1: \documentclass[11pt]{article}
2: \usepackage{mathpple}
3: \usepackage{amsfonts}
4: \usepackage{amsmath}
5: \usepackage{amstext}
6: \usepackage{amssymb}
7: \usepackage{amsthm}
8: \usepackage{amscd}
9: \usepackage{xypic}
10:
11: \usepackage{vmargin}
12: \setlength{\parindent}{0in}
13: \setlength{\parskip}{\medskipamount}
14:
15: \def\catmod{{\sf Mod}}
16: \def\sp{{\mathop{\rm Sp}}}
17: \def\gsp{{\mathop{\rm GSp}}}
18: \def\orth{{\mathop{\rm Orth}}}
19: \def\O{{\mathop{\rm O}}}
20: \def\u{{\mathop{\bf U}}}
21: \def\liegsp{{\mathop{\frak g\frak s\frak p}}}
22: \def\liesp{{\mathop{\frak s\frak p}}}
23: \def\qed{$\square$}
24:
25: \global\let\bar\undefined
26: \newcommand{\bar}[1]{{\overline{#1}}}
27:
28: \global\let\rank\undefined
29: \DeclareMathOperator{\rank}{rank}
30:
31: \def\arith{{\rm arith}}
32: \def\geom{{\rm geom}}
33: \def\cl{{\rm cl}}
34: \def\cgn{{_N\calc_g}}
35: \def\mgn{{_N\calm_g}}
36: %% \def\cgn{\calc_{g,N}}
37: %% \def\mgn{\calm_{g,N}}
38:
39: \def\defeq{:=}
40:
41:
42: \def\ff{{\mathord{\mathbb F}}}
43:
44:
45: \global\let\theorem\undefined
46: \global\let\lemma\undefined
47: \global\let\cor\undefined
48: \global\let\prop\undefined
49:
50: %\renewcommand{\thesection}{
51: %\ifnum \value{subsection}=0{\arabic{section}}\fi
52: %\ifnum \value{subsection}>0{\arabic{section}.\arabic{subsection}}\fi
53: %}
54: \renewcommand{\thesubsection}{\ifnum\value{subsection}=0{\arabic{section}}\fi\ifnum\value{subsection}>0{\arabic{section}.\arabic{subsection}}\fi}
55:
56: \numberwithin{equation}{subsection}
57: \numberwithin{table}{section}
58:
59: \newtheorem{theorem}[equation]{Theorem}
60: \newtheorem{lemma}[equation]{Lemma}
61: \newtheorem{prop}[equation]{Proposition}
62: \newtheorem{cor}[equation]{Corollary}
63:
64: \theoremstyle{remark}
65: \newtheorem{situation}[equation]{Situation}
66:
67:
68: % miscellaneous macros
69: \def\mypara{\paragraph}
70:
71:
72:
73: \def\ang{\ip}
74: \def\calc{{\cal C}}
75: \def\cald{{\cal D}}
76: \def\cale{{\cal E}}
77: \def\calf{{\cal F}}
78: \def\calh{{\cal H}}
79: \def\calm{{\cal M}}
80: \def\caln{{\cal N}}
81: \def\calo{{\cal O}}
82: \def\calp{{\cal P}}
83: \def\calt{{\cal T}}
84: \def\calu{{\cal U}}
85: \def\cross{\times}
86: \DeclareMathOperator{\aut}{Aut}
87: \DeclareMathOperator{\End}{End}
88: \DeclareMathOperator{\frob}{Fr}
89: \def\ff{{\mathord{\Bbb F}}}
90: \DeclareMathOperator{\gl}{GL}
91: \def\gth{\frak}
92: \DeclareMathOperator{\id}{id}
93: \def\inject{\hookrightarrow}
94: \def\integ{{\mathord{\Bbb Z}}}
95: \def\inv{^{-1}}
96: \def\iso{\cong}
97: \DeclareMathOperator{\jac}{Jac}
98: \DeclareMathOperator{\mat}{Mat}
99: \DeclareMathOperator{\mult}{mult}
100: \def\nat{{\mathord{\Bbb N}}}
101: \def\pf{\mypara{Proof}}
102: \def\ra{\rightarrow}
103: \def\rat{{\mathord{\Bbb Q}}}
104: \DeclareMathOperator{\spec}{Spec}
105: \def\units{^\cross}
106: \newcommand{\abs}[1]{{\left|#1\right|}}
107: \newcommand{\floor}[1]{{\lfloor #1 \rfloor}}
108: \newcommand{\half}[1]{\frac{#1}{2}}
109: \newcommand{\ip}[1]{{{\langle #1 \rangle}}}
110: \newcommand{\lie}[1]{{\gth #1}}
111: \newcommand{\oneover}[1]{\frac{1}{#1}}
112: \newcommand{\rest}[1]{|_{#1}}
113: \newcommand{\st}[1]{\{#1\}}
114: \newcommand{\til}[1]{{\widetilde{#1}}}
115:
116: \title{The distribution of class groups of function fields}
117: \author{Jeffrey D. Achter\\
118: {\tt j.achter@colostate.edu}
119: }
120: \date{\empty}
121: \begin{document}
122: \maketitle
123:
124: \begin{abstract}
125: For any sufficiently general family of curves over a finite field
126: $\ff_q$ and any elementary abelian $\ell$-group $H$ with $\ell$
127: relatively prime to $q$, we give an
128: explicit formula for the proportion of curves $C$ for which
129: $\jac(C)[\ell](\ff_q)\iso H$. In doing so, we prove a
130: conjecture of Friedman and Washington.
131: \end{abstract}
132:
133:
134: %\tableofcontents
135:
136:
137: In 1983, Cohen and Lenstra introduced heuristics \cite{cl} to explain
138: statistical observations about class groups of imaginary quadratic
139: fields. Their principle, although still unproven, remains an
140: important source of guidance in number theory. A concrete application
141: of their heuristics predicts that an abelian group occurs as a class
142: group of an imaginary quadratic field with frequency inversely
143: proportional to the size of its automorphism group.
144:
145: Six years later Friedman and Washington \cite{fw} addressed the
146: function field case. Fix a finite field $\ff_q$ and an abelian
147: $\ell$-group $H$, where $\ell$ is an odd prime relatively prime to
148: $q$. Friedman and Washington conjecture that $H$ occurs as the
149: $\ell$-Sylow part of the divisor class group of function fields over
150: $\ff_q$ with frequency inversely proportional to $\abs{\aut(H)}$. As
151: evidence for this, they prove that the uniform distribution of
152: Frobenius automorphisms of curves of genus $g$ in $\gl_{2g}(\integ_\ell)$ would imply
153: their conjecture. (Of course, autoduality of the Jacobian means that
154: these Frobenius elements are actually in $\gsp_{2g}(\integ_\ell)$; still,
155: \cite{fw} entertains the hope that this distinction is immaterial to
156: the problem.) Friedman and Washington observe that, since geometric
157: equidistribution results seem within reach, their conjecture may well
158: be tractable.
159:
160: Taking advantage of recent progress in equidistribution, we
161: prove a statement in the spirit of \cite{fw}. A special, yet typical,
162: case of our main result says the following.
163:
164: Let $\calc \ra \calm/\ff_q$ be a relative smooth, proper curve of
165: genus $g$ over a smooth, irreducible variety, and suppose that $\calc
166: \ra \calm$ has full $\ell$ monodromy. Let $\st{\ff_{q^{e_n}}}$ be a
167: tower of finite extensions of $\ff_q$ which is
168: cofinal in the collection of all finite extensions of $\ff_q$. For a
169: curve $C$, let $\jac(C)[\ell](k)$ denote the $k$-rational $\ell$-torsion subgroup of its Jacobian.
170: We give an explicit formula for a number $\alpha(g,r)$ so that:
171: %\begin{theorem}
172: \[
173: \lim_{n\ra\infty}\frac{\abs{\st{ x \in \calm(\ff_{q^{e_n}}) :
174: \jac(\calc_x)[\ell](\ff_{q^{e_n}}) \iso
175: (\integ/\ell)^r}}}{\abs{\calm(\ff_{q^{e_n}})}} = \alpha(g,r).
176: \]
177: %\end{theorem}
178:
179: Moreover, $\lim_{g \ra \infty} \alpha(g,r)$ exists. In this way, we
180: can formulate a version of our result which allows the genus of the
181: curves in question to change, too. The term $\alpha(g,r)$ should be
182: thought of as a sort of symplectic analogue of $\abs{\aut((\integ/\ell)^r)}\inv$.
183:
184: This result lets us essentially prove the original conjecture
185: of Friedman and Washington. (We will see in Section \ref{subsecfw} that the heuristic
186: they introduce, that statistics of $\gl_{2g}$ should track those of
187: $\sp_{2g}$, is a reasonable approximation but not literally true.)
188: Moreover, we significantly strengthen (Section \ref{subseche})
189: results of Cardon and Murty \cite{cardonmurty} on divisibility of
190: class groups of quadratic function fields.
191:
192: The family of curves $\calc \ra \calm$ is a concrete device for
193: enumerating function fields. The ``full $\ell$-monodromy''
194: constraint ensures that, as far as $\ell$-Sylow subgroups of class
195: groups are concerned, the family behaves like a general one.
196:
197: On one hand, the familiar moduli spaces $\mgn$ \cite{dm} of
198: proper smooth curves of genus $g$ equipped with principal Jacobi level
199: $N$ structure
200: have
201: this property. In fact, so do most versal families of curves
202: \cite{ekedahl}. In this sense, our main theorem describes a typical
203: collection of function fields of genus $g$.
204:
205: On the other hand, it's not hard to write down families of curves
206: which {\em do not} have this property. Generally speaking, extra
207: algebraic cycles on a family of curves force the $\ell$-adic monodromy
208: to lie in a proper subgroup of $\sp_{2g}$. As a specific caution, we
209: mention that if $d > 2$ then the family of curves $C_{d,f}: y^d =
210: f(x)$ does not have full monodromy. (In fact, for such curves,
211: $\integ[\zeta_d] \inject \End_{\bar\ff_q}\jac(C_{d,f})$. One
212: suspects \cite{zarendjac} that the $\ell$-adic monodromy is a unitary
213: group associated to $\rat(\zeta_d)$; at the very least, the monodromy
214: group is contained in such a group.)
215:
216: The first section of this paper collects results of Katz about
217: equidistribution of Frobenius elements on $\ell$-adic sheaves. The
218: second section investigates the combinatorics of $\sp_{2g}(\ff_\ell)$
219: and related groups. The next section combines these results to make
220: precise statements about the distribution of class groups of function
221: fields. The paper concludes with a series of applications of these
222: results, culminating in a proof of a modified Friedman-Washington
223: conjecture.
224:
225: I thank R. Pries for helpful comments on this paper.
226:
227: \section{$\ell$-adic monodromy}
228: \setcounter{equation}{0}
229:
230:
231: As noted above, Friedman and Washington foresaw that good
232: equidistribution theorems would allow one to prove Cohen-Lenstra type
233: results for function fields. Here, we recall the precise statements
234: we need. Our discussion follows section one of \cite{achhol}, which
235: itself is a recapitulation of parts of chapter nine of
236: \cite{katzsarnak}.
237:
238: Fix an odd prime $\ell$. Let $\calo_\lambda$ be the ring of integers
239: in some finite extension of $\integ_\ell$, and let $\Lambda =
240: \calo_\lambda/\lambda^n$ for some $n$. Let $V = V_\Lambda$ be a free,
241: rank $2g$ $\Lambda$-module equipped with a symplectic form
242: $\ip{\cdot,\cdot}$. The group of symplectic similitudes of
243: $(V,\ip{\cdot,\cdot})$ is
244: \[
245: \gsp(V,\ip{\cdot,\cdot}) = \st{
246: A \in \gl(V) | \exists \mult(A)\in \Lambda\units : \forall v,w\in V,
247: \ip{Av,Aw} = \mult(A)\ip{v,w}} \iso \gsp_{2g}(\Lambda).
248: \]
249:
250: The ``multiplicator'' $\mult$ is a character of
251: $\gsp(V,\ip{\cdot,\cdot})$, and its
252: kernel is the usual symplectic group $\sp_{2g}(\Lambda)$. For
253: $\xi\in \Lambda\units$, let $\gsp_{2g}^\xi(\Lambda) =
254: \mult\inv(\xi)$ be the
255: set of symplectic similitudes with multiplier $\xi$; each
256: $\gsp_{2g}^\xi$ is a torsor over $\sp_{2g}$. For $W\subset
257: \gsp_{2g}(\Lambda)$, let $W^\xi = W\cap \gsp_{2g}^\xi(\Lambda)$.
258:
259: Let $T \ra \spec \integ[1/\ell]$ be a connected normal scheme of
260: finite type, often $\spec\integ[1/\ell]$ itself. Let $\calm \ra T$ be
261: a scheme with smooth geometrically irreducible fibers; let
262: $\eta_\calm$ be a generic point of $\calm$. A local system $\calf$
263: of symplectic $\Lambda$-modules of rank $2g$ is equivalent to a continuous
264: representation $\rho_\calf:\pi_1(\calm,\bar\eta_\calm) \ra
265: \aut(\calf_{\bar\eta_\calm}) \iso \gsp_{2g}(\Lambda)$. We call the image of
266: this representation the (arithmetic) monodromy group of $\calf$.
267:
268: Let $k$ be a finite field and $t\in T(k)$. Then $\calm_t$ is a
269: $k$-scheme, and we distinguish the geometric fundamental group
270: $\pi_1^\geom(\calm_t) = \pi_1(\calm_t \cross \bar k)\subseteq \pi_1(\calm_t)$. The Galois
271: group of $k$ is (canonically isomorphic to) the quotient
272: $\pi_1(\calm_t)/\pi_1^\geom(\calm_t)$.
273:
274: We will require our sheaves to have uniform geometric monodromy group
275: $G_\geom\subseteq \sp_{2g}(\Lambda)$, in the sense that for every
276: finite field $k$ and $t\in T(k)$, the image of
277: $\rho_\calf(\pi_1^\geom(\calm_t))$ is $G_\geom$. With this assumption
278: the monodromy group $G$, the full image of $\rho_\calf$, is contained
279: in $\Lambda\cdot G_\geom \subseteq \gsp_{2g}(\Lambda)$. We let
280: $\xi(k)$ denote the image of $\frob_{k}$, the canonical generator of
281: $\pi_1(\spec k)$, in $G/G_\geom \subseteq \Lambda\units$.
282:
283: If $k$ is a finite field, then to a $k$-point $x\in \calm(k)$
284: one may associate its (conjugacy class of)
285: Frobenius $\frob_{x/k}$ in $\pi_1(\calm)$. Via $\rho_\calf$, the
286: Frobenius at $x$ acts on $\calf_{\bar\eta_\calm}$. Katz shows that these
287: Frobenius elements are equidistributed in the monodromy group.
288:
289: \begin{theorem}[Katz]\label{katz} Suppose $\calf$ has uniform geometric
290: monodromy group $G_\geom$ and arithmetic monodromy group $G$. Let
291: $W\subset G$ be stable under $G$-conjugation. There are effective
292: constants $\delta(\calm,\calf)$ and $A(\calm/T)$ so that, if $k$ is a
293: finite field with $\abs k >A(\calm/T)$ and $t:\spec k \ra T$ is an
294: inclusion, then
295: \[
296: \abs{
297: \frac{\abs{\st{ x\in \calm_t(k): \rho_\calf(\frob_{x,k})\in
298: W}}}{\abs{\calm_t(k)}}-
299: \frac{\abs{W^{\xi(k)}}}{\abs{G^{\xi(k)}}}
300: }
301: < \epsilon(\calm,\calf,k) \defeq \frac{\delta(\calm,\calf)}{\sqrt{\abs k}}.
302: \]
303: \end{theorem}
304:
305: \pf This is simply \cite[9.7.13]{katzsarnak}; see also
306: \cite[4.1]{chav}. While the result holds for any sheaf with finite
307: monodromy group, we will (almost; see Section \ref{subsececqn1}
308: below) always work with subgroups of $\gsp_{2g}(\Lambda)$.
309: \qed
310:
311:
312: Already, this deep theorem yields a method for computing the
313: proportion of curves in a family for which the $\ell$-Sylow subgroup
314: of the class group is isomorphic to a given group. Indeed, let $H$ be
315: any finite abelian group annihilated by $\ell^e$, and let
316: $\pi:\calc\ra\calm/T$ be a smooth, irreducible proper relative curve of
317: genus $g\ge 1$. For any finite field $k$ and $t\in T(k)$ we define
318: \begin{align}\label{defbeta}
319: \beta(\calc\ra\calm, t, \ell^e,H) & =
320: \frac{\abs{\st{x \in \calm_t(k): \jac(\calc_x)[\ell^e](k) \iso H}}}
321: {\abs{\calm(k)}}.
322: \end{align}
323: We will often assume the $k$-point $t:\spec k \ra T$ is fixed, and
324: simply write $\beta(\calc\ra\calm, k, \ell^e, H)$.
325:
326: There is a sheaf $\calf = \calf_{\calc,\ell^e}$ of abelian groups on
327: $\calm$ whose fiber at a geometric point $\bar x\in \calm$ is the
328: $\ell^e$-torsion of the Jacobian $\jac(\calc_{\bar x})[\ell^e]$; it
329: may be alternatively defined by $\calf = R^1\pi_!(\integ/\ell^e)$.
330: Suppose that this family has uniform geometric monodromy group
331: $G_\geom \subseteq \sp_{2g}(\integ/\ell^e)$ and arithmetic monodromy
332: group $G\subseteq \gsp_{2g}(\integ/\ell^e)$. Then Theorem \ref{katz} implies that, for any
333: sufficiently large finite field $k$ and fixed, suppressed $k$-point of
334: $T$, we have
335: \begin{align}\label{defepsilon}
336: \abs{\beta(\calc\ra\calm,k,\ell^e,H)-\frac{\abs{\st{ x\in G^{\xi(k)} : \ker(x-\id) \iso
337: H}}}{\abs{G^{\xi(k)}}}} &< \epsilon(\calm,\calf,k).
338: \end{align}
339: We will denote the right-hand term of this inequality by
340: $\epsilon_{\calc\ra\calm}(\ell^e,k)$. Note that for a fixed family of
341: curves $\calc\ra \calm$, as $\abs k \ra \infty$ we have
342: $\epsilon_{\calc\ra\calm}(\ell^e,k) \ra 0$.
343:
344: In the special case where $\ell H = 0$, the mod-$\ell$ monodromy group
345: is the full symplectic group, and $\st{k_n}$ is a collection of finite
346: fields equipped with maps to $T$; $\lim_{n\ra\infty}\abs{k_n} =
347: \infty$;
348: and, for $n\gg0$, $\abs{k_n}\equiv 1 \bmod \ell$; we have
349: \begin{align}\label{eqfullmono}
350: \lim_{n\ra\infty} \beta(\calc\ra\calm, k_n, \ell, H) &=
351: \frac{\abs{\st{ x \in \sp_{2g}(\ff_\ell) : \ker(x - \id) \iso
352: H}}}{\abs{\sp_{2g}(\ff_\ell)}}.
353: \end{align}
354: In the next section we explain how to compute the right-hand side of
355: \eqref{eqfullmono}.
356:
357: We collect the diverse notation and assumptions of this section in the
358: following:
359:
360: \begin{situation}\label{mysit} We suppose that $\ell$ is a fixed, odd prime;
361: $T$ is a
362: connected $\integ[1/\ell]$-scheme of finite type; $\calm \ra \calt$
363: is a smooth scheme with geometrically irreducible fibers; $\calc \ra
364: \calm$ is a proper, smooth relative curve of genus $g$;
365: $\calf_{\ell^e}$ is the sheaf of $\ell^e$-torsion on the Jacobian of
366: $\calc$; and $\calf$ has uniform geometric monodromy group $G_\geom$ and
367: arithmetic monodromy group $G$. We will say that $\calc \ra \calm$
368: has full $\ell^e$-monodromy if $G_\geom = \sp_{2g}(\integ/\ell^e)$.
369: Additionally, $\st{k_n}$ is a collection of finite fields, each
370: equipped with an inclusion $t_n:k_n \ra T$, such that
371: $\lim_{n\ra\infty}\abs{k_n} = \infty$, and we will often write
372: $\calm(k)$ for $\calm_{t_n}(k)$.
373: \end{situation}
374:
375: While any sequence of finite fields is allowed, psychologically it
376: seems to be easiest to think of either $\st{\ff_{q^{e_n}}}$, a tower of
377: extensions of a fixed finite field $\ff_q$, or $\st{\ff_{p_n}}$, a
378: collection of finite fields of ever-larger prime order.
379:
380: \section{Matrices with given fixed space}\label{secsp}
381:
382:
383: \def\sl{{\rm SL}}
384: \def\so{{\rm SO}}
385:
386: Katz's work on mod-$\ell$ monodromy reduces the calculation of $\beta$
387: to a calculation in a symplectic group $\sp_{2g}(\ff_\ell)$. In
388: this section we go to some length to calculate precisely the
389: proportion of elements with a certain behavior. We note that \cite{fw}
390: avoids these difficulties in two ways. First, the authors compute in
391: $\gl_n$, where the relevant Lie theory is more transparent,
392: rather than in $\sp_{2g}$. Second, they compute with
393: elements of $\mat_n$, with the hope that if $x$ is equidistributed in
394: $\gl_n$, then $x-\id$ is equidistributed in $\mat_n$. Unfortunately,
395: these choices mean that the conjectural description of class group
396: frequencies in \cite{fw} differs slightly from the actual frequencies;
397: we take up this point in more detail in Section \ref{subsecfw}.
398:
399: \subsection{Symplectic matrices over finite fields}\label{subsecsymp}
400:
401: Fix a finite field $\ff$ with $\ell$ elements, where $\ell$ is a power
402: of an odd prime. (In our applications $\ell$ will itself be prime,
403: but this assumption is not necessary for the present computation.)
404:
405: Our main goal is a formula for
406: \begin{eqnarray}\label{defalpha}
407: \alpha(g,r)&\defeq& \frac{\abs{\st{x\in \sp_{2g}(\ff) : \ker(x-\id)\iso \ff^r}}}{\abs{\sp_{2g}(\ff)}}.
408: \end{eqnarray}
409: The method presented here should work for any of the classical
410: families of finite groups of Lie type. Still, since it is the symplectic group which arises
411: most naturally in questions about the typical function field, we have
412: chosen to focus our efforts on groups of type $C$. The reader will
413: notice our heavy reliance on the paper \cite{springersteinberg} of
414: Springer and Steinberg.
415:
416: Consistent with the notation introduced in the previous section, we
417: view $\sp_{2g}(\ff)$ as the group of automorphisms of a
418: $2g$-dimensional $\ff$-vector space $V_g$ equipped with a symplectic
419: form $\ang{\cdot,\cdot}_g$. Unless otherwise noted, an $r$-subspace of
420: $V_g$ means any subspace $W\subset V_g$ for which
421: $(W,\ang{\cdot,\cdot}_g\rest W)\iso (V_r, \ang{\cdot,\cdot}_r)$.
422: We will need to isolate the subspace of $V_g$ on which a given
423: element $x\in \sp(V_g)$ acts unipotently.
424:
425: \begin{lemma}\label{eigenspace} Suppose $x\in \sp(V)\iso \sp_{2g}(\ff)$. Then there are subspaces $E_1(x)$
426: and $E_1(x)^\perp$ such that $V\iso E_1(X) \oplus E_1(x)^\perp$;
427: $x\rest{E_1(x)}$ is unipotent; and $x-\id$ is invertible on
428: $E_1(x)^\perp$.
429: \end{lemma}
430:
431: %% \begin{lemma} Suppose $x\in \sp(V)\iso \sp_{2g}(\ff)$. Then there are subspaces $E_1(x)$
432: %% and $E_1(x)^\perp$ such that $V\iso E_1(X) \oplus E_1(x)^\perp$;
433: %% $x\rest{E_1(x)}$ is unipotent; and $x-\id$ is invertible on
434: %% $E_1(x)^\perp$.
435: %% \end{lemma}
436:
437: \pf We assume that $x-\id$ is not invertible, as otherwise the statement is trivial.
438: Write $V_{\bar\ff}$ as the direct sum of generalized eigenspaces
439: for $x$, $V_{\bar\ff} = \oplus V_{\bar\ff}(\lambda)$ where
440: $V_{\bar\ff}(\lambda)$ is the kernel of $(x-\lambda\id)^{2g}$ on
441: $V_{\bar\ff}$.
442:
443: Suppose that $\lambda\mu\not = 1$. We will prove, by induction on
444: $m+n$, that
445: \[
446: \ip{\ker(x-\lambda\id)^m\rest{V_{\bar\ff}},
447: \ker(x-\mu\id)^n\rest{V_{\bar\ff}}} = 0.
448: \]
449: For the base case $m=n=1$,
450: suppose $xu = \lambda u$ and $x v = \mu v$. Because $x$ preserves the
451: symplectic form, we have $\ip{u,v} =
452: \ip{xu,xv} = \ip{\lambda u, \mu v} = \lambda\mu\ip{u,v}$. Since
453: $\lambda\mu\not = 1$, this forces $\ip{u,v} = 0$.
454:
455: We now treat the inductive step. Suppose that $u\in
456: \ker(x-\lambda\id)^m \rest{V_{\bar\ff}}$ and $v\in \ker(x-\mu\id)^n
457: \rest{V_{\bar\ff}}$. Without loss of generality, assume that $m\ge n
458: \ge 1$. Then $xu = u' + \lambda u$, where $u' \in
459: \ker(x-\lambda\id)^{m-1}\rest{V_{\bar\ff}}$, and $xv = v'+\mu v$ for
460: some $v'\in \ker(x-\mu\id)^{n-1}\rest{V_{\bar\ff}}$. (If $n=1$, this
461: simply means that $v'=0$.) We then have $\ip{u,v} =
462: \ip{xu,xv} = \ip{u'+\lambda u, v' + \mu v} = \ip{u',v'}+\lambda
463: \ip{u,v'} + \mu\ip{u',v} + \lambda\mu\ip{u,v}$. By the inductive
464: hypothesis, the first three terms in the last expression vanish. This
465: leaves us with $\ip{u,v} = \lambda\mu\ip{u,v}$; again, $\ip{u,v} = 0$.
466:
467: This shows that, if $\lambda\mu\not =1$, then
468: $\ip{V_{\bar\ff}(\lambda), V_{\bar\ff}(\mu)} = 0$. Since the pairing
469: $\ip{\cdot,\cdot}$ is nondegenerate on $V$, we conclude that the pairing
470: $\ip{\cdot,\cdot}: V_{\bar\ff}(\lambda) \cross
471: V_{\bar\ff}(\lambda\inv) \ra \bar\ff$ is nondegenerate. In
472: particular, $V_{\bar\ff}(1)$ is self-dual under $\ip{\cdot,\cdot}$.
473:
474: Now, the generalized eigenspace associated to $1$ is defined over
475: $\ff$; therefore, its orthogonal complement is, too. Returning to
476: the $\ff$-vector space $V$, we find that $E_1(x) \defeq
477: \ker(x-\id)^{2g}\subset V$ is a symplectic subspace of $V$.
478: Therefore, there exists a canonical decomposition $V = E_1(x)
479: \oplus E_1(x)^\perp$, where $1$ is not an eigenvalue of the action of
480: $x$ on $E_1(x)^\perp$.\qed
481:
482:
483: We define the following quantities associated to $\sp_{2g}(\ff)$. Let
484: $\nu(g)$ be the number of elements in $\sp_{2g}(\ff)$; let $U(g)$ be
485: the number of unipotent elements in $\sp_{2g}(\ff)$; and let $S(g,r)$
486: be the number of $r$-subspaces of $V_g$. Let $\Phi(g)$ be the number
487: of elements $x\in\sp_{2g}$ for which $x-\id$ is invertible, and let
488: $\phi(g) = \Phi(g)/\nu(g)$ be the proportion of symplectic matrices
489: with this property. For convenience, we define $\Phi(0) = 1$.
490:
491: Recall that $\alpha(g,r)$ is the proportion of elements $x\in
492: \sp_{2g}(\ff)$ for which $\ker(x-\id) \iso \ff^r$. Let $U(g,r)$ be
493: the number of {\em unipotent} elements $u$ of $\sp_{2g}(\ff)$ for
494: which $\ker(u-\id) \iso \ff^r$. These quantities enjoy the following
495: relations.
496:
497: \begin{lemma}\label{calcphi} With all notation as above, let $\lambda(j) =
498: \ell^{2j-1}(\ell^{2j}-1)$. Then $\nu(g) =
499: \prod_{j=1}^g \lambda(j)$; $S(g,r) = \nu(g)/(\nu(r)\nu(g-r))$; $U(j) =
500: \ell^{2j^2}$;
501: \begin{align}
502: \alpha(g,r) &= \oneover{\nu(g)}\sum_{j=1}^g S(g,j)
503: U(j,r)\Phi(g-j)\label{eqcalcalpha};\\
504: \intertext{and}
505: \Phi(g)& = \nu(g) - \sum_{j=1}^g S(g,j)U(j) \Phi(g-j).\label{eqcalcphi}
506: \end{align}
507: \end{lemma}
508:
509: \pf The calculation of $\nu$ and $S$ is standard geometric algebra
510: \cite[III.6]{artingeomalg}. One proves that the symplectic group acts simply
511: transitively on symplectic bases for $V_g$, and that $\lambda(g)$
512: counts the number of symplectic pairs in $V_g$. A theorem of
513: Steinberg (\cite[8.14]{humphreysconjclass} or \cite{springersteinberg}) says that the number of unipotent elements in a finite group
514: $G$ of Lie type is $\ell^{\dim G - \rank G}$. Therefore $U(g)$, the
515: number of unipotent elements in $\sp_{2g}(\ff)$, is $\ell^{2g^2}$.
516:
517: By Lemma \ref{eigenspace}, any $x\in
518: \sp_{2g}(\ff)$ determines a decomposition $V = E_1(x) \oplus
519: E_1(x)^\perp$, where $x$ acts unipotently on $E_1(x)$ and $(x-\id)$ is
520: invertible on $E_1(x)^\perp$.
521: Therefore, any element of the
522: symplectic group determines, and is determined by, the data of a
523: subspace $W\subset V$; a unipotent element $u\in \sp(W)$; and an
524: element $y\in \sp(W^\perp)$ for which $(y-\id)$ is invertible. If $x$
525: corresponds in this way to the triple $(W,u,y)$, then $\ker(x-\id) =
526: \ker(u-\id)\rest W$.
527:
528: Equations \eqref{eqcalcalpha} and \eqref{eqcalcphi} follow swiftly.
529: The right-hand side of \eqref{eqcalcalpha} enumerates all choices of
530: data $(W,u,y)$ where $W$ is a $j$-subspace of $V$, $u$ is a unipotent
531: element of $\sp(W)$ with $\ker(u-\id) \iso \ff^r$, and $y\in
532: \sp(W^\perp)$ with $y-\id$ invertible, all normalized by the size of
533: the symplectic group.
534:
535: To calculate $\Phi(g)$ and thus derive \eqref{eqcalcphi}, we simply subtract from
536: $\nu(g)$ the number of symplectic elements with nontrivial unipotent
537: part. We enumerate triples $(W,u,y)$ as before, where $W$ is a
538: positive-dimensional subspace of $V$. If $W \iso V_j$, then $U(j)$
539: counts the number of choices for $u$, while $\Phi(g-j)$ is, by
540: definition, the number of choices for $y$.\qed
541:
542: Equation \eqref{eqcalcalpha}, combined with Proposition \ref{calcu}
543: below, allows the explicit computation of $\alpha(g,r)$ in any
544: particular case. The results of this calculation for $g \le 3$ are
545: shown in Table \ref{bigtable}.
546:
547: \begin{prop}\label{calcu} The number of unipotent elements $u$ in $\sp_{2g}(\ff)$ such
548: that $\ker(u-\id)\iso \ff^r$ is
549: \begin{align}
550: U(g,r) &= \nu(g) \sum_{{\bf d}: 0 < d_1 \le d_2 \le \cdots \le d_r}
551: \left(
552: \ell^{\half 1 \left( \sum_i s_i^2 - \sum_i r_i^2 + \sum_{i\text{ even}}
553: r_i\right)}
554: \cdot
555: \prod_{i\text{ odd}} \nu(r_i/2)
556: \cdot
557: \prod_{i\text{ even}}\nu_\orth(r_i)
558: \right)\inv\label{eqbigmess}
559: \end{align}
560: where the sum is over all partitions ${\bf d}$ of $\dim V_g$ into $r$
561: parts such that odd parts occur with even multiplicity; $r_i = \abs{\st{
562: j : d_j =i }}$; $s_i= \sum_{j\ge i} r_i$; and
563: \[
564: \nu_\orth(n) = \begin{cases}
565: \ell^{m^2}\prod_{i=1}^m (\ell^{2i}-1) & n = 2m+1 \\
566: \ell^{m^2-2m}\prod_{i=1}^{m}(\ell^{2i}-1) & n = 2m
567: \end{cases}
568: .
569: \]
570: \end{prop}
571:
572:
573: \pf Let $G = \sp_{2g}$. We start by identifying the relevant
574: $G(\bar\ff)$-conjugacy classes of unipotent elements.
575: As in, say, \cite[6.20]{humphreysconjclass},
576: let
577: $\calu = \calu(G)$ be the unipotent variety of $G$; it parametrizes
578: all unipotent elements of $G$. Similarly, let $\caln$ be the
579: nilpotent variety of $\lie g$, the Lie algebra of $G$. The Cayley
580: transform is a $G$-equivariant isomorphism
581: \begin{equation*}
582: \xymatrix{
583: \calu \ar[r]& \caln\\
584: x
585: \ar@{|->}[r] & (1-x)(1+x)\inv.
586: }
587: \end{equation*}
588:
589:
590: Thus, it suffices to count those $y\in \caln(\ff)$ with
591: nullspace of rank $r$.
592:
593: Happily, enumeration of nilpotent elements is a classical result.
594: Moreover, the description makes it easy to pick out those with the
595: appropriate rank.
596: To give a nilpotent orbit in $\lie{s}\lie{l}_n$ is to describe its Jordan
597: normal form; a similar classification exists for arbitrary Lie groups. We have
598: the classical bijection
599: (\cite[5.1.1]{collingwoodmcgovern}, \cite[7.11]{humphreysconjclass})
600: between nilpotent orbits of $\lie g$ and the partitions of $2g$ for
601: which odd parts occur with even multiplicity.
602: The dimension of the nullspace of an element in a nilpotent
603: orbit corresponding to a given partition is the number of elements in
604: that partition. Therefore, the desired (geometric) nilpotent orbits are
605: represented by suitable partitions with exactly $r$ pieces. Each of
606: these conjugacy classes has a representative in $G(\ff)$, and
607: the summation in equation \eqref{eqbigmess} thus ranges over all
608: $G(\bar\ff)$-conjugacy classes of unipotent elements $x$ in $G(\ff)$
609: for which $\ker(x-\id) \iso \ff^r$.
610:
611: We now explain how these $G(\bar\ff)$ conjugacy classes behave
612: over $G(\ff)$, and compute the isomorphism class of the
613: centralizer (still in $G(\ff)$) of an element of such a
614: conjugacy class. By doing so, we are able to compute the size of the
615: relevant conjugacy class.
616:
617: We proceed as in \cite[IV.2]{springersteinberg}. Fix a geometric
618: conjugacy class corresponding to a partition ${\bf d}$ of $g$, and let
619: $I = I({\bf d})= \st{ i : i\text{ even and }r_i >0 }$. Jordan factors
620: corresponding to even members $d_i$ split into two conjugacy classes
621: over $\ff$. Therefore, to give a $G(\ff)$-conjugacy class inside the
622: $G(\bar\ff)$ conjugacy class ${\bf d}$ is to give a map of sets $c: I
623: \ra \st{-1,+1}$.
624:
625: Let $u$ be a representative for the $G(\ff)$-conjugacy class
626: corresponding to ${\bf d}$ and a choice of assignments $c$. A
627: theorem of Springer and Steinberg \cite[IV.2.26-8]{springersteinberg} computes the isomorphism class of
628: the centralizer $Z = Z_G(u)$ in $G$. It is the semidirect product of
629: a unipotent radical, $R$, and the centralizer $C$ of a certain
630: torus associated to $u$. (Note that \cite{springersteinberg} computes
631: the connected component of the centralizer, and then later accounts
632: for multiple components.) The dimension of the Lie algebra
633: of $R$ is $\half 1 \left( \sum_i s_i^2 - \sum_i r_i^2 + \sum_{i\text{ even}}
634: r_i\right)$. The reductive group $C$ is isomorphic to
635: \[
636: \prod_{i\text{ odd}} \sp_{r_i}(\ff) \cdot \prod_{i\text{ even}}
637: \O^{c(i)}_{r_i}(\ff).
638: \]
639: Here, if $r_i$ is even then $\O^{+1}_{r_i}(\ff)$ denotes the rank $r_i$
640: orthogonal group of Witt defect $0$ over $\ff$, while
641: $\O^{-1}_{r_i}(\ff)$ is the orthogonal group of Witt defect $1$. For
642: odd $r_i$, $\O^{\pm 1}_{r_i}(\ff)$ is the (unique) orthogonal group of
643: rank $r_i$.
644:
645: Therefore, the size of the set of elements in $G(\ff)$ which belong to
646: the $G(\bar\ff)$-conjugacy class represented by ${\bf d}$ is
647: \begin{align*}
648: %% \frac{\abs{\sp_{2g}(\ff)}}{\abs R \prod_{i\text{ odd}}
649: %% \abs{\sp_{r_i}(\ff)}} \cdot \left(
650: %% \sum_{c: I \ra \st{-1,+1}} \prod_{i\in I}
651: %% \abs{\O_{r_i}^{c(i)}(\ff)}\inv \right) &=
652: \sum_{c:I \ra \st{-1,+1}}
653: \frac{\abs{\sp_{2g}(\ff)}}{\abs R \prod_{i\text{ odd}}
654: \abs{\sp_{r_i}(\ff)}}
655: \prod_{i\in I}\abs{\O_{r_i}^{c(i)}(\ff)}\inv
656: &=
657: \frac{\nu(g)}{
658: \abs R \cdot
659: %\ell^{\half 1 \left( \sum_i s_i^2 - \sum_i r_i^2 + \sum_{i\text{ even}}
660: %r_i\right)}
661: \prod_{i\text{ odd}}\nu(r_i/2)} \cdot \prod_{i\in I}
662: \left(
663: \abs{\O_{r_i}^{(-1)}(\ff)}\inv + \abs{\O_{r_i}^{(+1)}(\ff)}\inv\right)
664: \\
665: &=
666: \frac{\nu(g)}{
667: \ell^{\half 1 \left( \sum_i s_i^2 - \sum_i r_i^2 + \sum_{i\text{ even}}
668: r_i\right)} \prod_{i\text{ odd}}\nu(r_i/2)} \cdot \prod_{i\in I}
669: \nu_\orth(r_i),
670: \end{align*}
671: where
672: \[
673: \nu_\orth(n) = \begin{cases}
674: \ell^{m^2}\prod_{i=1}^m (\ell^{2i}-1) & n = 2m+1 \\
675: \ell^{m^2-2m}\prod_{i=1}^{m}(\ell^{2i}-1) & n = 2m
676: \end{cases}
677: .
678: \]
679: Note that $\nu_\orth(2m+1)$ is simply the number of elements in
680: $\so_{2m+1}(\ff)$, while $\nu_\orth(2m+1)$ is the harmonic mean of
681: $\abs{\so^{(-1)}_{2m}(\ff)}$ and $\abs{\so^{(+1)}_{2m}(\ff)}$. By
682: summing over suitable geometric conjugacy classes ${\bf d}$ we obtain
683: equation \eqref{eqbigmess}.\qed
684:
685: \begin{lemma}\label{alphainf} The limits
686: \[
687: \phi(\infty)\defeq \lim_{g\ra\infty} \phi(g)\text{ and
688: }\alpha(\infty,r) \defeq \lim_{g\ra\infty}\alpha(g,r)
689: \]
690: exist.
691: \end{lemma}
692:
693: \pf By Lemma \ref{calcphi},
694: \begin{eqnarray*}
695: \phi(g) & = & \oneover{\nu(g)}(\nu(g) - \sum_{j=1}^g S(g,j)U(j) \Phi(g-j))
696: \\
697: & = & 1 - \sum_{j=1}^g \frac{U(j)\Phi(g-j)}{\nu(j)\nu(g-j)}.
698: \end{eqnarray*}
699: Now, $\Phi(g-j)$ is necessarily less than $\nu(g-j)$, while
700: $U(j)/\nu(j) < \ell^{-j}$. Therefore, $\lim_{g\ra\infty}\phi(g)$
701: exists. Similarly, consider
702: \[
703: \alpha(g,r) = \sum_{j=1}^g \frac{U(j,r)\Phi(g-j)}{\nu(j)\nu(g-j)}.
704: \]
705: Again, $U(j,r)/\nu(j) \le U(j) / \nu(j) < \ell^{-j}$, so that $\lim_{g\ra\infty}\alpha(g,r)$ converges.\qed
706:
707: %% \subsection{Finite abelian $\ell$-groups}
708: %% Given an arbitrary finite
709: %% abelian $\ell$-group $H$, let
710: %% \[
711: %% \alpha(g,H,\ell^e) = \frac{\abs{\st{x\in \sp_{2g}(\integ/\ell^e)
712: %% :\ker(x-\id) \iso H \bmod \ell^e}}}{\abs{\sp_{2g}(\integ/\ell^e)}}.
713: %% \]
714: %% While Section \ref{subsecsymp} explains how to compute this quantity
715: %% in the special case where $e=1$, we do not know a general answer.
716: %% Still, in any particular case it is not hard to calculate this value,
717: %% and we can give the following brief indications.
718:
719: %% First, $\alpha(g,H,\ell^n)$ stabilizes as $n$
720: %% gets large. Indeed, if $\ell^e H = 0$ then $\alpha(g,H,\ell^e) = \alpha(g,H,\ell^{e+m})$,
721: %% since the projection $\sp_{2g}(\integ/\ell^{e+n}) \ra
722: %% \sp_{2g}(\integ/\ell^e)$ is a surjection of groups.
723:
724: %% Second, the limit $\lim_{g\ra\infty}\alpha(g,H,\ell^e)$ exists. If we
725: %% define quantities such as $U(j,r,\ell^e)$ in the expected way, they
726: %% are sufficiently close to $U(j,r,\ell) = U(j,r)$ that the proof of
727: %% Lemma \ref{alphainf} goes through.
728:
729:
730: \subsection{Unitary groups}\label{subsecunitary}
731: The methods of Section \ref{subsecsymp} work for any family of
732: classical Lie groups. Since unitary groups also come up in certain natural
733: applications (see Section \ref{extrigonal}), we briefly indicate how the
734: argument works for $\u_n$. Because $\u_n$ is a twist of $\gl_n$, the
735: details are actually somewhat simpler. We preserve all notation from Section
736: \ref{subsecsymp}, using the subscript $\u$ to denote the appropriate
737: group.
738:
739: So, let $\u_n$ denote the unitary group in $n$ variables over $\ff$.
740: Implicit in this definition is a nontrivial involution $\sigma$ of
741: $\ff$; let $m$ be $\sqrt{\ell}$, the size of the fixed field of $\sigma$.
742: The number of elements in $\u_n$ is
743: \[
744: \nu_\u(n) = m^{\half 1(n^2-n)} \prod_{i=1}^n(m^i - (-1)^i);
745: \]
746: the number of unipotent elements is $U_\u(n) = m^{n^2-n}$; and
747: $S_\u(n,r) = \nu_\u(n)/(\nu_\u(r)\nu_\u(n-r))$. Moreover, the
748: number of unitary matrices for which $1$ is not an eigenvalue is
749: \[
750: \Phi_\u(n) = \nu_\u(n) - \sum_{j=1}^n S_\u(n,j)U_\u(j)\Phi_\u(n-j).
751: \]
752: Since the unitary group is a form of the general linear group,
753: unipotent classes are parametrized by (unrestricted) partitions of
754: $n$. Moreover, $\u_n(\bar\ff)$ and $\u_n(\ff)$ conjugacy coincide. The
755: centralizer of a unipotent element corresponding to the partition $0 <
756: d_1 \le \cdots \le d_r$ of $n$ is connected, and has
757: \[
758: m^{\sum s_i^2 - \sum r_i^2} \prod \nu_\u(d_i)
759: \]
760: elements. All other results of Section \ref{subsecsymp}, including
761: the existence of $\alpha_\u(\infty, r)$, carry over.
762:
763:
764:
765: \section{Class groups of families of curves}
766:
767: Let $H$ be a finite abelian group with $\ell^e H =0$. We would like
768: to compute the chance that $H$ is the $\ell$-part of the class group
769: of a function field. As ``sample space'' of function fields we choose
770: the fibers of
771: any relative curve $\calc \ra \calm/T$ as in Situation \ref{mysit}
772: with full $\ell^e$-monodromy.
773:
774: In practice, general families
775: of curves tend to have full $\ell^e$-monodromy; see, for instance, the
776: introduction to \cite{ekedahl}. As a concrete example, fix a natural
777: number $N\ge 3$ relatively prime to $p$ and consider $\cgn\ra \mgn$,
778: the universal curve of genus $g$ with principal Jacobi structure of
779: level $N$. By \cite[5.15-5.16]{dm}, this family of curves has
780: full $\ell^e$ monodromy. Indeed, any versal family of curves has full
781: monodromy at most primes \cite[2.2]{achhol}. We also expect
782: (see Section \ref{subseche}) that a general family of
783: hyperelliptic curves has full $\ell^e$-monodromy.
784:
785: The equidistribution results in the first section let us
786: detect the occurrence of $H$ in class groups of function fields.
787: Recall (Equation \eqref{defbeta}) that this is measured by
788: $\beta(\calc\ra\calm,k, \ell^e,H)$.
789:
790: \begin{theorem}\label{maintheorem} Let $H$ be a finite abelian $\ell$-group.
791: As in Situation \ref{mysit}, let $\calc \ra \calm/T$ be a relative
792: curve with full $\ell^e$-monodromy and let $\st{k_n}$ be a collection
793: of finite fields. Suppose that for $n\gg 0$, $\xi(k_n) \equiv \xi \bmod
794: \ell^e$. There exists an effective constant $\delta(\calc \ra \calm)$
795: so that, for $n$ sufficiently large,
796: \begin{align*}
797: \abs{\beta(\calc\ra\calm, k_n, \ell^e, H) - \alpha^\xi(g,H,\ell^e)} &< \epsilon_{\calc \ra \calm}(\ell^e,k_n) \defeq
798: \frac{\delta(\calc\ra\calm)}{\sqrt{\abs{k_n}}},\\
799: \intertext{and thus }
800: \lim_{n\ra\infty}\beta(\calc\ra\calm,k_n,\ell^e,H)& = \alpha^\xi(g,H,\ell^e),
801: \end{align*}
802: where
803: \[
804: \alpha^\xi(g,H,\ell^e) = \frac{\abs{\st{x \in
805: \gsp^\xi_{2g}(\integ/\ell^e) : \ker( x-\id) \iso H
806: }}}{\abs{\sp_{2g}(\integ/\ell^e)}}.
807: \]
808: In the special case where $e = 1$ and $\xi = 1$, this term is computed
809: by Lemma \ref{calcphi} and Proposition \ref{calcu}.
810: \end{theorem}
811:
812: \pf As above consider the lisse sheaf $\calf = \calf_{\calc,\ell^e}$
813: on $\calm$ which associates, to each geometric point $\bar x$, the
814: $\ell^e$-torsion of the Jacobian of $\calc_{\bar x}$. Let $x\in \calm(k_n)$
815: be any point. The divisor class group of $\calc_x$ is
816: $\jac(\calc_x)(k_n)$, the $k_n$-rational points of the Picard variety.
817: By definition, the $\ell^e$-torsion of this group is the subgroup of $\calf_x$ fixed by $\frob_{x/k_n}$. Thus, in the
818: notation of the first section, $\jac(\calc_x)[\ell^e](k_n) \iso
819: \ker (\rho_\calf(\frob_{x/k_n})-\id)$, and
820: \[
821: \beta(\calc\ra\calm, k_n, \ell^e, H) = \frac{\abs{\st{ x \in \calm(k_n) : \ker
822: (\rho_\calf(\frob_{x/k_n})-\id) \iso H}}}{\abs{\calm(k_n)}}.
823: \]
824: In general, Theorem \ref{katz} finishes the proof. For the special case where
825: $H$ is an elementary abelian $\ell$-group and $\abs{k_n}\equiv 1 \bmod \ell$,
826: Section \ref{secsp}
827: provides an algorithm for computing the appropriate quantity.\qed
828:
829: %% Note that if $\ell^{e-1}H$, then Theorem \ref{maintheorem} states that
830: %% $\alpha(g,H,\ell^e)$ is the proportion of curves for which the
831: %% $\ell$-Sylow subgroup of the class group is isomorphic to $H$.
832:
833: In some applications, it is useful to be able to consider a family of
834: curves with unbounded genus. To employ our methods, we
835: need the size of the field of constants to grow more swiftly than the
836: error terms $\epsilon$ of \eqref{defepsilon}.
837:
838: \begin{theorem}\label{thgenusunbound} Let $H$ be the elementary abelian $\ell$-group
839: $(\integ/\ell)^r$. As in Situation \ref{mysit}, let $\st{\calc_n \ra \calm_n/T_n}_{n\in\nat}$ be a collection of relative
840: smooth proper curves of genus $g_n$ with full $\ell$-monodromy, and
841: let $\st{k_n}$ be a collection of finite fields, each equipped with $t_n: \spec k_n \ra T_n$.
842: Suppose that $\lim_{n\ra\infty}g_n=\infty$;
843: $\lim_{n\ra\infty}\epsilon_{\calc_n \ra \calm_n}(\ell,k_n) = 0$; and
844: for $n\gg 0$, $\xi(k_n) = 1$. Then
845: \[
846: \lim_{n\ra\infty}\beta(\calc_n\ra\calm_n, k_n, \ell, H) =
847: \alpha(\infty,r),
848: \]
849: where $\alpha(\infty,r)$ is computed in Lemma \ref{alphainf}.
850: \end{theorem}
851:
852: \pf The analysis is the same as that in Theorem \ref{maintheorem}. For $n$
853: sufficiently large that $\xi(k_n) = 1$, we have
854: \[
855: \abs{\beta(\calc_n \ra \calm_n, k_n,\ell,H) -
856: \alpha(g_n, r)} <
857: \epsilon_{\calc_n \ra
858: \calm_n}(\ell,k_n)
859: .\]
860: By Lemma \ref{alphainf}, $\lim_{n\ra\infty} \alpha(g_n,r)$ exists,
861: with limit $\alpha(\infty,r)$. By hypothesis,
862: $\lim_{n\ra\infty}\epsilon_{\calc_n \ra \calm_n}(\ell,k_n)=0$; the
863: theorem then follows.\qed
864:
865: As predicted in \cite{fw}, the divisor class groups of curves satisfy
866: a Cohen-Lenstra type result. Recent research also addresses the
867: distribution of other ideal class groups of function fields
868: \cite{cardonmurty, friesen01, pacelli}. These studies work with an
869: explicit affine model for a family of curves. To ease notation
870: somewhat, we work with a relative curve $\calc \ra \calm/k_0$ over a
871: fixed finite field $k_0$, and specify an affine model by introducing a
872: nonempty collection of sections $S = \st{ \sigma_1, \cdots, \sigma_n :
873: \calm \ra \calc}$ with disjoint image. Since we need to pass to
874: extension fields to apply our main result, we assume each
875: $\sigma_i$ is defined over the base field, $k_0$. For a curve $C$ and a
876: nonempty finite set of points $S$, let $\calo_{C,S} = \cap_{P\not\in
877: S} \calo_P$ be the ring of functions regular outside $S$. Let
878: $\cl(\calo_{C,S})$ be the ideal class group of this Dedekind domain,
879: and let $\cl(\calo_{C,S})_\ell$ be the $\ell$-Sylow part of that
880: group.
881:
882: The techniques of this paper don't yield exact formulae for the
883: frequency with which a given group $H$ occurs as
884: $\cl(\calo_{C,S})_\ell$. Still, we can at least give bounds
885: for the occurrence of $\ell$-Sylow subgroups of
886: given rank; these bounds are nontrivial if the genus of $C$ is larger than $\abs S$. Let $\rank_\ell H =
887: \dim_{\integ/\ell}H/\ell H$.
888:
889: \begin{cor}\label{coraffine} Let $\calc\ra
890: \calm/k_0$ be a smooth proper relative curve of genus $g$ with full $\ell$-monodromy.
891: Let $S$ be a nonempty finite set of
892: sections $\sigma : \calm \ra \calc$ with disjoint image inside
893: $\calc$, and let $S_x = \cup_{\sigma\in S}\sigma(x)$. Let $k$ be a sufficiently large finite extension of $k_0$
894: with $\abs{k} \equiv 1 \bmod \ell$. For any nonnegative integer $r$,
895: \begin{align}
896: \frac{\abs{\st{x\in \calm(k) : \rank_\ell \cl(\calo_{\calc_x,S_x}) \le r}}}{\abs{\calm(k)}}
897: &\ge
898: \sum_{j = 0}^r \phi(g,j) - \epsilon_{\calc \ra \calm}(k,\ell),\label{eqsmallrank}\\
899: \intertext{while}
900: \frac{\abs{\st{x\in \calm(k) : \rank_\ell
901: \cl(\calo_{\calc_x,S_x}) \ge r}}}{\abs{\calm(k)}}
902: & \ge \sum_{j=r+\abs S}^g \phi(g,j) - \epsilon_{\calc \ra \calm}(k,\ell).\label{eqbigrank}
903: \end{align}
904: \end{cor}
905:
906: \pf Given Theorem \ref{maintheorem}, all that's necessary is to relate the
907: ideal class group to the divisor class group.
908: By a theorem of F. K. Schmidt
909: \cite[proposition 1]{rosensclass}, there is an exact sequence of groups
910: \begin{equation}\label{classdiag}
911: \xymatrix{
912: 0 \ar[r] &
913: 0 \ar[r] & \frac{\cald(\calc_x,S_x)^0}{\calp(\calc_x,S_x)}\ar[r] & \jac(\calc_x)(k) \ar[r]&\cl(\calo_{\calc_x,S_x})\ar[r]
914: & 0,
915: }
916: \end{equation}
917: where $\frac{\cald(\calc_x,S_x)^0}{\calp(\calc_x,S_x)}$ is the class
918: group of divisors of degree zero represented by divisor classes
919: supported at $S$. (The sequence \eqref{classdiag} is exact on the right because the
920: sections $\sigma$ are defined over $k$.) On one hand, this shows that the
921: $\ell$-rank of the ideal class group is no bigger than that of the the
922: divisor class group. On the other hand, the $\ell$-rank of the kernel
923: of the surjection $\jac(\calc_x)(k) \ra \cl(\calo_{\calc_x, S_x})$ is
924: at most $\abs S$. These two observations yield inequalities
925: \eqref{eqsmallrank} and \eqref{eqbigrank}, respectively.\qed
926:
927: \section{Examples}
928:
929:
930: We conclude by working out some examples of these considerations.
931: Specifically, we show how Theorem \ref{maintheorem} and its variants,
932: in conjunction with the calculations in Section \ref{secsp}, let us
933: recover results of \cite{lenstra}; justify a heuristic used in
934: \cite{gekeler}; improve the main results of \cite{cardonmurty}, and a
935: special case of \cite{pacelli}; and discuss the conjecture
936: of \cite{fw}.
937:
938: For the most part, we work over a fixed finite field $k \iso \ff_q$.
939: We often phrase our results in terms of $\alpha(g,r)$, which is
940: computed by Lemma \ref{calcphi} and Proposition \ref{calcu}. Values of $\alpha(g,r)$
941: for $g \le 3$ are shown in Table \ref{bigtable}.
942:
943: \begin{table}
944: \[
945: \begin{array}{||ll|l||}
946: \hline\hline
947: g & r & \alpha(g,r)\\
948: \hline
949: 1 & 0 &{\frac {{\ell}^{2}-\ell-1}{{\ell}^{2}-1}}\\
950: 1 & 1 &\frac{1}{\ell}\\
951: 1 & 2 &{\frac {1}{\ell \left( {\ell}^{2}-1 \right) }}\\
952: \hline
953: 2 & 0 &{\frac {{\ell}^{6}-{\ell}^{5}-{\ell}^{4}+\ell+1}{
954: \left( {\ell}^{2}-1 \right) \left( {\ell}^{4}-1 \right) }}\\
955: 2 & 1 &{\frac {{\ell}^{3}-\ell-1}{{\ell}^{2} \left( {\ell}^{2}
956: -1 \right) }}\\
957: 2 & 2 &{\frac {{\ell}^{3}-\ell-1}{{\ell}^{2} \left( {\ell}^{2}
958: -1 \right) ^{2}}}\\
959: 2 & 3 &{\frac {1}{ \left( {\ell}^{2}-1 \right) {\ell}^{4
960: }}}\\
961: 2 & 4 &{\frac {1}{{\ell}^{4} \left( {\ell}^{2}-1
962: \right) \left( {\ell}^{4}-1 \right) }}\\
963: \hline
964: 3 & 0 &{\frac {{\ell}^{12}-{\ell}^{11}-{\ell}^{10}+{\ell}^{7}+
965: {\ell}^{5}+{\ell}^{4}-{\ell}^{3}-\ell-1}{ \left( {\ell}^{2}-1 \right) \left( {\ell}^{4}
966: -1 \right) \left( {\ell}^{6}-1 \right) }}\\
967: 3 & 1 &{\frac {{\ell}^{8}-{\ell}^{6}+{\ell}^{2}-{\ell}^{5}+\ell-{
968: \ell}^{4}+1}{{\ell}^{3} \left( {\ell}^{2}-1 \right) \left( {\ell}^{4}-1 \right) }
969: }\\
970: 3 & 2 &{\frac {{\ell}^{8}-{\ell}^{6}+{\ell}^{2}-{\ell}^{5}+\ell-{
971: \ell}^{4}+1}{{\ell}^{3} \left( {\ell}^{2}-1 \right) ^{2} \left( {\ell}^{4}-1
972: \right) }}\\
973: 3 & 3 &{\frac {{\ell}^{5}-{\ell}^{3}-1}{{\ell}^{7} \left( {
974: \ell}^{2}-1 \right) ^{2}}}\\
975: 3 & 4 &{\frac {{\ell}^{5}-{\ell}^{3}-1}{{\ell}^{7} \left( {
976: \ell}^{2}-1 \right) ^{2} \left( {\ell}^{4}-1 \right) }}\\
977: 3 & 5 &{\frac {1}{ \left( {\ell}^{2}-1 \right)
978: \left( {\ell}^{4}-1 \right) {\ell}^{9}}}\\
979: 3 & 6 &{\frac {1}{{\ell}^{9} \left( {\ell}^{2}-1
980: \right) \left( {\ell}^{4}-1 \right) \left( {\ell}^{6}-1 \right) }}\\
981: \hline\hline
982: \end{array}
983: \]
984: \caption{\label{bigtable}The proportion of symplectic matrices of dimension $2g$ with
985: fixed subspace of exact dimension $r$, as computed in Lemma
986: \ref{calcphi} and Proposition \ref{calcu}.}
987: \end{table}
988:
989:
990:
991: \subsection{Elliptic curves, $q\equiv 1 \bmod \ell$}
992:
993: The $\ell$-torsion of a random elliptic curve $E$ is
994: \begin{equation}\label{eqecq1}
995: E[\ell](k) \iso \left\{
996: \begin{array}{ll}
997: \st 1 & \text{with probability close to } {\frac {{\ell}^{2}-\ell-1}{{\ell}^{2}-1}}\\
998: \integ/\ell & \text{with probability close to }\oneover\ell \\
999: (\integ/\ell)^2 & \text{with probability close to } \oneover{\ell(\ell^2-1)}
1000: \end{array}
1001: \right.
1002: \end{equation}
1003: in the following sense.
1004:
1005: Let $\cale \ra \calm$ be a non-isotrivial family of elliptic curves,
1006: such as the Legendre family (with affine model) $y^2 =
1007: x(x-1)(x-\lambda)$ over the $\lambda$-line. Such a family is
1008: versal, and therefore \cite[2.2]{achhol} has full $\ell$-monodromy for
1009: almost all $\ell$; fix one such $\ell$. With a slight simplification
1010: of the notation of Equation \eqref{defbeta}, let
1011: \[
1012: \beta_{\cale\ra\calm}(k,r) = \frac{\abs{\st{x\in\calm(k):
1013: \cale_x[\ell](k)\iso (\integ/\ell)^r}}}{\abs{\calm(k)}}
1014: \]
1015: be the proportion of elliptic curves in our family, defined over $k$,
1016: for which the $\ell$-torsion subgroup is isomorphic to
1017: $(\integ/\ell)^r$. Suppose that $\abs k \equiv 1 \bmod \ell$ and
1018: $\abs k$ is sufficiently large. Then Theorem \ref{maintheorem} says that
1019: \[
1020: \abs{\beta_{\cale\ra\calm}(k,r) - \alpha(1,r)} \le
1021: \epsilon_{\cale\ra\calm}(\ell,k),
1022: \]
1023: where the error term decays as $1/\sqrt{\abs k}$, and $\alpha(1,r)$,
1024: defined in Equation \eqref{defalpha}, may be read off from the first
1025: section of Table \ref{bigtable}.
1026:
1027: \subsection{Elliptic curves, $q\not\equiv 1 \bmod \ell$}\label{subsececqn1}
1028:
1029: In the situation $\cale \ra \calm$ considered above, suppose that $k$
1030: is a large finite field for which $\abs{k}\equiv \xi \not\equiv 1
1031: \bmod\ell$. Again, Theorem \ref{maintheorem} says that
1032: \begin{equation}\label{eqecqn1}
1033: \abs{\beta_{\cale\ra\calm}(k,r) - \alpha^\xi(1,r)} \le
1034: \epsilon_{\cale\ra\calm}(\ell,k),
1035: \end{equation}
1036: where
1037: \[
1038: \alpha^\xi(1,r) \defeq\frac{\abs{\st{x \in \gsp_2^\xi(\integ/\ell) :\ker(x-\id)
1039: \iso (\integ/\ell)^r }}}{\abs{\sl_2(\integ/\ell)}}.
1040: \]
1041: We have not computed $\alpha^\xi(g,r)$ in general, but it is not hard
1042: to compute $\alpha^\xi(1,r)$ directly (see also \cite[3.3]{achhol}):
1043: if $\xi\not = 1$, then
1044: \begin{equation}\label{eqcalcalphaxi}
1045: \alpha^\xi(1,r)
1046: =
1047: \left\{
1048: \begin{array}{ll}
1049: \frac{\ell-2}{\ell-1}& r = 0\\
1050: \oneover{\ell-1}& r = 1 \\
1051: 0& r =2
1052: \end{array}
1053: \right..
1054: \end{equation}
1055: Note that this is compatible with the familiar result (use the Weil
1056: pairing) that if $k$ has no $\ell^{th}$ root of unity, then an
1057: elliptic curve over $k$ cannot have all its $\ell$-torsion defined
1058: over $k$.
1059:
1060: Taken together, Equations \eqref{eqecq1}, \eqref{eqecqn1} and
1061: \eqref{eqcalcalphaxi}, in the special case where $k$ is a prime field
1062: $\ff_p$, fully recover Theorem 1.14 of \cite{lenstra}.
1063:
1064: In \cite{gekeler}, Gekeler studies the distribution of Frobenius
1065: elements of elliptic curves over $\ff_p$, taken as elements of
1066: $\gl_2(\integ_\ell)$. Among
1067: other results, he computes the proportion of elements in
1068: $\gl_2(\integ/\ell^e)$ with given trace and determinant. (This is
1069: easier than the analogous question in $\sp_{2g}(\integ/\ell^e)$, first
1070: because conjugacy and stable conjugacy coincide in $\gl_2$, and second
1071: because of the severe constraints on Jordan blocks of $2\times 2$
1072: matrices.) Combining \cite[4.4]{gekeler} and Theorem
1073: \ref{maintheorem} allows one to compute the proportion of elliptic
1074: curves with $\ell^e$-torsion isomorphic to a given abelian $\ell$-group
1075: $H$.
1076:
1077: Moreover, we can justify a heuristic
1078: used in section 3 of \cite{gekeler}. There, it is asserted that if
1079: $m$ and $n$ are relatively prime, then for a fixed elliptic curve
1080: $E/k$, the actions of Frobenius on $E[m](\bar k)$ and $E[n](\bar k)$
1081: are independent, at least if $m\cdot n$ is small relative to
1082: $\sqrt{\abs k}$. Indeed, let $\calc \ra \calm\ra k_0$ be any
1083: relative curve with full $mn$-monodromy; for simplicity, assume that
1084: $\abs{k_0}\equiv 1 \bmod mn$. Then Frobenius elements of Jacobians
1085: of curves $\calc_x$ are, by Theorem \ref{katz}, equidistributed in
1086: $\sp_{2g}(\integ/mn)$. Since this latter group is isomorphic to
1087: $\sp_{2g}(\integ/m)\oplus \sp_{2g}(\integ/n)$, Gekeler's claim
1088: follows.
1089:
1090:
1091: \subsection{Quadratic function fields}\label{subseche}
1092:
1093: %If $\deg M = d$ is even, this is called a real quadratic field.
1094: %If $\deg M = d$ is odd, then imaginary. In the former case, the genus is $d/2-1$; in the latter, $(d-1)/2$. In general, the genus is $\floor{\half{d-1}}$.
1095:
1096: Attention has recently turned to the explicit construction of ideal
1097: classes of given order in the class groups of quadratic function
1098: fields $\ff_q(x,\sqrt{f(x)})$. Friesen computes both empirical
1099: \cite{friesen00} and analytic \cite{friesen01} bounds for the chance
1100: that $\ell$ divides the class number of $\ff_q(x)[y]/(y^2 - f(x))$,
1101: where $f$ is a quartic polynomial. Cardon and Murty
1102: \cite{cardonmurty} show that there are at least $q^{d (\half 1 +
1103: \oneover \ell)}$ imaginary quadratic extensions $K = \ff_q(x,\sqrt
1104: {f(x)})$ of $\ff_q(x)$ where $\deg f \le d$ and the ideal class group
1105: of $K$ has an element of prime order $\ell \ge 3$. While as $q$
1106: gets large this produces arbitrarily large families of quadratic
1107: function fields with class number divisible by $\ell$, it is a
1108: vanishingly small proportion of all quadratic function fields.
1109:
1110: We can use Corollary \ref{coraffine} to compute the {\em
1111: proportion} of quadratic function fields with class number divisible
1112: by $\ell$, and thereby strengthen these results.
1113:
1114:
1115: Suppose $q$ is a power of an odd prime and that $q\equiv 1 \bmod
1116: \ell$. We let $k = \ff_q$, and let $\st{k_n}$ be any collection of
1117: finite extensions of $k$ with $\lim_{n\ra\infty} \abs{k_n} = \infty$.
1118:
1119: Let $\calh_d$ be the space of separable monic polynomials $f(x)$ of
1120: degree $d$. Over it lies $\calc_d$, the curve with affine model $y^2
1121: = f(x)$; it is a hyperelliptic curve of genus $\floor{\half{d-1}}$.
1122: We work under the hypothesis that $\calc_d \ra \calh_d$ has full
1123: $\ell$-monodromy. For
1124: odd $d$, this is implied by unpublished work of J.K. Yu
1125: \cite[10.5.10]{katzsarnak}; we will treat the general case in a
1126: future work.
1127:
1128: The function field of the curve with affine model $y^2 = f(x)$ is
1129: called an imaginary quadratic function field if $d = \deg f$ is odd,
1130: and a real quadratic function field otherwise. We address these cases
1131: separately.
1132:
1133:
1134: If $d$ is odd, then there is a single point ``at infinity'' in this
1135: affine model; the left-hand term of \ref{classdiag} is
1136: trivial, and the ideal class group of this ring is isomorphic to the
1137: $\ff_q$-rational points of the Jacobian of the associated proper
1138: curve. We see that, for instance,
1139: \[
1140: \lim_{n\ra\infty}\frac{\abs{\st{f(x) \in k_n[x]: \deg f = d,
1141: f\text{ monic}, \ell |
1142: \cl(k_n[x,\sqrt{f(x)}])}}}
1143: {{\abs{\st{f(x) \in k_n[x] :
1144: \deg f = d, f\text{ monic}}}}}
1145: \]
1146: is equal to
1147: \[
1148: \lim_{n\ra\infty}\frac{\abs{\st{f(x) \in k_n[x]: \deg f = d,
1149: f\text{ monic, separable}, \ell |
1150: \cl(k_n[x,\sqrt{f(x)}])}}}
1151: {{\abs{\st{f(x) \in k_n[x] :
1152: \deg f = d, f\text{ monic, separable}}}}},
1153: \]
1154: since most polynomials are separable, which is in turn equal to
1155: \[
1156: \lim_{n\ra\infty} \frac{\abs{ \st{ x\in \calh_d(k_n): \ell |
1157: \abs{\jac(\calc_x)[\ell](k_n)}}}}{\abs{\calh_d(k_n)}},
1158: \]
1159: or $1-\alpha(g,0)$.
1160:
1161:
1162: If $d$ is even, then there are two points at infinity, and the
1163: regulator term in \ref{classdiag} is an abelian group on a single
1164: generator.
1165: Therefore, for any curve $C/k_n$ with affine model $C^{{\rm aff}}: y^2 =
1166: f(x)$, we have
1167: \[
1168: \rank_\ell(\jac(C)[\ell](k_n)) \ge
1169: \rank_\ell(\cl(\calo_{C^{{\rm aff}}}))\ge
1170: \rank_\ell(\jac(C)[\ell](k_n)) - 1,
1171: \]
1172: and the chance that $\ell$ divides the class group of the affine
1173: coordinate ring is bounded from below by
1174: \[
1175: \sum_{r=2}^g \alpha(g,r) - \epsilon_{\calc_d \ra \calh_d}(k_n,\ell).
1176: \]
1177: Thus, we can significantly strengthen the main conclusion of
1178: \cite{cardonmurty}; as $n\ra\infty$, there is an element of order
1179: $\ell$ in the class group of $k_n[x][y]/(y^2 -f(x))$ for a positive
1180: proportion of monic degree $d$ polynomials $f(x)\in k_n[x]$.
1181:
1182: Note that Theorem \ref{thgenusunbound} allows one to make uniform statements about
1183: curves of the form $y^2 = f(x)$ as $\deg f \ra \infty$, provided that
1184: the size of the base field grows sufficiently quickly.
1185:
1186: \subsection{Cyclic cubic fields}\label{extrigonal}
1187: Pacelli \cite{pacelli} looks at curves $y^d = f(x)$, and obtains
1188: results (for general $d$) similar to those of \cite{cardonmurty} for $d=2$.
1189: As mentioned in the introduction, the general family of curves (with
1190: affine model) $y^d = f(x)$ cannot have full $\bmod\ \ell$ monodromy,
1191: because of the extra automorphisms this family possesses. Still, by
1192: computing in the appropriate monodromy group one can calculate
1193: divisibility of class numbers for these families. We expect
1194: \cite{zarendjac} that the monodromy group is a unitary group
1195: associated to $\rat(\zeta_d)$.
1196:
1197: We take up these considerations in the special case where $d=3$, the
1198: degree of $f$ is $4$, and $3$ is invertible in the base field. Let
1199: $\calc \ra \calp$ be the family of curves with affine model $y^3 =
1200: f(x)$, where $f$ ranges over all separable polynomials of degree $4$.
1201: Each fiber $\calc_x$ has genus $3$. Moreover, since there is an
1202: obvious action of a cyclic group of order $3$ on $\calc$, the Jacobian
1203: $\jac(\calc)$ admits an action by $\integ[\zeta_3]$. The action on
1204: the tangent space at the identity of any fiber has signature $(2,1)$,
1205: since actions of type $(3,0)$ are rigid. Therefore, under the Torelli
1206: map, $\calc \ra\calp$ becomes identified with an open subset of the
1207: Picard modular variety associated to $\integ[\zeta_3]$. Using
1208: transcendental arguments \cite{holzapfel} and the theory of
1209: compactification \cite {larsenthesis}, one knows that for almost all
1210: $\ell$, the full $\ell$-adic monodromy group of this family is
1211: $G(\integ_\ell)$, where $G$ is the unitary group in three variables
1212: associated to $\integ[\zeta_3]$.
1213:
1214: Suppose, then, that $\calc \ra \calp$ has $\ell$-monodromy group
1215: $G(\integ/\ell)$. If $\integ[\zeta_3]$ is inert at $\ell$, then
1216: $G(\integ/\ell)$ is an example of the unitary groups studied in
1217: Section \ref{subsecunitary}. In particular, we see that:
1218: \[
1219: \jac(y^3 =f(x))[\ell](k) \iso
1220: \left\{
1221: \begin{array}{ll}
1222: \st 1 & \text{with probability close to } {\frac {\ell \left( {\ell}^{5}-{\ell}^{3}-1 \right)
1223: }{ \left( \ell+1 \right) \left( {\ell}^{2}-1 \right) \left( {\ell}^{3}+1
1224: \right) }}\\
1225: (\integ/\ell)^2 & \text{with probability close to } {\frac {{\ell}^{5}-{\ell}^{2}+{\ell}^{4}-\ell-1}{{\ell}^{2
1226: } \left( \ell+1 \right) ^{2} \left( {\ell}^{2}-1 \right) }}\\
1227: (\integ/\ell)^4 & \text{with probability close to }{\frac {{\ell}^{3}+{\ell}^{2}-1}{ \left( \ell+1
1228: \right) ^{2} \left( {\ell}^{2}-1 \right) {\ell}^{3}}}\\
1229: (\integ/\ell)^6 & \text{with probability close to }{\frac {1}{{\ell}^{3} \left( \ell+1 \right)
1230: \left( {\ell}^{2}-1 \right) \left( {\ell}^{3}+1 \right) }}
1231: \end{array}
1232: \right..
1233: \]
1234: (If $\integ[\zeta_d]$ splits at $\ell$, then $G(\integ_\ell)$ is
1235: isomorphic to a general linear group, and a similar, but easier,
1236: calculation applies.)
1237:
1238:
1239: \subsection{The Friedman-Washington conjecture}\label{subsecfw}
1240:
1241: Let $\calc \ra \calm$ be a family of curves of genus $g$ with full
1242: $\ell$-monodromy; this corresponds to any suitably general family of
1243: curves. Let $k$ be a large finite field, say with $\abs k \equiv 1
1244: \bmod \ell$. Then Theorem \ref{maintheorem} says that the proportion
1245: of fibers $\calc_x$, for $x\in \calm(k)$, with $\jac(\calc_x)[\ell](k)
1246: \iso (\integ/\ell)^r$ is $\alpha(g,r)$; see Table \ref{bigtable} for
1247: the first few values of $\alpha(g,r)$.
1248:
1249:
1250: Friedman and Washington \cite{fw} give a conjectural
1251: description of the frequency with which a given abelian $\ell$-group
1252: occurs as the $\ell$-Sylow part of the divisor class group of a
1253: function field. While they formulate their conjecture in terms of
1254: hyperelliptic curves in order to preserve the analogy with the Cohen-Lenstra
1255: heuristics, all of their arguments depend on merely having a
1256: sufficiently general family of curves. Given our
1257: expectation (Section \ref{subseche}) that hyperelliptic curves behave,
1258: in terms of $\ell$-monodromy, like general curves, we compare the
1259: predictions of \cite{fw} to the results of Theorem \ref{maintheorem}
1260: for $\calc\ra\calm$ with full monodromy.
1261:
1262: To facilitate this comparison we estimate the chance that the
1263: $\ell$-part of the class group of a curve is trivial. Let
1264: $\phi_{\gl}(n)$ denote the proportion of elements $x\in \gl_n$ for
1265: which $x-\id$ is invertible. It is shown that for large $n$ $\phi_{\gl}(n)$
1266: approaches
1267: \[
1268: \til\phi_{\gl}(n) \defeq \prod_{j=1}^n ( 1 - \ell^{-j}).
1269: \]
1270: In the special case of genus $2$, Friedman and Washington
1271: predict that the proportion of curves with trivial $\ell$-class group
1272: is (close to) $\til \phi_{\gl}(4)$, while Theorem \ref{maintheorem} says
1273: that this proportion is actually $\alpha(2,0) = {\frac {{\ell}^{6}-{\ell}^{5}-{\ell}^{4}+\ell+1}{
1274: \left( {\ell}^{2}-1 \right) \left( {\ell}^{4}-1 \right) }}$. The
1275: gap between $\til\phi_{\gl}(4)$ and $\alpha(2,0)$ is of order
1276: $1/\ell^2$.
1277:
1278: Now, \cite[p.131]{fw} expresses the hope that this discrepancy disappears for large
1279: genus; unfortunately, this difference persists. The proportion of curves of
1280: (arbitrarily large) genus with trivial $\ell$-group approaches (the
1281: well-defined limit; see Lemma \ref{alphainf}) $\phi(\infty)$, which by
1282: \cite[3.3]{achhol} is $1 -\ell/(\ell^2-1) + O(1/\ell^3)$. The difference
1283: between the conjectural estimate of \cite{fw} and the actual value
1284: remains of order $1/\ell^2$, even as the genus of the curves in
1285: question gets arbitrarily large.
1286:
1287:
1288: \bibliographystyle{plain}
1289: \bibliography{jda}
1290:
1291: \end{document}
1292: