
1796: to the earlier formulation, but has the advantage that it does not involve
1797: considering equivalence classes, so it is easier to work with.
1798: Below, we use the language of schemes, so a variety is an integral separated
1799: scheme of finite type over an algebraically closed field.
1800: We continue to assume that curves are smooth and projective.
1801:
1802: Recall that there is a natural bijection between rank $n$ locally free
1803: sheaves on a variety and rank $n$ vector bundles over it.
1804: For example, the sheaf of differential forms on a variety $X$,
1805: $\Om_X$, corresponds to the cotangent bundle $T^*X$,
1806: and can be viewed as the sections of this bundle. Given a curve $C$,
1807: the locally free rank 2 sheaf of $\tau$-differentials on $C$, as defined in \cite{Ros},
1808: and also denoted $\Omt_C$,
1809: corresponds exactly to the set of rational $\tau$-forms, as defined above,
1810: with a $\tau$-form $\omt$ corresponding to a section $s \in \Omt(U)$,
1811: for open $U \sub C$.
1812:
1813: Given a rank 2 bundle, one can consider all the rank 1 subbundles.
1814: (To be precise, a rank 1 subbundle is a rank 1 bundle together
1815: with a bundle embedding.) Passing to sheaves, rank
1816: 1 subbundles correspond to invertible (i.e., rank 1) subsheaves.
1817: Note that it is possible for one rank 1 sheaf to be properly embedded in
1818: another (in which case the quotient will be a torsion sheaf), essentially
1819: because this can also happen with free rank 1 modules. Nevertheless,
1820: in our situation, any invertible subsheaf of a rank 2 locally free sheaf
1821: embeds in a unique maximal invertible subsheaf, as described below.
1822:
1823: We have seen that \taufs\ correspond to sections of a rank 2 locally
1824: free sheaf. Since the $\seq$-equivalence class of such a form $\omt$
1825: consists of those forms $g\omt, g \in K(C)$, such a class should correspond
1826: to a (unique) maximal subsheaf containing $\omt$ as a section on an open set.
1827: The following easy lemma makes this precise.
1828:
1829: \bla
1830: Let $C$ be a smooth curve, $\MH$ a rank 2 locally free sheaf on $C$,
1831: and $s \in \MH(U)$ a section. Then there is a unique maximal
1832: invertible subsheaf $\MF \sub \MH$ such that $s \in \MF(U)$.
1833: Further, the quotient $\MH / \MF$ is an invertible sheaf.
1834: \ela
1835:
1836: \prf
1837: By possibly passing to a smaller open set, we may assume that $U = \Spec\ R$
1838: is affine, $\MH(U) \cong \tilde{M}$, $M = \<m,n \>$ a free rank 2
1839: $R$-module, and $s = am + bn\in M, a,b \in R$. Since $R$ is a UFD
1840: (e.g.\ \cite{Mat}, Thm.\ 20.1), $a$ and $b$ have a greatest common divisor
1841: $c$. Letting $a' = a/c$, $b'=b/c$, and $s'=a'm + b'n$, the module
1842: $N = Rs' \sub M$ is the unique largest rank 1 free submodule of $M$
1843: containing $s$. Let $d,e \in R$ be such that $da'+eb'=1$, and let
1844: $L = R(em-dn)$. Then it is easy to check that $M = N \op L$, and hence that
1845: there is a short exact sequence $0 \lra N \lra M \lra L \lra 0$.
1846:
1847: We want to define $\MF$ such that $\MF(U) = \tilde{N}$. Since a sheaf is
1848: determined by its stalk at each point $P$, it suffices to show how, for each
1849: $P \in C - U$, $\MF(U)$ determines a unique maximal stalk $\MF_P$
1850: that will make $\MF$ locally free.
1851:
1852: For each such $P$, choose an affine scheme $\ W \sub U \cup \{P\}$
1853: containing $P$, $W \cong \Spec\ S$, such that $\MH(W)$ is free on $W$.
1854: Note that $ W':= W - P =
1855: \Spec\ S_f$, for some $f \in S$, and $P = V(f)$. Let $\MH(W') = \tilde{J}$
1856: and choose basis elements $\<i,j\> \in J$ such that $\MF(W') = \tilde{I}$,
1857: where $I := Ri \sub J$. We then define $\MF_P$ to be the submodule of
1858: $\tilde{J}_P$ generated by $\<i\>$.
1859: Clearly, this determines an extension of $\MF$ to $U \cup P$ and thus,
1860: continuing in this way to all of $C$. (As an aside, note that one could
1861: also define $\MF_P$ to be generated by $f^mi$, for any positive $m$.
1862: Choosing $m > 1$ would yield a non-maximal invertible subsheaf.)
1863: \qed
1864:
1865: \msk
1866:
1867: It is easy to see that two sections determine the same invertible subsheaf
1868: if and only if, on some non-empty open set on which they are both defined,
1869: each is a multiple of the other. In particular, $\sim$ equivalence classes
1870: of \taufs\ correspond to maximal invertible subsheaves, as desired.
1871:
1872: \bdf
1873: Let $C$ be a curve. A {\em $\tau$-invertible sheaf} on $C$, denoted $\MFT$,
1874: is a maximal invertible subsheaf of $\Omt_C$, i.e, an invertible sheaf $\MF$,
1875: together with an embedding $f:\MF\ra\Omt_C$.
1876: \edf
1877:
1878: \brm
1879: For any curve $C$, there is a {\em trivial} \tis, corresponding to
1880: the natural embedding of $\O_C$ into $\Omt_C$. Clearly, this \tis\
1881: corresponds to the $\sim$-equivalence class of trivial \taufs.
1882: Below, by \tis\ we will always mean non-trivial \tis.
1883: \erm
1884:
1885:
1886: We can reformulate Corollary~\ref{formz} as follows.
1887:
1888: \bpro
1889: Let $C$ be a curve. There is a natural bijection between strongly
1890: minimal sets living on $C$ and $\tau$-invertible sheaves on $C$.
1891: \epro
1892:
1893: As before, given a \tis\ $\MFT$ on $C$, let $\Xi(C,\MFT)$ denote the
1894: corresponding strongly minimal\ set.
1895:
1896: %The following lemma is clear.
1897: %\bla
1898: %Let $C,C'$ be curves, $f:C\ra C'$ a morphism, and $\MFT$ a $\tau$-invertible
1899: %sheaf on $C'$. Then the pullback $f^*\MFT$ is a $\tau$-invertible sheaf
1900: %on $C$.
1901: %\ela
1902:
1903: We now introduce the analog of a global \tauf.
1904: Since we are dealing with equivalence classes of \taufs,
1905: the correct notion is that of being a \tauf\ that is $\seq$-equivalent
1906: to a global \tauf. Passing to $\tau$-invertible sheaves, these are
1907: the sheaves $\MFT$ that contain a global section.
1908:
1909: \bdf
1910: Let $C$ be a curve, $\MFT$ a \tis. Say that $\MFT$ is a {\em global \tis}
1911: if $H^0(C,\MFT) \neq 0$, that is, if $\MFT$ has a global section.
1912: \edf
1913:
1914: Next, we define the appropriate notion of a $\tau$-invertible
1915: sheaf being essential, which is more subtle than for 1-forms.
1916: First, we establish the following results.
1917:
1918: \bpro
1919: \label{injpull}
1920: Let $f: X \ra Y$ be a finite morphism of curves.
1921: There is an exact sequence of sheaves on $X$,
1922: $$
1923: 0 \lra f^*\Omt_Y \sr{\al}{\lra} \Omt_X \lra \Om_{X/Y} \lra 0
1924: $$
1925: \epro
1926:
1927: \prf
1928: To check that $\alpha$ is injective, it suffice to check this at the
1929: stalk of the generic point. Here, injectivity is obvious.
1930: \qed
1931:
1932: The following proposition follows immediately.
1933:
1934: \bpro
1935: \label{tauinvpull}
1936: Let $f:X \ra Y$ be a finite morphism of curves, and $\MFT$ a (non-trivial)
1937: \tis\ on $Y$.
1938: Then there is a unique \tis\ $\MGT$ on $Y$ such that $\al(f^*\MFT) \sub \MGT$.
1939: %the map $f^*\Omt_Y \lra \Omt_X$
1940: %determines an injective homomorphism of $f^*\MFT$ into $\MGT$.
1941: Further, there is the following commutative diagram, with exact rows,
1942: $$
1943: \xymatrix{
1944: 0 \ar[r] & f^*\MFT \ar[r]\ar[d] & \MGT \ar[r]\ar[d] & \Om_{X/Y} \ar[r]\ar[d] & 0 \\
1945: 0 \ar[r] & f^*\Omt_Y \ar[r]^\al & \Omt_X \ar[r] & \Om_{X/Y} \ar[r] & 0,\\
1946: }
1947: $$
1948: such that each vertical arrow is an injection, and $\MH$ is a torsion sheaf.
1949: \epro
1950:
1951: \bpr
1952: The only point that needs to be checked is that the cokernel of the map
1953: $f^*\MFT \lra \MGT$ is $\Om_{X/Y}$. To verify this, it suffices to compute
1954: the map locally around each ramification point of $X$.
1955: \epr
1956:
1957: \brm
1958: This proposition says that, given a morphism of curves,
1959: $f:X\ra Y$, the pullback of a \tis\ is not necessarily
1960: a \tis, even though the pullback of a $\tau$-form is a $\tau$-form.
1961: By analogy, the pullback of a 1-form on $Y$ is a 1-form on $X$
1962: though, in general, $f^*\Om_Y \neq \Om_X$. But there is an
1963: injective homomorphism $f^*\Om_Y \ra \Om_X$, just as there is
1964: an injective homomorphism $f^*\MFT \ra \MGT$.
1965:
1966: Note that given a morphism between curves (or more generally varieties),
1967: $f: C_1 \lra C_2$, the pullback of the trivial \tis\ on $C_2$ is equal to
1968: the trivial \tis\ on $C_1$, precisely because $f^*\O_{C_2} = \O_{C_1}$.
1969: \erm
1970:
1971: \bdf
1972: In the notation of the previous proposition, we say that $\MGT$ is the
1973: {\em weak pullback} of $\MFT$ (along $f$).
1974: In case the natural map $f^*\MFT \ra \MGT$ is an isomorphism, we say
1975: that $\MGT$ is the {\em strong pullback} of $\MFT$.
1976: \edf
1977:
1978: The following lemma follows immediately from the definitions.
1979:
1980: \bla
1981: Let $f: X \ra Y$ be a morphism of curves, $\omt$ a $\tau$-form on $Y$,
1982: and $\MFT$ the \tis\ corresponding to $\omt / \sim$. Then $\MGT$, the weak
1983: pullback of $\MFT$, is the \tis\ corresponding to $f^*\omt$.
1984: \ela
1985:
1986: We can now reformulate Lemma~\ref{taupullback} in terms of \tiss.
1987: \bla
1988: Let $f: X \ra Y$ be a morphism of curves, $\MFT$ a \tis\ on $Y$,
1989: and $\MGT$ the weak pullback of $\MFT$ on $X$. Then
1990: $f^{-1}\Xi(Y,\MFT)=\Xi(X,\MGT)$.
1991: \ela
1992:
1993: \bdf
1994: Let $X$ be a curve. A \tis\ $\MGT$ on $X$ is {\em essential}
1995: if there does not exist a curve $Y$ and a morphism $f:X\ra Y$
1996: such that $\MGT$ is the weak pullback of some \tis\ on $Y$.
1997: \edf
1998:
1999: \brm
2000: Let $X$ be a curve, and $f:X \ra Y$ a regular map to another curve $Y$.
2001: By Proposition~\ref{tauinvpull}, there is a natural embedding of $\O_X$-sheaves,
2002: $\alpha: f^*\Omt_Y \lra \Omt_X$. Notice that for any \tis\
2003: $\MGT\sub \Omt_X$, there is maximal invertible subsheaf of
2004: $\alpha(f^*\Omt_Y)$ that embeds in $\MGT$,
2005: namely $\MGT \cap \alpha(f^*\Omt_Y)$.
2006: Let us call this subsheaf $\MH$. If $\MGT$ is essential, then
2007: $\MH$ cannot be of the form $f^*\MFT$, for $\MFT$ a \tis\ on $Y$.
2008:
2009: In particular, although every section of $f^*\Omt_Y$ is naturally a
2010: function on $\tau X$, not every such function is the pullback along
2011: $f$ of a $\tau$-form on $Y$. Functions arising in this way
2012: naturally correspond instead to sections of $f^{-1}\Omt_Y$, but this
2013: is not an $\O_X$-sheaf. Recall that by definition,
2014: $f^*\Omt_Y = f^{-1}\Omt_Y \ot_{f^{-1}\O_Y}\O_X$.
2015:
2016: As sections of a \tis\ $\MFT$ on $Y$ correspond to a $\sim$-equivalence class of functions
2017: on $\tau Y$, sections of $f^{-1}\MFT$ correspond to the pullback of these
2018: functions to $\tau X$ along the lifting map $f\prl$, as defined in Section~\ref{prolong}.
2019: As $f^*\MFT = f^{-1}\MFT\ot_{f^{-1}\O_Y}\O_X$, sections of $f^*\MFT$ correspond to
2020: products of functions from $f^{-1}\MFT$ with rational functions on $X$.
2021: \erm
2022:
2023: \brm
2024: Observe that a $\tau$-form $\omt$ is $\sim$-essential, as defined in
2025: Section~\ref{algcurves},
2026: if and only if the \tis\ corresponding to $\omt / \sim$ is essential.
2027: \erm
2028:
2029: We can now restate Theorem~\ref{genprop}.
2030:
2031: \bthm
2032: Let $C$ be a smooth projective curve of genus $g \geq 1$. Let $\MFT$ be
2033: an essential \tis. Then there is a finite set $A \sub \Xi(C,\MFT)$
2034: such that $\Xi(C,\MFT) - A$ is strictly minimal. Further, for any
2035: definable equivalence relation with finite classes on $\Xi(C,\MFT)$,
2036: over any set of parameters, almost every class has one element.
2037: Two such sets, $\Xi(C_1,\mathcal{F}^\tau_1)$ and
2038: $\Xi(C_2,\mathcal{F}^\tau_2)$ are
2039: non-orthogonal if and only if there is an isomorphism $f:C_1 \ra C_2$ such that
2040: $\mathcal{F}^\tau_1$ is the weak pullback of $f^{*}\mathcal{F}^\tau_2$.
2041: \ethm
2042:
2043: We now complete the proof of Proposition~\ref{globdim}. Given a (locally free)
2044: sheaf $\MF$ on a projective variety $V$, let $\hh^0(V,\MF)= \dim_K H^0(V,\MF)$,
2045: the dimension of the space of global sections of $\MF$.
2046:
2047: Given a curve $C$, we will call $\hh^0(C,\Omt_C)$ the {\em $\tau$-rank} of $C$,
2048: denoted $g^\tau_C$ or simply $g^\tau$. The following proposition shows that $g^\tau$
2049: can be easily determined by the Kodaira-Spencer rank of $C$, which we now explain.
2050:
2051: By \cite{Ros}, we have the following short exact sequence
2052: $$
2053: \xi:\qquad \ 0 \lra \O_C \lra \Omt_C \lra \Om_{C/K} \lra 0.
2054: $$
2055: Recall that $\xi$ is called an extension of $\Om_{C/K}$ by $\O_C$,
2056: and that there is a natural bijection between the set of such extensions and
2057: the cohomology class $\Ext^1(\Om_{C/K}, \O_C)$, such that
2058: the split extension corresponds to $0\in \Ext^1(\Om_{C/K}, \O_C)$
2059: (e.g., \cite{CW}, Theorem 3.4.3). In our case, this mapping is given as follows.
2060: Given $f:C \lra \Spec\, K$, consider the sequence
2061: $$
2062: \eta: \qquad 0 \lra f^*\Om_K \lra \Om_C \lra \Om_{C/K} \lra 0
2063: $$
2064: (with $\Om_C$ projective). Applying $\Hom(-,\O_C)$ yields the exact sequence
2065: $$
2066: \Hom(\Om_C,\O_C) \lra \Hom(f^*\Om_K, \O_C) \sr{\partial}{\lra}
2067: \Ext^1(\Om_{C/K},\O_C) \lra 0
2068: $$
2069: Given $x\in \Ext^1(\Om_{C/K},\O_C)$, choose $\beta \in \Hom(f^*\Om_K, \O_C)$
2070: such that $\pd(\beta)=x$. We then let $\zeta$, the pushout of $\eta$ along $\beta$,
2071: be the extension corresponding to $x$.
2072: $$
2073: \xymatrix{
2074: \eta: & 0 \ar[r] & f^*\Om_K \ar[r]\ar[d]^\beta
2075: & \Om_C \ar[r]\ar[d] & \Om_{C/K} \ar[r]\ar[d] & 0 \\
2076: %
2077: \zeta: & 0 \ar[r] & \O_C \ar[r] & \MG \ar[r] & \Om_{C/K} \ar[r] & 0
2078: }
2079: $$
2080: Thus, by \cite{Ros}, the extension $\xi$ corresponds to the element
2081: $\pd(\dltl) \in \Ext^1(\Om_{C/K},\O_C)$, where $\dltl$ is determined by the derivation
2082: on $K$.
2083:
2084: There are natural isomorphisms $\Ext^1(\Om_{C/K},\O_C) \cong
2085: \Ext^1(\O_C, \Theta_{C/K}) \cong H^1(C,\Theta_{C/K})$,
2086: with $\Theta_{C/K} = \Om_{C/K}^\vee$, the tangent sheaf
2087: of $C$ (the dual of $\Om_{C/K}$) (e.g., \cite{RH}, pages 234-5).
2088: Thus, there is a natural map, which we also denote $\pd$,
2089: $\pd:\Hom(f^*\Om_K,C) \lra H^1(C,\Theta_{C/K})$,
2090: which Buium~(\cite{Bui93}, p. 1395) calls the Kodaira-Spencer map of $C$.
2091: The element $\pd(\dltl) \in \Ext^1(\Om_{C/K},\O_C)$, or $H^1(C,\Theta_{C/K})$,
2092: is the Kodaira-Spencer class of $C$.
2093:
2094: From the sequence $\eta$, corresponding to the Kodaira-Spencer class of $C$,
2095: one obtains a long exact sequence of cohomology groups.
2096: $$
2097: 0 \lra \HH^0(C,\O_C) \lra \HH^0(C,\Omt_C) \lra \HH^0(C,\Om_{C/K})
2098: \stackrel{\partial}{\lra}
2099: \HH^1(C,\O_C) \lra \ldots
2100: $$
2101: Let us call $\dim_K \partial (\HH^0(C,\Om_{C/K}))$ the {\em Kodaira-Spencer} rank of
2102: $C$, denoted $KS_C$. Clearly, if the Kodaira-Spencer class of $C$ is trivial,
2103: that is, if $\eta$ splits, then $KS_C = 0$, though the converse does not hold in
2104: general.
2105: For curves of genus 1, though, $\eta$ splits if and only if $KS_C = 0$.
2106: This follows, for example,
2107: from Atiyah's classification of rank 2 bundles over elliptic curves \cite{At}.
2108: This yields a complete understanding of global $\tau$-forms on elliptic curves.
2109:
2110: The following proposition is equivalent to Proposition~\ref{globdim} above.
2111:
2112: \bpro
2113: \label{cohom}
2114: Let $C$ be a smooth projective curve.
2115: Then $1 \leq h^0(C,\Omt_C)\leq g + 1$.
2116: Precisely, $h^0(C,\Omt_C) = g+1 - KS_C$.
2117: In particular, if $C$ is defined over $k$, the field of constants, then
2118: $h^0(C,\Omt_C) = g+1 $.
2119: \epro
2120:
2121: \prf
2122: The first two statements follow from remarks above. The third follows from
2123: the fact that a curve defined over $k$ has trivial Kodaira-Spencer class.
2124: (Since $H^1(C,\Theta_{C/K})$ classifies $TC$-torsors (e.g.,
2125: see~\cite{JM}), this is equivalent to the fact that for such curves,
2126: $\tau C \cong_C TC$.)
2127: \qed
2128:
2129:
2130:
2131: \ignore{
2132: To prove the third,
2133: recall that the class $H^1(C,\Theta_{C/K})$ classifies $TC$-torsors (e.g.,
2134: see~\cite{JM}), that and Buium~(\cite{Bui93}, p. 1396) proves that the
2135: {\em Kodaira-Spencer class}\
2136: $\pd(\dltl)$ corresponds with the $TC$-torsor $\tau C$. The torsor is
2137: trivial, that is, isomorphic to $TC$, if and only if $\pd(\dltl) =0$, which
2138: holds exactly when the original sequence $\xi$ splits. Putting everything together,
2139: one gets that $\xi$ splits if and only if $\tau C \cong_C TC$, which happens exactly
2140: when $C$ is defined over $k$, as desired. This completes the proof.
2141: \qed
2142: }
2143: %%end{ignore}
2144:
2145: \subsection{Definability results}
2146:
2147: Recall that the {\em degree} of an invertible sheaf on a variety is defined to
2148: be the degree of the corresponding divisor. We now observe that there is
2149: a (uniform) bound on the degree of a $\tau$-invertible sheaf on a curve.
2150:
2151: \bla
2152: Let $V$ be a projective curve, $\MF$ a coherent sheaf on $V$. Then for any
2153: invertible subsheaf $\MG$ of $V$, $\deg \MG \leq \hh^0(\MF) + g(C) - 1$.
2154: \ela
2155:
2156: \bpr
2157: By the Riemann-Roch theorem,
2158: $\deg \MG = \hh^0(\MG) - \hh^0(\om \otimes \MG^{-1}) + g - 1 \leq
2159: \hh^0(\MG) + g - 1 $. Since $\hh^0(\MG) \leq \hh^0(\MF)$, this yields the
2160: desired bound.
2161: \epr
2162:
2163: \bco
2164: For any curve $C$, the degree of any \tis\ is $\leq 2g$.
2165: \eco
2166:
2167: The next result then follows by standard arguments.
2168:
2169: \bco
2170: For any curve $C$, the set of global $\tau$-forms on $C$ is definable,
2171: as is the equivalence relation $\sim$ on these forms. Further, this remains
2172: true for families of curves (of fixed genus). Finally, the dimension of
2173: the space of global $\tau$-forms, as a $K$-vector space, is uniformly definable.
2174: \eco
2175:
2176: Observe that, by Proposition~\ref{cohom}, this last claim is equivalent to the known fact
2177: that the Kodaira-Spencer rank of a curve is definable.
2178:
2179: %% ??
2180:
2181: \ignore{
2182: \newpage
2183:
2184: We can now state one of our main results.
2185: All curves are smooth projective curves over an uncountable algebraically
2186: closed field $K$. Given a curve $C$ and a sheaf $\MF$, recall that
2187: $h^i(C,\MF) = \dim_K(H^i(C,\MF))$. (I'll specify the notion of
2188: generic curve later.)
2189:
2190: \bthm
2191: Let $C$ be a generic curve of genus $g \geq 2$. Then every 'non-trivial'
2192: global $\tau$-invertible sheaf is essential.
2193: \ethm
2194:
2195: The proof uses the following proposition, which was provided by L. Ein.
2196:
2197: \bpro
2198: Let $C$ be a generic curve of genus $g \geq 2$. For any curve $C'$,
2199: morphism $f:C \ra C'$, and invertible sheaf $\MF$ on $C'$,
2200: if $h^0(C',\MF) = 0$, then $h^0(C,f^*\MF) = 0$.
2201: \epro
2202:
2203: \prf
2204: The following fact is well-known. (The proof involves a counting
2205: argument. The moduli space of curves of genus $g \geq 2$ is
2206: $(3g-3)$-dimensional and, by the Hurwitz formula, if $f: C_1 \ra C_2$
2207: is a morphism between curves, then $g(C_2) \leq g(C_1)$.)
2208:
2209: \bla
2210: Let $C$ be a generic curve of genus $g \geq 2$, $C$ any curve, and
2211: $f:C\ra C'$ a morphism of degree $\geq 2$. Then $C' \cong \P^1$.
2212: \ela
2213:
2214: We now assume that $C'\cong\P^1$, and let $\MF$ be an invertible sheaf
2215: on $\P^1$, $\MF=\O_{\P^1}(n), n\in\Z$. Then
2216: $
2217: h^0(C,f^*\MF)=h^0(\P^1,f_*f^*\MF)=h^0(C',\MF\otimes f_*\O_C).
2218: $
2219: The first equality is obvious, and the second follows immediately from
2220: the Projection Formula (\cite{RH}, p.\ 124).
2221:
2222: We claim that $f_*\O_C$ is a locally free sheaf of the form
2223: $\O_{\P^1}\oplus(\oplus_i\O_{\P^1}(a_i)),$
2224: with $a_i < 0$, for all $i$. We have a short exact sequence,
2225: $$
2226: 0\lra \O_{\P^1}\stackrel{\alpha}\lra f_*\O_C\lra \ME \lra 0.
2227: $$
2228: Let $\beta: f_*\O_C \ra \O_{\P^1}$ be the map $\beta = \frac{1}{d}t$,
2229: where $t$ is the standard trace map. Then $\beta\circ\alpha$ is the
2230: identity on $\O_{\P^1}$, so the above sequence splits,
2231: i.e., $f_*\O_C=\O_{\P^1}\oplus\ME$.
2232:
2233: Next, observe that $\ME$ is locally free. To show this, it is sufficient
2234: to show that $\ME_x$ is locally free, for all $x\in \PO$
2235: (\cite{RH}, p.\ 124). For the generic point $x_0\in \PO$,
2236: we get an exact sequence of vector spaces over the field $K(y)$.
2237: $$
2238: 0\lra \O_{\P^1,x_0}\lra f_*\O_{C,x_0}\lra \ME_{x_0} \lra 0
2239: $$
2240: as desired. For non-generic $x\in\PO$, we get a corresponding short
2241: exact sequence, with $f_*\O_{C,x}$ a finitely generated module over
2242: $\O_{\PO,x}$, which is a discrete valuation ring, and hence a
2243: principal ideal domain. Clearly, $f_*\O_{C,x}$ is torsion free, and hence free
2244: ([L], p.\ 147). Likewise, one can easily show that $\ME_x$ is torsion free,
2245: and hence free.
2246:
2247: Every locally free sheaf on $\P^1$ is of the form $\oplus_i\O_{P_1}(a_i)$,
2248: $a_i\in\Z$ (\cite{RH}, p.\ 384).
2249: Observe that $\ME$ has no global sections, so it must be of
2250: the form $\oplus_i\O_{P_1}(a_i), a_i < 0$. Indeed, one has the following exact
2251: sequence.
2252: $$
2253: 0\lra H^0(\PO,\O_\PO)\lra H^0(\PO,f_*\O_C)\lra H^0(\PO,\ME)
2254: \lra H^1(\PO,\O_\PO)
2255: $$
2256: Since $h^0(\PO,f_*\O_C)=1$ and $h^1(\PO,\O_\PO)=\textup{genus}(\PO)=0$,
2257: we get $h^0(\PO,\ME)=0$, as desired. This completes the proof of the claim.
2258:
2259: Finally, let $\MF$ be an invertible sheaf on $\PO$ with $h^0(\PO,\MF)=0$,
2260: so $\MF=\O_\PO(b), b < 0$. By above, to show that $h^0(C,f^*\MF)=0$,
2261: it suffices to show $h^0(C,\MF\otimes f_*\O_C)=0$. But
2262: $$
2263: \MF\otimes f_*\O_C=\O_\PO(b)\otimes(\O_{\P^1}\oplus(\oplus_i\O_{P_1}(a_i)))=
2264: \O_\PO(b) \oplus (\oplus_i\O_{P_1}(a_i+b))
2265: $$
2266: which obviously has no global sections,
2267: completing the proof of the proposition.
2268: \qed
2269:
2270: We introduce a second, similar equivalence relation on \taufs\
2271: that looks somewhat less natural but will be more useful for
2272: later applications.
2273:
2274: \bdf
2275: Let $C$ be a curve, $\omt \in \Omt(C)$. The {\em unary set} of $\omt$ is
2276: $$
2277: M_{\omt} = \{(a, u): a \in C-Z_{\omt}, u\in \tau C_a \text{ and }\omt_a(u) = 1\}.
2278: $$
2279: This is a rational section of the algebraic variety $\tau C$,
2280: and thus birational to $C$.
2281:
2282:
2283: Say that $\omt_1, \omt_2 \in \Omt (C)$ are {\em $\seq$-equivalent},
2284: written $\omt_1\teq\omt_2$, if
2285: $M_{\omt_1}$ and $M_{\omt_2}$ are almost equal.
2286: \edf
2287:
2288: \bla
2289: \label{teqcl}
2290: Let $C$ be a curve, $\omt_1, \omt_2 \in \Omt(C)$, both non-trivial.
2291: Then $\omt_1$ and $\omt_2$ are $\seq$-equivalent if and only if there is a rational
2292: function $f\in K(C)$ such that $\omt_1=f(\omt_2-1) + 1$.
2293: \ela
2294:
2295: \prf
2296: If $\omt_1=f(\omt_2-1) + 1$, then it is clear $M_{\omt_1}$ and
2297: $M_{\omt_2}$ are almost equal. The other direction follows as
2298: with $\seq$-equivalence.
2299: \qed
2300:
2301: \bla
2302: \label{taunull}
2303: Let $C$ be a curve, $\omt_1, \omt_2 \in \Omt(C)$, both non-trivial.
2304: Suppose that $M_{\omt_1} \cap M_{\omt_2}$ is infinite. Then
2305: $\omt_1$ and $\omt_2$ are $\seq$-equivalent. In other words,
2306: if $M_{\omt_1} \cap M_{\omt_2}$ is infinite, then
2307: $M_{\omt_1}$ and $M_{\omt_2}$ are almost equal.
2308: \ela
2309:
2310: \prf
2311: Like the proof of Lemma~\ref{sigmanull}.
2312: \qed
2313:
2314: %\bqu
2315: %Are the global $\tau$-forms on any abelian variety exactly the invariant
2316: %$\tau$-forms,
2317: %since any abelian variety is a quotient of a Jacobian variety [I think]?
2318: %\equ
2319:
2320: To prove the second statement, we first determine $\ker(\dltl)$ and $\ker(\be)$.
2321: Setting $R = K$ in the above diagram, one obtains
2322: $$
2323: \xymatrix{
2324: 0 \sr[r] &\Om_K \ar[r]^\alpha \ar[d]^{\tilde{\dl}} & \Om_K \ar[r]
2325: \ar[d]^\beta & \Om_{K/K}\ar[r]\ar[d]^{=} & 0 \\
2326: 0 \ar[r] &K \ar[r]^{\iota} & \Omt_K \ar[r]^{\lm} & \Om_{K/K} \ar[r] & 0
2327: }
2328: $$
2329: with $\Om_{K/K}=0$, $\dltl$ and $\beta$ surjective.
2330: By definition, $\ker(\be)=\< \dl(a)db-\dl(b)da | a,b\in K\>\sub \Om_K$,
2331: which we call $M$. By the Snake Lemma, $\ker(\dltl)= M$ also,
2332: so one gets a short exact sequence,
2333: $0\lra M \lra \Om_K \lra K \lra 0$.
2334:
2335: Now let $R$ be any $K$-algebra. Tensoring the previous sequence with $R$,
2336: one gets, $0\lra R\ot_K M \lra R \ot_K\Om_K \lra R \lra 0$. By the definition
2337: of $\Omt_R$, one also has $0 \lra R\ot_KM\lra\Om_R\lra\Omt_R\lra 0$.
2338: By Lemma~\ref{tdfeq}, these sequences fit together to form the following
2339: diagram.
2340: $$
2341: \xymatrix{
2342: 0 \ar[r] & R \ot_K M \ar[r] \ar[d]^= & R\ot_K \Om_K \ar[r]^\dltl \ar[d]^\al
2343: & R \ar[r] \ar[d]^\iota & 0 \\
2344: 0\ar[r] & R \ot_K M \ar[r] & \Om_R \ar[r]^\be & \Omt_R \ar[r] & 0
2345: }
2346: $$
2347: By the Snake Lemma again, $\Ker(\al) \cong \Ker (\iota)$, which completes the
2348: proof.
2349:
2350: There is also hidden my original construction of $\tau$-differentials.
2351:
2352:
2353: \brm
2354: There is a more algebraic way to construct the module of 1-forms
2355: on a variety, via the notion of K\"{a}hler differentials.
2356: Let $R = K[V]$ be a $K$-algebra, e.g.\ the ring of regular functions on an
2357: affine $K$-variety.
2358: Then one defines the $R$-module $\Omega_{R/K}$ as a certain universal object.
2359: For more information, see Eisenbud~(\cite{Eis}, Chapter 16)
2360: or Hartshorne~(\cite{RH}, Chapter~II.8).
2361: %
2362: We introduce K\"{a}hler $\tau$-differentials, which should
2363: provide a purely algebraic construction of \taufs. I will develop
2364: %\marginpar{Expand.}
2365: and investigate these ideas later.
2366: \erm
2367: %
2368: Let $A$ be a ring, $B$ an $A$-algebra, and $M$ a $B$-module. Recall that a
2369: derivation is an abelian group homomorphism $d:B\lra M$ satisfying the Leibniz
2370: rule, $d(bc)=bdc+cdb$. The derivation is $A$-linear if it is a homomorphism of
2371: $A$-modules.
2372: %
2373: \bdf
2374: Let $(A,\dl)$ be a differential ring and $B$ an $A$-algebra.
2375: \ben
2376: \item
2377: A $B$-$\tau$-module is a pair $M^\tau=(M,\iota_M)$ such that $M$ is a
2378: $B$-module and $\iota_M$ is an injective $B$-module map from $B$ itself
2379: to $M$, $\iota_M : B\lra M$. (Thus, a $B$-$\tau$-module is a
2380: $B$-module together with a distinguished free rank 1 submodule of $M$.)
2381: %
2382: Given two $B$-$\tau$-modules $M_0^\tau$ and $M_1^\tau$, a
2383: homomorphism $f$
2384: from $M_0^\tau$ to $M_1^\tau$ is a $B$-module homomomorphism from $M_0$
2385: to $M_1$ such that for all $b\in B$, $f\circ \iota_{M_0} (b)=\iota_{M_1}(b)$.
2386: %
2387: \item
2388: A derivation $\tau$ from $B$ to a $B$-$\tau$-module $M^\tau$ is
2389: $A$-quasilinear if for all $a\in A$, $\tau a = \iota (\dl (a))$.
2390: %
2391: \een
2392: \edf
2393: %
2394: The following lemma explains the choice of the term `quasilinear'.
2395: (The formula for $\tau(ab)$ bears a certain resemblance to Pillay \&
2396: Ziegler's definition of a $\partial$-module in their paper,
2397: beginning of Section 3. They have an additive endomorphism $D:V\lra V$
2398: such that for $c\in K$, $v \in V$, $D(cv)= cDv+\dl(c)v$.)
2399: %
2400: \bla Let $(A,\dl)$ be a differential ring, $B$ an $A$-algebra, $M^\tau$ a
2401: $B$-$\tau$-module, and $\tau :B\lra M^\tau$ an $A$-quasilinear derivation.
2402: Then for $a\in A$, $b\in B$, $\tau(ab)=a\tau b + \dl (a) \iota b$.
2403: %
2404: In particular, if $\dl$ is the trivial derivation on $A$, then
2405: $\tau$ is an $A$-linear derivation.
2406: \ela
2407: %
2408: \prf
2409: Let $a\in A, b \in B$. Then $\tau(ab)=$
2410: $$
2411: a\tau(b) + b\tau(a)=
2412: a\tau b + b\iota(\dl(a))=
2413: a\tau b + \iota(b\dl(a))=
2414: a\tau b + \dl(a)\iota(b).
2415: $$
2416: %
2417: The second statement follows immediately.
2418: %
2419: \qed
2420: %
2421: \bdf
2422: The module of K\"{a}hler $\tau$-differentials of $B$ over $A$,
2423: written $\Omt_{A/B}$, is the $B$-$\tau$-module generated by
2424: $\{\tau b : b \in B\} \cup \{\iota b : b \in B\}$, with the following relations
2425: holding for all $a\in A$, $b,b' \in B$.
2426: $$
2427: \tau(bb')=b\tau b' + b'\tau b
2428: $$
2429: %
2430: $$
2431: \tau(a)=\iota(\dl(a))
2432: $$
2433: The map $\tau : B \lra\Omt_{A/B}$, sending $b$ to $\tau b$ is an
2434: $A$-quasilinear derivation, called the universal
2435: $A$-quasilinear derivation.
2436: \edf
2437: %
2438: The next lemma follows immediately from the previous definition.
2439: %
2440: \bla
2441: $\Omt_{A/B}$ has the expected universal property. That is,
2442: given a $B$-$\tau$-module $M^\tau$ and a $\tau$-derivation
2443: $\sigma: B \lra M$, there is a unique homomorphism of $B$-$\tau$-modules
2444: $\phi: \Omt_{A/B}\lra M$ such that $\sigma=\phi\circ\tau$.
2445: %
2446: Equivalently, there is a natural bijection,
2447: $$
2448: \text{Der}^\tau_B(B,M)\cong\text{Hom}^\tau(\Omt_{B/A},M).
2449: $$
2450: \ela
2451: %
2452: The significance of the following lemma will become clearer after
2453: the introduction below of the $\Lm$-map from \taufs\ to 1-forms
2454: in Lemma~\ref{lambda}. The proof is immediate.
2455: %
2456: \bla
2457: Let $(A,\dl)$ be a differential ring, $B$ an $A$-algebra. There
2458: is a canonical homomorphism $\widehat{\Lambda}$ from $\Omt_{B/A}$ to $\Om_{B/A}$,
2459: that sends each $\tau b \in \Omt_{B/A}$ to $db \in \Om_{B/A}$ and has kernel
2460: $\iota (B)$.
2461: %
2462: Thus, there is a $B$-module isomorphism, $\Omt_{B/A}\cong\Om_{B/A}\oplus B.$
2463: \ela
2464: }
2465: %% end ignore
2466:
2467: \bibliography{jsl-smsets.bib}
2468: \bibliographystyle{alpha}
2469:
2470: \end{document}
2471:
2472: