1: \documentclass[11pt,reqno]{amsart}
2: \setlength{\textwidth}{14cm}
3:
4: \usepackage[dvips]{graphicx,color}
5: \usepackage{amssymb,amscd,latexsym,epsfig}
6: \usepackage[arrow,matrix,line]{xy}
7: \usepackage[mathscr]{euscript}
8:
9: \DeclareFontFamily{OT1}{rsfs}{}
10: \DeclareFontShape{OT1}{rsfs}{n}{it}{<-> rsfs10}{}
11: \DeclareMathAlphabet{\curly}{OT1}{rsfs}{n}{it}
12:
13: \newcommand{\rt}[1]{\stackrel{#1\,}{\rightarrow}}
14: \newcommand{\Rt}[1]{\stackrel{#1\,}{\longrightarrow}}
15: \newcommand\comp{{\!\,}_{{}^\circ}}
16: \newcommand\B{\operatorname{Bl}}
17: \newcommand\C{\mathbb C}
18: \newcommand\Q{\mathbb Q}
19: \newcommand\R{\mathbb R}
20: \newcommand\G{\mathbb G}
21: \renewcommand\S{\mathscr S}
22: \newcommand\Z{\mathcal Z}
23: \newcommand\PP{\mathbb P}
24: \renewcommand\P{\mathcal P}
25: \newcommand{\T}{\mathcal T}
26: \newcommand\X{\curly X}
27: \newcommand\XX{\,\widehat{\!\curly X}{}}
28: \newcommand\Y{\curly Y}
29: \renewcommand\L{\mathcal L}
30: \newcommand\LL{\curly L}
31: \renewcommand\O{\mathcal O}
32: \newcommand\I{\curly I}
33: \newcommand\into{\hookrightarrow}
34: \newcommand\res{\arrowvert_}
35: \newcommand\To{\longrightarrow}
36: \newcommand\half{\frac12}
37: \newcommand\ip{{\mbox\,\_\hspace{-1.5mm}\shortmid\hspace{1.5pt}}}
38: \renewcommand\_{^{\ }_}
39: \newcommand{\mat}[4]{\left(\begin{array}{cc} \!\!#1 & #2\!\! \\ \!\!#3 &
40: #4\!\!\end{array}\right)}
41: \newcommand{\rk}{\operatorname{rank}}
42: \newcommand{\coker}{\operatorname{coker}}
43: \newcommand{\im}{\operatorname{im}}
44: \newcommand{\Hom}{\operatorname{Hom}}
45: \newcommand{\Ext}{\operatorname{Ext}}
46: \newcommand{\Aut}{\operatorname{Aut}}
47: \newcommand{\Pic}{\operatorname{Pic}}
48: \newcommand{\Proj}{\operatorname{Proj}\,}
49: \newcommand{\Spec}{\operatorname{Spec}\,}
50: \newcommand{\Hilb}{\operatorname{Hilb}}
51: \newcommand{\Grass}{\operatorname{Grass}}
52:
53: \makeatletter \@addtoreset{equation}{section} \makeatother
54: \renewcommand{\theequation}{\thesection.\arabic{equation}}
55:
56: \newtheorem{thm}[equation]{Theorem}
57: \newtheorem{lem}[equation]{Lemma}
58: \newtheorem{cor}[equation]{Corollary}
59: \newtheorem{prop}[equation]{Proposition}
60: \newtheorem{conj}[equation]{Conjecture}
61:
62: \theoremstyle{definition}
63: \newtheorem{defn}[equation]{Definition}
64: \newtheorem{rmk}[equation]{Remark}
65: \newtheorem{rmks}[equation]{Remarks}
66:
67:
68: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
69:
70: \title[The Hilbert-Mumford criterion]
71: {\textbf{A study of the Hilbert-Mumford criterion for the stability of
72: projective varieties}}
73: \author[J. Ross and R. P. Thomas]{Julius Ross and Richard Thomas}
74:
75: \begin{document}
76:
77: \begin{abstract} \noindent
78: We make a systematic study of the Hilbert-Mumford criterion for different
79: notions of stability for polarised algebraic
80: varieties $(X,L)$; in particular for K- and Chow stability. For each type of stability this leads to a concept of
81: slope $\mu$ for varieties and
82: their subschemes; if $(X,L)$ is semistable then $\mu(Z)\le\mu(X)$ for all
83: $Z\subset X$. We give examples such as curves, canonical models and Calabi-Yaus.
84: We prove various foundational technical results towards understanding
85: the converse, leading to partial results; in particular this gives a geometric
86: (rather than combinatorial) proof of the stability of smooth curves.
87: \end{abstract}
88:
89: \maketitle
90:
91:
92: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
93:
94:
95: \section{Introduction; slope stability}
96:
97: Geometric Invariant Theory \cite{GIT} has been very successful in forming
98: moduli spaces of (semi)stable coherent sheaves over polarised
99: algebraic varieties $(X,L)$ \cite{HL}. Moduli space is constructed as a quotient
100: of a subset of a Quot scheme by a group, and the Hilbert-Mumford criterion
101: is applied to 1-parameter subgroups. Their weights are found to be dominated
102: by positive linear combinations of weights of particularly simple 1-parameter
103: subgroups corresponding to a degeneration
104: of a sheaf $E$ into a splitting $F\oplus E/F$, for some subsheaf $F\le E$.
105: Thus stability of sheaves is governed by subsheaves; calculating the corresponding
106: weights leads to the notion of slope stability (the exact form of the slope
107: depending on the linearisation used on Quot).
108:
109: Forming moduli spaces of (semi)stable varieties themselves using GIT has
110: proved much more difficult, and has mainly been accomplished for canonically
111: polarised varieties using the Chow linearisation, due to work of Mumford
112: \cite{GIT, Mu}, Gieseker \cite{Gi}, and, in a little more generality, Viehweg
113: \cite{V}. Roughly
114: speaking one expects varieties polarised by their canonical bundle $L=K_X$
115: to be automatically stable since their moduli functor is
116: already separated because of the birational invariance of spaces of sections
117: of powers of the canonical bundle. In the general case no geometric
118: criterion for (in)stability has emerged. This is because the Hilbert-Mumford
119: criterion has not been successfully simplified or interpreted for varieties;
120: instead Viehweg proved deep positivity results to produce the group-invariant
121: sections of the appropriate line bundle directly (for varieties with semi-ample
122: canonical bundle). Kollar \cite{Ko} and others have turned to other methods
123: for producing moduli of varieties, but new impetus to understanding stability
124: has come from the link between K-stability and the existence
125: of K\"ahler-Einstein and constant scalar curvature K\"ahler metrics \cite{Ti1,
126: Do1} to which we apply our methods in \cite{RT}.
127:
128: Our approach uses the Hilbert-Mumford criterion, as pioneered by
129: Mumford \cite{Mu}. He calculates the relevant weights in terms of
130: blow-ups of $X\times\C$ in subschemes supported on thickenings of $X\times\{0\}$;
131: one of our main results (Corollary \ref{ref}) reduces the calculation to
132: a sum of weights of blow-ups \emph{in the scheme-theoretic central fibre}.
133: (In a sense which is made clear by the proof the construction
134: turns the horizontal thickenings of Mumford ``vertical", into the
135: central fibre.)
136:
137: Taking just one such blow-up in a subscheme $Z\subset X$ gives the ``deformation
138: to the normal cone of $Z$", analogous
139: to the simple 1-parameter subgroups that arise in the GIT of sheaves.
140: This gives a numerical condition for $Z$ to destabilise $X$, and so
141: a notion of ``slope stability", by analogy with the sheaf theory which we
142: now review briefly.
143:
144: For a sheaf $E$ over $(X,L)$, the \emph{reduced} Hilbert polynomial $p\_E$
145: is the monic version of the Hilbert polynomial
146: $\P\_{\!E}(r)=\chi(E\otimes L^r)=a_0r^n+a_1r^{n-1}+\ldots\,$:
147: $$
148: p\_E(r)=\frac{\chi(E\otimes L^r)}{a_0}=r^n+\mu(E)r^{n-1}+\ldots\,,
149: $$
150: and the \emph{slope} of $E$ is its leading nontrivial coefficient,
151: \begin{equation} \label{mu}
152: \mu(E)=a_1/a_0.
153: \end{equation}
154: (This differs from the usual definition deg$\,(E)/\rk(E)$ by unimportant
155: terms that depend only on the geometry of $(X,L)$.)
156: Then we say that $E$ is \emph{semistable} if for all proper coherent subsheaves
157: $F\le E$,
158: $$
159: p\_F\preceq p\_E\,.
160: $$
161: For \emph{Gieseker semistability}, $\preceq$ means there exists an $r_0>0$
162: such that $p\_F(r)\le p\_{E}(r)$ for all $r\ge r_0$; for \emph{slope semistability}
163: the inequality is at the level of the $r^{n-1}$ coefficients:
164: \begin{equation} \label{mu2}
165: \mu(F)\le\mu(E).
166: \end{equation}
167: If the inequalities are all strict then $E$ is \emph{stable}, and this agrees
168: with the appropriate GIT notions for different choices of linearisation (i.e.\
169: choice of equivariant line bundle on the Quot scheme; in fact Jun Li's
170: line bundle \cite{Li} is only semi-ample, but he extends GIT to this case.)
171:
172: Due to different choices
173: of linearisation and asymptotics there are also many notions of stability
174: (Hilbert, Chow, asymptotic Hilbert, asymptotic Chow and K-stability) for
175: varieties, defined in the next section, and so many versions of slope. We
176: first describe the one relevant to K-stability.
177:
178: Given a subscheme $Z$ of a polarised algebraic variety $(X,L)$, define
179: $$
180: \epsilon(Z)=\sup\{c\in\Q_{\,>0}\colon L^r\otimes\I_{\!Z}^{cr}\text{ is globally
181: generated } \forall r\gg0 \text{ with } cr\in\mathbb N\}.
182: $$
183: The Hilbert-Samuel polynomial for \emph{fixed} $x$,
184: \begin{equation} \label{HS}
185: \chi(L^r\otimes\I^{xr}_{\!Z})=a_0(x)r^n+a_1(x)r^{n-1}+\ldots\,,\quad r\gg0,
186: \ rx\in\mathbb N,
187: \end{equation}
188: defines $a_i(x)$ which are polynomials in $x$ (see Section \ref{K})
189: and so extend by the same formulae to
190: $x\in\R$. Then analogously to (\ref{mu}) we define the \emph{K-slope} of
191: $\I_Z$ (with respect to $L$ and $c\in(0,\epsilon(Z)]$) to be
192: \begin{equation} \label{def:slope}
193: \mu_c(\I_Z)=\mu_c(\I\_Z,L):=\frac{\int_0^c\big(a_1(x)+\frac{a_0'(x)}2\big)dx}
194: {\int_0^ca_0(x)dx}\,.
195: \end{equation}
196: Setting $Z=\emptyset$ defines the slope of $X$ (precisely: of $\O_X$ with
197: respect to $L$ and $c$) as
198: \begin{equation} \label{def:X}
199: \mu(X)=\mu(X,L)=\frac{a_1}{a_0}\,,
200: \end{equation}
201: independently of $c$. Here $a_i$ are the coefficients of the
202: Hilbert polynomial $\chi(L^r)=a_0r^n+a_1r^{n-1}+\ldots\,,\ a_0=a_0(0)$
203: and, for $X$ normal, $a_1=a_1(0)$ (\ref{normal}).
204:
205: Setting $\mu(Z):=\sup\_{0<c\le\epsilon(Z)}\mu_c(\I_Z)$, we say that
206: $(X,L)$ is \emph{K-slope semistable} if for all proper $Z\subset X$,
207: $$
208: \mu(Z)\le\mu(X), \qquad\qquad\text{i.e.}\quad \frac{\int_0^c\left(a_1(x)+
209: \frac{a_0'(x)}2\right)dx}{\int_0^ca_0(x)dx}\le\frac{a_1}{a_0}\quad
210: \forall c\in(0,\epsilon(Z)].
211: $$
212: Cf. (\ref{mu2}).
213: The definition of slope stability is slightly trickier (Definition \ref{defstab});
214: for this reason we work with $\mu_c(\I_Z)$ instead of $\mu(Z)$. Then (Theorem
215: \ref{thm:kstableslopestable}) $X$ is slope (semi)stable if it is K-(semi)stable.
216:
217: The $a_0'/2$ ``correction term" in the definition of slope arises from the
218: difference between the Hilbert polynomial of a 2-component normal crossing
219: variety and the sum of those of its components (whereas for sheaves, $\P\_{\!E}=
220: \P\_{\!F}+\P\_{\!E/F}$ for any $F\le E$). Simon Donaldson
221: pointed out that his analysis of stability of toric varieties \cite{Do2}
222: throws up a boundary term similar to our $(a_0(c)-a_0)/2=\int_0^ca_0'/2$;
223: we explain this in \cite{RT}.
224:
225: Similarly for $X\subseteq\PP^N=\PP(H^0(X,L)^*)$ (embedded in projective space
226: by the space of sections of its polarisation $L=\O_X(1)$ for ease of exposition;
227: see Section \ref{sec:chow} for the general case), a subscheme $Z\subset
228: X$ and an \emph{integer} $0<c\le\epsilon(Z)$,
229: we define the \emph{Chow slope} of $\I_Z$ to be
230: $$
231: Ch_c(\I_Z):=\frac{\sum_{i=1}^ch^0(\I_{\!Z}^i(1))}{\int_0^ca_0(x)dx}\,.
232: $$
233: Setting $Z=\emptyset$ gives $Ch(\O_X)=Ch(X):=\frac{h^0(\O_X(1))}{a_0}$;
234: then Chow (semi)stability implies Chow slope (semi)stability (Theorem
235: \ref{thmchow}):
236: $$
237: Ch_c(\I_Z){\ }^{\ <\,}_{(\le)}\ Ch(X).
238: $$
239: Asymptotic Chow is more like Gieseker stability and so our slope criterion
240: for that (\ref{chow}) is more complicated.
241:
242: Section \ref{K} describes the various stability notions uniformly, while
243: Section \ref{Kslope} introduces slope stability via the deformation to the
244: normal cone. Arbitrary 1-parameter subgroups are studied in Section \ref{testconfigs},
245: calculating their weights in terms of a sequence of simple blow ups in Corollary
246: \ref{ref}. In Section \ref{converse} these weights are shown to be those
247: of a deformation to the normal cone under certain circumstances, giving a
248: partial
249: converse to ``stability $\Rightarrow$ slope stability". We used to think
250: this could be done in general, but the failure of the \emph{thickenings}
251: of certain flat families of subschemes to themselves be flat prevents
252: us from carrying out the programme in full. We study
253: when the thickenings \emph{are} flat, and get round the problem with a basechange
254: trick in some situations. This deals with the curve case completely, i.e.\
255: stability and slope stability are equivalent there,
256: giving geometric proofs of the K- and asymptotic Chow stability of curves
257: of genus $\ge1$. As far as we know all previous proofs used analysis and
258: combinatorics
259: respectively. Section \ref{sec:chow} is devoted to Chow stability
260: and Section \ref{egs} to examples. Many more examples, such as projective
261: bundles, appear in \cite{RT}, in particular showing that K- and slope
262: stability are also equivalent for projective bundles over a curve.
263:
264: \noindent \textbf{Acknowledgements.}
265: We would like to thank Brian
266: Conrad, Kevin Costello, Dale Cutkosky, Simon Donaldson, David Eisenbud,
267: Daniel Huybrechts, Frances Kirwan, Miles Reid and
268: Bal\'azs Szendr\"oi for useful conversations. The book
269: \cite{HL} and paper \cite{Mu} have been very useful to us. The authors
270: were supported by an EPSRC PhD studentship and a Royal Society
271: university research fellowship respectively. The second author would also
272: like to thank Oscar Garcia-Prada and Peter Newstead for supporting a visit
273: to CSIC, Madrid, where much of this work was done.
274:
275: \section{Notation} \label{notation}
276:
277: Throughout this paper $Z$ will denote a closed subscheme of a proper
278: irreducible polarised scheme $(X,L)$ of dimension $n=\dim X$; for much of
279: the paper $X$ will be a normal irreducible variety. By $jZ$ we mean the subscheme
280: which is the $j$-fold thickening of $Z$ defined by its ideal sheaf $\I_{jZ}:=
281: \I_{\!Z}^j$. We denote
282: the blow up along $Z$ by $\pi\colon\widehat X\to X$, with exceptional divisor
283: $E$. Then $\pi^*\I_{\!Z}^j=\O(-jE)$ and there exists a $p$ such that
284: $\pi_*\O(-jE)=\I_{\!Z}^j$ for all $j\ge p$. For convenience
285: we often suppress pullback maps, mix multiplicative and additive
286: notation for line bundles, and use the same letter to denote a divisor
287: and the associated line bundle. For example on $\widehat{X}$, we
288: denote $(\pi^*L(-E))^{\otimes k}$ by $L^k-kE$. Worse still, where it
289: does not cause confusion, $L^k$ may also denote $c_1(L)^{\cap k}$.
290:
291: For brevity we often denote sheaf cohomology on a space $X$ by $H^i_X$;
292: this never means local cohomology.
293:
294: Any vector space $V$ with a $\C^\times$-action splits into one-dimensional
295: weight spaces $V=\bigoplus_iV_i$, where $t\in\C^\times$ acts on $V_i$ by
296: $t^{w_i}$. The $w_i$ are the \emph{weights} of the action, and $w(V)=\sum_iw_i$
297: is the \emph{total weight} of the action; i.e.\ the weight of the induced
298: action on $\Lambda^{\max}V$.
299:
300: On any family over $\C=\Spec\C[t]$, $t$ will denote the pullback of the
301: standard coordinate on $\C$.
302:
303: A line bundle $L$ is \emph{semi-ample} \cite{La} if its high
304: powers are globally generated (i.e.\ basepoint free). In this paper all
305: semi-ample line bundles will have the additional property that
306: the contraction they define is birational; that is $L$ is the pull back of
307: an ample line bundle from a birational scheme. (It is important that this
308: contraction can be trivial, i.e.\ semi-ample includes ample.)
309:
310: Given a (semi-)ample line bundle $L$ on $X$, the \emph{Seshadri constant}
311: of $Z$ is
312: \begin{eqnarray} \nonumber
313: \epsilon(Z) = \epsilon(Z,X,L)&=&\sup\,\{c\in\Q_{\,>0}\colon L^r\otimes
314: \I_{\!Z}^{cr} \text{ is globally generated for } r\gg0\} \\ \label{def:seshadri}
315: &=&\max\,\{ c \in \Q_{\,>0}: L - cE \,\, \text{is nef on }\widehat X \}.
316: \end{eqnarray}
317:
318: Given a pair of ideals $J\subset I\subset\O_X$, we say that $J$ \emph{saturates}
319: I if there exists $i>0$ such that $JI^{i-1}=I^i$. Equivalently,
320: on the blow up $p\colon\widehat X\to X$ of $X$ along $I$ with exceptional
321: divisor $E$, the natural inclusion $p^*J\to\O(-E)$ is an isomorphism.
322: (This equivalence is a tautology: the zero set of the sections
323: of $p^*J\subseteq\O(-E)$ has homogeneous ideal the saturation of
324: $\oplus_iJI^{i-1}\subseteq\oplus_iI^i$ (\cite{Ha} Exercise II.5.10); this
325: is all of $\oplus_iI^i$ if and only if the zero set is empty, if and only
326: if the sections generate $\O(-E)$.)
327:
328: Similarly, give a line bundle $L$ on $X$, the global sections of $L\otimes
329: I$ generate a subsheaf $L\otimes J\subset L\otimes I$, defining an ideal
330: $J$. We say that \emph{the global sections of $L\otimes I$
331: saturate} $I$ if $J$ saturates $I$. Equivalently, the global
332: sections of $L\otimes I$ generate the line bundle $L(-E)$ on $\widehat
333: X$. This is weaker than (i.e.\ is implied by) $L\otimes I$ being globally
334: generated.
335:
336:
337: \section{Hilbert, Chow and K-stability, and test configurations} \label{K}
338:
339: Fix a polynomial $\P$ of degree $n$ and consider any $n$-dimensional
340: proper polarised scheme $(X,L)$ whose Hilbert polynomial equals $\P$:
341: $$
342: \P(r)=\chi\_X(L^r) = a_0r^n+a_1r^{n-1}+\ldots\,,
343: $$
344: where
345: $a_0=\frac{1}{n!}\int_X c_1(L)^n=\frac{L^n}{n!}$ and, for smooth $X$,
346: $$
347: a_1=\frac{1}{2(n-1)!}\int_X c_1(X)c_1(L)^{n-1}=
348: -\frac{K_X.L^{n-1}}{2(n-1)!}\,.
349: $$
350: Fix $r>0$ such that $L^r$ is both very ample on $X$, and
351: \begin{equation} \label{rlarge}
352: H^i(L^r)=0 \quad\text{for }i>0, \quad\Rightarrow\ H^0(L^r)^*\cong\C^{\P(r)}.
353: \end{equation}
354: Then $L^r$ embeds $X$ in $\PP=\PP^{\P(r)-1}$, defining a point of the
355: Hilbert scheme $\Hilb=\Hilb_{\P',K}$ of subschemes of $\PP$ with
356: Hilbert polynomial $\P'(K)=\P(Kr)$. Then there is a $K_0$ such that for
357: \emph{all} points $\{X\}$ of Hilb and $K\ge K_0$ we have an exact sequence
358: \begin{equation} \label{eq:seshilb}
359: 0\to H^0_\PP(\I_X(K))\to S^K \C^{\P(r)*}\cong S^K H_X^0(L^r)\to
360: H^0_X(L^{Kr})\to0.
361: \end{equation}
362: This (see for example \cite{V}) defines Hilb as a closed subscheme of the
363: Grassmannian
364: $$
365: \G=\Grass(S^K \C^{\P(r)*},\P(Kr)).
366: $$
367:
368: So $(X,L^r)$ and a choice of isomorphism $H^0(L^r)^*\cong\C^{\P(r)}$
369: give us a point $x=x_{r,K}$ in $\G$, with different choices of isomorphism
370: corresponding (up to scale) to the orbits of
371: $SL(\P(r),\C)$ on $\PP^{\P(r)-1}$. A $g\in SL(\P(r),\C)$ acting on $\C^{\P(r)}$
372: induces an action
373: \begin{equation} \label{convention}
374: (S^Kg^*)^{-1} \quad\text{on }\ S^K \C^{\P(r)*},
375: \end{equation}
376: and so one on $\G=\Grass(S^K \C^{\P(r)*},\P(Kr))$. It is this action that
377: commutes with the action on $\Hilb\subset\G$ induced by that on $\PP$.
378: Points in the orbit of $x\in\G$ correspond to
379: the projective transformations of $X$.
380:
381: From (\ref{eq:seshilb}), the fibre over $x\in\Hilb$ of the
382: hyperplane bundle on $\G$ is naturally isomorphic to
383: \begin{equation} \label{eq:line}
384: \O_{\G}(1)\res{x}=\Lambda^{\max} H^0_X(L^{Kr})\otimes
385: \left(\Lambda^{\max} S^K H^0_X(L^r)\right)^*.
386: \end{equation}
387:
388: \begin{defn}
389: $(X,L)$ is Hilbert (semi)stable with respect to $r$ if the point $x_{r,K}\in\Hilb$
390: is GIT (semi)stable for the action of $SL(\P(r),\C)$, linearised on (\ref{eq:line}),
391: for all $K\gg0$.
392: \end{defn}
393:
394: \emph{Asymptotic Hilbert stability} is defined to mean Hilbert stability
395: for all sufficiently large $r$.
396: By picking a different line bundle on the Hilbert scheme (i.e.\ a different
397: projective embedding of Hilb -- the beautiful Chow embedding \cite{Mu})
398: we also get the notion of \emph{Chow stability} with respect to $r$
399: and \emph{asymptotic Chow stability}. We need the concept
400: of a \emph{test configuration}, as defined in the foundational paper \cite{Do2}.
401:
402: \begin{defn}
403: A \emph{test configuration} for a polarised variety $(X,l)$ is
404: \begin{enumerate}
405: \item A proper flat morphism $\pi\colon\X\to\C$,
406: \item An action of $\C^\times$ on $\X$ covering the usual
407: action of $\C^\times$ on $\C$,
408: \item An equivariant very ample line bundle $\L$ on $\X$,
409: \end{enumerate}
410: such that the fibre $(\X_t,\L|_{\X_t})$ is isomorphic to $(X,l)$ for
411: one, and so all, $t\in\C\backslash\{0\}$.
412:
413: A test configuration is called a \emph{product configuration}
414: if $\X\cong X\times\C$, and a \emph{trivial configuration} if in addition
415: $\C^\times$ acts only on the second factor.
416:
417: We will often need a weaker concept which we call a \emph{semi test configuration}
418: where $\L$ is just globally generated.
419: \end{defn}
420:
421: \begin{prop}
422: In the situation of (\ref{rlarge}), a 1-parameter subgroup of $GL(\P(r),\C)$
423: is equivalent to the data of a test configuration for $(X,L^r)$.
424: \end{prop}
425:
426: \begin{proof}
427: The action of a 1-parameter subgroup of $GL(\P(r),\C)$ on
428: $X\subseteq\PP(H^0_X(L^r)^*)$
429: clearly gives a test configuration $(\X,\L)$ for $(X,L^r)$ with $\L$ the
430: pull back of the (equivariant) line bundle $\O_\PP(1)$.
431:
432: Conversely the subgroup
433: can be recovered from the test configuration via the induced $\C^\times$-action
434: on the dual of the vector space $(\pi_*\L)|_{\{0\}}$. Since $\pi_*\L$ is
435: torsion-free (by flatness) over the curve $\C$ it is a vector bundle, and
436: $\X$ sits inside the projectivisation of its dual by the very ampleness of
437: $\L$. Its general fibre has dimension $\P(r)$, so $(\pi_*\L)|_{\{0\}}$ does
438: too.
439:
440: Pick a trivialisation of $\pi_*\L$ over $\C$, identifying it with
441: $(\pi_*\L)|_{\{0\}}\times\C$. This has a diagonal $\C^\times$-action, inducing
442: one on $\PP(\pi_*\L)^*\supset\X$ which yields the original test configuration;
443: thus these two operations are mutual inverses.
444:
445: (Note that in fact $(\pi_*\L)|_{\{0\}}\cong H^0_\X(\L)\big/tH^0_\X(\L)$.
446: The map $\leftarrow$ is just restriction; we need to define its inverse $\to$.
447: Any element of $(\pi_*\L)|_{\{0\}}$ can be lifted to a meromorphic section
448: $s$ of $\L$ on $\X$ that is regular on $\X_0$.
449: The projection of its polar locus to $\C$ does not contain $\{0\}$
450: and so is a finite number of points in $\C$. Choosing a polynomial $p$ with
451: high order zeros at these points such that $p(0)=1$, $p.s$ is a holomorphic
452: section with the same class as $s$ in $(\pi_*\L)|_{\{0\}}$ since $p-1\in(t)$.
453: Then $[p.s]$ defines the required class in $H^0_\X(\L)\big/tH^0_\X(\L)$.)
454: \end{proof}
455:
456: Denote the weight of the induced $\C^\times$-action on
457: $(\pi_*\L^K)|_{\{0\}}=H^0_\X(\L^K)\big/tH^0_\X(\L^K)$ by $w(Kr)$.
458: (We enumerate by $k:=Kr$ since $(\X,\L^K)$ is a test configuration for $(X,L^{Kr})$.
459: The confused reader may set $r=1$ temporarily.) For $K\gg0,\
460: (\pi_*\L^K)|_{\{0\}}$ is $H^0_{\X_0}(\L^K)$ and
461: $w(k)=w(Kr)$ is a polynomial of degree $n+1$ in $k=Kr$ by the equivariant
462: Riemann-Roch theorem. (It is important that we do \emph{not} modify
463: $w(k)$ to be this polynomial for small $k$, so for instance $w(r)$ really
464: is the weight on
465: $(\pi_*\L)|_{\{0\}}$.) To make the $\C^\times$-action special
466: linear on $(\pi_*\L)|_{\{0\}}$ we first pull back the family by the cover
467: $\C\to\C,\ t\mapsto t^{r\P(r)}$, making the action of weight $r\P(r)w(r)$.
468: Composing with the trivial action which scales the $\L$-fibres with weight
469: $-rw(r)$ scales $\Lambda^{\max}(\pi_*\L)|_{\{0\}}$ with weight
470: $-r\P(r)w(r)$, cancelling out the previous action. (The extra factor of $r$
471: is to make the formula (\ref{normalised}) nicer, and is natural if $(X,\L)$
472: is the $r$th twist of a test configuration for $(X,L)$.)
473: Since this new action acts with zero total weight on $S^K(\pi_*\L)|_{\{0\}}$,
474: it acts on the line $\O_\G(1)|_{\{\X_0\}}=
475: \Lambda^{\max}H^0_{\X_0}(\L^{Kr})\otimes
476: \left(\Lambda^{\max}S^K(\pi_*\L)|_{\{0\}}\right)^*$ (compare (\ref{eq:line})
477: which was for $(X,L^r)$ with no higher cohomology of $L^r$) with
478: \emph{normalised weight}
479: \begin{equation} \label{normalised}
480: \tilde{w}_{r,Kr}=\tilde{w}_{r,k}=w(k)r\P(r)-w(r)k\P(k), \qquad k:=Kr.
481: \end{equation}
482: The normalised weight is a polynomial $\sum_{i=0}^{n+1}e_i(r)k^i$
483: of degree $n+1$ in $k$ for $k\gg0$,
484: with coefficients which are also polynomial of degree $n+1$ in $r$ for $r\gg0$:
485: $e_i(r)=\sum_{j=0}^{n+1}e_{i,j}r^j$ for $r\gg0$. The normalisation means
486: that the coefficient of
487: $(rk)^{n+1}$ vanishes: $e_{n+1,n+1}=0$. The Hilbert-Mumford
488: criterion relates this weight to stability as follows.
489:
490: \begin{thm} \label{thm:instability}
491: A polarised variety $(X,L)$ is \emph{stable} if and only if
492: $$
493: \tilde{w}_{r,k} \succ 0 \quad \forall \text{ nontrivial
494: test configurations for } (X,L^r),
495: $$
496: where $\succ$ and $\forall$ mean the following for the
497: different notions of stability:
498: \begin{itemize}
499: \item Hilbert stable with respect to $r$: for any nontrivial
500: test configuration for $(X,L^r),\ \tilde{w}_{r,k}>0$ for all $k\gg0$,
501: \item Asymptotically Hilbert stable: for all $r\gg0$, any nontrivial
502: test configuration for $(X,L^r)$ has $\tilde{w}_{r,k}>0$ for all $k\gg0$,
503: \item Chow stable with respect to $r$: for any nontrivial
504: test configuration for $(X,L^r)$, the leading $k^{n+1}$
505: coefficient $e_{n+1}(r)$ of $\tilde{w}_{r,k}$ is positive: $e_{n+1}(r)>0$,
506: \item Asymptotically Chow stable: for all $r\gg0$, any nontrivial
507: test configuration for $(X,L^r)$ has $e_{n+1}(r)>0$,
508: \item K-stable: for all $r\gg0$, any nontrivial test configuration for $(X,L^r)$
509: has leading
510: coefficient $e_{n,n+1}$ of $e_{n+1}(r)$ (the \emph{Donaldson-Futaki invariant}
511: of the test configuration) positive: $e_{n,n+1}>0$.
512: \end{itemize}
513: Furthermore the result holds if we replace ``stable" with
514: ``semistable" and strict inequalities with non strict inequalities throughout.
515:
516: Finally $(X,L)$ is \emph{polystable} if it is semistable and any destabilising
517: test configuration (i.e.\ one which is not strictly stable) is a product
518: configuration.
519: \end{thm}
520:
521: The increasing number of test configurations that have to be tested in the
522: second, fourth and fifth definitions
523: as $r\to\infty$ currently prevent us from adding K-stability to the left
524: of the following consequences of Theorem \ref{thm:instability}. \\
525:
526: Asymptotically Chow stable $\Rightarrow$ Asymptotically Hilbert stable
527: $\Rightarrow$ Asymptotically Hilbert semistable $\Rightarrow$ Asymptotically
528: Chow semistable $\Rightarrow$ K-semistable. \\
529:
530: The relation between K-stability and asymptotic Chow stability is analogous
531: to the relation between slope stability and Gieseker stability for vector
532: bundles, and a geometric criterion for asymptotic Chow stability would show
533: it was implied by K-stability; we only have a necessary condition (Theorem
534: \ref{chow}).
535:
536: The fact that Chow stability is controlled by the coefficient $e_{n+1}(r)$
537: is due to Mumford \cite{Mu}. The definition of K-stability above is due to
538: Donaldson, adapting Tian's original differential-geometric definition to
539: allow nonnormal central fibres $\X_0$ (though what
540: is called properly semistable in \cite{Ti2} and stable in
541: \cite{Do2} is what we call K-polystable). Tian \cite{Ti1}
542: defines a line bundle (the ``CM polarisation") on $\Hilb$ such that K-stability
543: is exactly stability in the sense of the Hilbert-Mumford criterion for this
544: line bundle \cite{PT}. K-stability is probably not a bona
545: fide GIT notion: Tian's polarisation may not be ample, and
546: the number of test configurations increases as $r$ tends to infinity.
547: However it is K-polystability that is conjecturally related to
548: the existence of constant scalar curvature K\"ahler metrics; we apply our
549: methods to these in \cite{RT}.
550:
551: The definition of polystability says that any destabilising test configuration
552: comes from a $\C^\times$-action on $X$ with the appropriate weight 0 (i.e.\
553: Donaldson-Futaki invariant 0 in the K-polystability case). This corresponds to an
554: orbit in Hilb which is closed in the semistable points but with possibly
555: higher dimensional stabilisers; equivalently the orbit in the total space
556: of the dual of the polarising line bundle over Hilb is closed. There seems
557: to be no universally accepted name for this; we use polystability by analogy
558: with bundles.
559:
560: A test configuration $(\X,\L)$ for $(X,L^r)$ can be twisted to get another,
561: $(\X,\L^K)$, for $(X,L^{Kr})$; for $K\gg0$ this will have no higher cohomology.
562: Since the definition of K-stability is unchanged if
563: $L$ is replaced by some power, for this notion we can allow $\L$ to
564: be an ample $\Q$\,-line bundle in the definition of a test configuration.
565:
566: Letting $F:=e_{n,n+1}$ denote the Donaldson-Futaki invariant and writing
567: the unnormalised weight $w(k)$ as $b_0k^{n+1}+b_1k^n+\ldots\,$, we see that
568: \begin{equation} \label{Futaki}
569: F=b_0a_1-b_1a_0,
570: \end{equation}
571: and $-a_0^{-2}F$ is the coefficient of $k^{-1}$ in the expansion of
572: $w(k)/k\P(k)$. When the central fibre $\X_0$ is smooth,
573: $F=\frac{a_0}{4}\nu$, where $\nu$ is the usual Calabi-Futaki invariant
574: of $c_1(L)$ and the vector field generated by the $S^1$-action
575: \cite{Do2}. \\
576:
577: Without loss of generality we now take $r=1$.
578: An arbitrary test configuration $(\Y,\O_\Y(1))$ for $(X,L)$ is, by definition,
579: $\C^\times$-isomorphic to the trivial test configuration $(X\times\C,L)$
580: away from the central fibre. It is therefore $\C^\times$-birational to
581: $(X\times\C,L)$ and so is dominated by a blow up $(\X,\L)$ of $X\times\C$
582: in a $\C^\times$-invariant ideal $I$ supported on (a thickening
583: of) the central fibre $X\times\{0\}$:
584: \begin{eqnarray}
585: (\X,\L)\ =&(\B_I(X\times\C),L(-E))&\Rt{\phi}\ (\Y,\O_\Y(1)) \nonumber \\
586: &\downarrow p \label{any} \\ &(X\times\C,L)\,. \nonumber
587: \end{eqnarray}
588: Here we use the canonical $\C^\times$-action
589: on $\B_I(X\times\C)$ inherited from that on $I$, and its linearisation on
590: the line bundle $\L:=L-E=p^*(L\otimes I)$, where $E$ denotes the exceptional
591: divisor. $L-E=\phi^*\O_\Y(1)$ and the horizontal arrow in (\ref{any}) is
592: an equivariant map of equivariant
593: polarisations (although $L-E$ may be only semi-ample), whereas
594: $p$ does not respect the polarisation.
595:
596: Mumford (\cite{Mu} section 3 of the proof of Theorem 2.9) essentially shows
597: that any test configuration is of this form, where $I=I_r$ is of the form
598: \begin{equation} \label{ideals}
599: I_r=\I_0+t\I_1+t^2\I_2+\ldots+t^{r-1}\I_{r-1}+(t^r)\ \subset\ \O_X\otimes\C[t].
600: \end{equation}
601: Here the ideals $\I_0\subseteq\I_1\subseteq\ldots\subseteq\I_r
602: \subseteq\O_X$ correspond to subschemes $Z_0\supseteq Z_1\supseteq\ldots\supseteq
603: Z_{r-1}$ of $(X,L)$, so $I_r$ is $\C^\times$-invariant and corresponds to
604: a subscheme of $X\times\C$ supported on (the $r$-fold
605: thickening of) the central fibre $X\times\{0\}$.
606:
607: This can be proved by writing the $\C^\times$-invariant ideal $I$ as a sum
608: of weight spaces: defining $\I_j$ in terms of the weight-$j$ piece $t^j\I_j$
609: we get the weight space decomposition (\ref{ideals}). Or, embedding a test
610: configuration in some $\PP^N\times\C$, Mumford's result applies directly.
611:
612: For example, given a $\C^\times$-action on $X$ with ``repulsive fixed point
613: set" $Z$ (that part of the fixed point set with negative weight spaces in
614: its normal cone), there is a blow up of $X\times\C$ supported on $Z\times\{0\}$ in which the proper transform of $X\times\{0\}$ can be blown down to give
615: back $X\times\C$ but with a nontrivial $\C^\times$-action.
616:
617: Since $(\X,\L)$ is a \emph{semi} test configuration (and
618: because later we will want to replace an ample $(X,L)$ with a semi-ample
619: $(\widehat X,p^*L)$ for some resolution of singularities $p$) we will
620: prove many results in the generality of semi test configurations.
621:
622: General test configurations (\ref{any}) will be studied in Section \ref{testconfigs};
623: next we look at the simplest case (beyond product configurations) of $r=1,\
624: I=\I_Z+(t)$ in (\ref{any}): the \emph{deformation of $X$ to the normal cone
625: of $Z$}.
626:
627:
628: \section{Deformation to the normal cone and K-slope stability} \label{Kslope}
629:
630: Given any proper subscheme $Z\subset X$ we get a canonical test
631: configuration $\X$, the blow up of $X\times\C$ along $Z\times\{0\}$ with
632: exceptional divisor $P$. This deformation to the normal cone has central
633: fibre $\X_0=\widehat X\cup_E P$, where $\widehat X$ is
634: the blow up of $X$ along $Z$ with exceptional divisor $E$. When $Z$ and
635: $X$ are both smooth $P=\PP(\nu\oplus\underline\C)$ is the projective
636: completion of the normal bundle $\nu$ of $Z$ in $X$, glued along
637: $E=\PP(\nu)$ to the blow up of $X$. For more on the deformation to the
638: normal cone see \cite{Fu}; for a diagram see Section \ref{testconfigs}.
639:
640: Let $\pi$ be the composite of the projections
641: $\X\to X\times\C\to X$. For $L$ an ample line bundle on $X$ and $c$ a positive
642: rational number let $\L_c$ be the $\Q\,$-line bundle $\pi^*L-cP$. This
643: restricts to $L$ on the general fibre of $\X\rightarrow\C$, and is ample
644: for $c$ sufficiently small.
645:
646: \begin{prop} \label{ample}
647: Fix $L$ an ample (respectively semi-ample) line bundle on $X$.
648: For $c\in(0,\epsilon(Z))\cap\Q$ (\ref{def:seshadri}), the $\Q$-line bundle
649: $\L_c$ is ample (semi-ample) on
650: $\X$. If $\epsilon(Z)\in\Q$ then $\L_{\epsilon(Z)}$ is nef. If in
651: addition the global sections of $L^k\otimes\I_{\!Z}^{\epsilon(Z)k}$
652: saturate (see Section \ref{notation}) for $k\gg0$ then $\L_{\epsilon(Z)}$
653: is semi-ample.
654: \end{prop}
655:
656: \begin{proof}
657: Note that if $k,kc\in\mathbb N$ and $L^k$ is globally generated
658: and the sections of $L^k\otimes\I_{\!Z}^{ck}$ saturate, then $\L_c^k$
659: is globally generated on $X\times\C$ by the
660: sections $\pi^*H^0(L^k\otimes\I_{\!Z}^{ck})+t^{ck}\pi^*H^0(L^k)$. Algebraically
661: this is the statement that on $X\times\C$, $\I_{\!Z}^{ck}+(t^{ck})$ saturates
662: $(\I_Z+(t))^{ck}$. Geometrically it says that the global sections saturating
663: $L^k\otimes\I_{\!Z}^{ck}$ generate $L^k-ckP$ on $\X$ away from the zero
664: section $\PP(\underline\C\to Z)\cong Z$ of the cone $P\to Z$, over which
665: the sections $t^{ck}H^0(L^k)$ generate.
666:
667: Putting $c=\epsilon$ now gives the third claim. For the first
668: two we claim that we may assume that $L$ is ample by replacing $(X,L)$ by
669: $(Y,\O_Y(1)):=\Proj\bigoplus_kH^0(L^k)$ if necessary. For $k$ sufficiently
670: large that $L^k\otimes\I_{\!Z}$ is globally
671: generated, $H^0(L^k\otimes\I_{\!Z})\subset H^0(L^k)=H^0(\O_Y(k))$
672: generates a subsheaf $\O_Y(k)\otimes\I_{Z_0}\subset\O_Y(k)$ and so a subscheme
673: $Z_0\subset Y$ such the pullback of $\I_{Z_0}$ to $X$ is $\I_Z$. Thus it
674: is sufficient to prove the result for $Z_0\subset(Y,\O_Y(1))$ and then pullback
675: to get the result for $Z\subset(X,L)$.
676:
677: On $\widehat X$, $L-cE$ is in the ample cone for small $c$ and
678: on its boundary for $c=\epsilon(Z)$, so by Kleiman \cite{Kl}
679: $L-cE$ is ample for rational $c<\epsilon(Z)$. So $L^k-ckE$ is globally generated
680: for $k\gg0$; equivalently the global sections of $L^k\otimes\I_{\!Z}^{ck}$
681: saturate $\I_{\!Z}^{ck}$ (for $k$ sufficiently large that the pushdown of
682: $\O(-ckE)$ to $X$ is $\I_{\!Z}^{ck}$). $L^k$ is also globally generated so
683: again this implies that $\L_c^k$ is globally generated. Thus $\L_c$
684: is nef for $c\in(0,\epsilon(Z))$, but it is ample for small $c$ since $\pi^*L$
685: is ample. Thus by Kleiman again, $\L_c$ is actually ample and $\L_{\epsilon(Z)}$
686: is nef.
687: \end{proof}
688:
689: The obvious $\C^\times$-action on $X\times\C$ (acting trivially on the
690: $X$ factor) lifts to an action on the deformation to the normal cone
691: $(\X,\L_c)$ which, on the central fibre $\X_0=\widehat
692: X\cup_EP$, is trivial on $\widehat X$. When $Z$ and $X$
693: are both smooth, $\lambda\in\C^\times$ acts on
694: $P=\PP(\nu\oplus\underline \C)$ as $\operatorname{diag}(1,\lambda)$.
695:
696: \begin{thm} \label{degen}
697: Fix $L$ ample and $c\in(0,\epsilon(Z))\cap\Q$. Then $(\X,\L_c)$ is a flat
698: family of polarised schemes, and, for all $k\gg0,\ ck\in\mathbb N$, the
699: total weight of the induced action on $H^0(\X_0,\L_c^k)$ is
700: $$
701: w(k)=-\sum_{j=1}^{ck} j\,h^0\!\left(L^k\otimes(\I_{\!Z}^{ck-j}/
702: \I_{\!Z}^{ck-j+1})\right)=\sum_{j=1}^{ck} \chi(L^k\otimes
703: \I_{\!Z}^j)-ckh^0(L^k).
704: $$
705: \end{thm}
706:
707: \begin{proof} From the definition of the blow up of a subscheme, $\X=\Proj
708: \bigoplus_{k\ge 0} \S_k$, where
709: \begin{eqnarray} \nonumber
710: \S_k &=& (\I_{\!Z \times \{0\}})^k = (\I_Z+(t))^k \\
711: &=& \bigoplus_{j=0}^{k-1} t^j\I_{\!Z}^{k-j}\ \oplus\
712: t^k\C[t]\O_X\ \subset\ \C[t]\otimes\O_X. \label{S}
713: \end{eqnarray}
714: It follows that for all $k\gg0$, the pushdown of $-kP$ to $X\times\C$ is
715: $\bigoplus_{j=0}^{k-1}t^j\I_{\!Z}^{k-j}\oplus t^k\C[t]\O_X$, with the higher
716: pushdowns zero. By the ampleness of $L-cP$ (\ref{ample}), for $k\gg0$ we
717: have the vanishing of
718: $$
719: H^i_\X((L-cP)^k) = H^i_{X\times\C}(\pi_*(L^k-ckP)) =
720: \bigoplus_{j=0}^{ck-1} t^j H^i_X(L^k\otimes\I_{\!Z}^{ck-j})\oplus
721: t^{ck}\C[t]H^i_X(L^k),
722: $$
723: for $i>0$. In particular then,
724: \begin{equation} \label{vanish}
725: H^i_X(L^k\otimes\I_{\!Z}^{ck-j})=0\quad\text{for}\ j=0,\ldots,ck, \quad
726: i>0.
727: \end{equation}
728:
729: Now
730: $$
731: \X_0=\Proj\bigoplus_{k\ge0}\S_k/t\S_k,
732: $$
733: where by (\ref{S}),
734: \begin{equation} \label{St}
735: \S_k/t\S_k=\I_{\!Z}^k\ \oplus\ t\!\left(\!\I_{\!Z}^{k-1}/\I_{\!Z}^k\right)\
736: \oplus\,\ldots\,\oplus\ t^k\big(\O_X/\I_Z\big).
737: \end{equation}
738: Replacing $\I_Z$ by $\I^c_{\!Z}$ does not change the blow up (just the corresponding
739: exceptional line bundle) so that, for $k\gg0$,
740: \begin{equation} \label{wspace}
741: H^0_{\X_0}(\L_c^k)=H^0_X(L^k\otimes\I_{\!Z}^{ck})\ \oplus\,\bigoplus_{j=1}^{ck}
742: t^j H^0_X(L^k\otimes(\I_{\!Z}^{ck-j}/\I_{\!Z}^{ck-j+1})).
743: \end{equation}
744: The vanishing (\ref{vanish}) of
745: $H^1(L^k\otimes\I_{\!Z}^{ck-j+1})$ for $1\le j\le ck$ means that the dimension
746: of this is
747: $$
748: h^0_{\X_0}(\L_c^k)=
749: h^0_X(L^k\otimes\I_{\!Z}^{ck})+\sum_{j=1}^{ck}\left(h^0_X(L^k\otimes \I_{\!Z}^{ck-j})-
750: h^0_X(L^k\otimes\I_{\!Z}^{ck-j+1})\right),
751: $$
752: which is $h^0_X(L^k)$. By (\cite{Ha} Theorem III.9.9) this proves flatness
753: of the family.
754:
755: Now (\ref{wspace}) is the weight space decomposition with respect to the
756: $\C^\times$-action: $\C^\times$ acts on $t$ with weight $-1$ and trivially
757: on $X$ and so on $\I_Z$. Therefore
758: \begin{eqnarray}
759: w(k) &=& -\sum_{j=1}^{ck} jh^0\left(L^k\otimes(\I_{\!Z}^{ck-j}/
760: \I_{\!Z}^{ck-j+1})\right) \nonumber \\
761: &=&-\sum_{j=1}^{ck} j\left(h^0(L^k\otimes \I_{\!Z}^{ck-j})-
762: h^0(L^k\otimes \I_{\!Z}^{ck-j+1})\right) \label{noco} \\
763: &=& \sum_{j=1}^{ck} (ck-j+1) h^0(L^k\otimes\I_{\!Z}^j)-\sum_{j=0}^{ck-1}
764: (ck-j)h^0(L^k\otimes \I_{\!Z}^j) \nonumber \\
765: &=& \sum_{j=1}^{ck} h^0(L^k\otimes \I_{\!Z}^j)-ckh^0(L^k)
766: \ =\ \sum_{j=1}^{ck} \chi(L^k\otimes \I_{\!Z}^j)-ckh^0(L^k), \nonumber
767: \end{eqnarray}
768: where the second and last equalities follow from the vanishing (\ref{vanish})
769: of $H^i(L^k\otimes\I_{\!Z}^{ck-j+1})$ for $1\le j\le ck$ and $i>0$.
770: \end{proof}
771:
772: We may only take $c=\epsilon(Z)$ if the global sections
773: of $L^k\otimes\I_{\!Z}^{\epsilon(Z)k}$ saturate for $k\gg0$. In this case
774: $\L_c$ is semi-ample, pulled back from a contraction $p$ from $(\X,\L_c)\to\C$
775: to $\Proj\bigoplus_{k\gg0}H^0_\X(\L_c^k)$, which we call $(\Y,\O_\Y(1))\to\C$.
776: By construction, $p^*\colon H^0_\Y(\O_\Y(k))\to H^0_\X(\L_c^k)$ is then
777: an isomorphism. By Lemma 2.13 of \cite{Mu}, $\Y\to\C$ is also a flat family,
778: so $(\Y,\O_\Y(k))\to\C$ is a test configuration for $k\gg0$.
779:
780: \begin{thm} \label{ss}
781: Let $Z$ be a proper subscheme of an irreducible polarised algebraic
782: variety $(X,L)$, and suppose that $c=\epsilon(Z)\in\Q$ and the global sections
783: of $L^k\otimes\I_{\!Z}^{\epsilon(Z)k}$ saturate for $k\gg0,\ ck\in\mathbb
784: N$. Letting $(\Y,\O_\Y(1))$ be the contraction of $(\X,\L_c)$ constructed
785: above, the weight of the action on $\Lambda^{\max}H^0_{\Y_0}(\O_\Y(k))$ is
786: $$
787: w(k)=\sum_{j=1}^{ck}\chi(L^k\otimes\I_{\!Z}^j)-ckh^0(L^k)=
788: \sum_{j=1}^{ck}\chi(L^k\otimes\I_{\!Z}^j)-ckh^0(L^k)+O(k^{n-1}).
789: $$
790: \end{thm}
791:
792: \begin{proof}
793: By (\ref{S}), for $k\gg0$,
794: $$
795: H^0_\X(\L_c^k)=H^0_{X\times\C}(L^k\otimes\S_{ck})\cong
796: \bigoplus_{j=0}^{ck-1}t^jH^0_X(L^k\otimes\I_{\!Z}^{k-j})\ \oplus\ t^{ck}\C[t]H^0_X(L^k),
797: $$
798: yielding
799: \begin{equation} \label{qt}
800: \frac{H^0_\X(\L_c^k)}{tH^0_\X(\L_c^k)}\cong H^0_X(L^k\otimes\I_{\!Z}^{ck})\
801: \oplus\ \bigoplus_{j=1}^{ck}\,t^j\frac{H^0_X(L^k\otimes\I_{\!Z}^{ck-j})}
802: {H^0_X(L^k\otimes\I_{\!Z}^{ck-j+1})}\,.
803: \end{equation}
804: By the isomorphism $p^*\colon H^0_\Y(\O_\Y(k))\to H^0_\X(\L_c^k)$ this is
805: the weight space decomposition of $H^0_\Y(\O_\Y(k))\big/tH^0_\Y(\O_\Y(k))$,
806: which by flatness of $\Y\to\C$ and ampleness of $\O_\Y(1)$ is
807: $H^0_{\Y_0}(\O_\Y(k))$. So the total weight is
808: $$
809: -\sum_{j=1}^{ck} j\left(h^0(L^k\otimes \I_{\!Z}^{ck-j})-
810: h^0(L^k\otimes \I_{\!Z}^{ck-j+1})\right).
811: $$
812: Just as in (\ref{noco}) this is $\sum_{j=1}^{ck}h^0(L^k\otimes\I_{\!Z}^j)-ckh^0(L^k)$.
813: Then replacing $h^0$ by $\chi$ gives an error bounded by $\sum_{j=1}^{ck}
814: h^{\ge1}_X(L^k\otimes\I_{\!Z}^{ck-j+1})$, where $h^{\ge1}:=\sum_{i=1}^nh^i$.
815: So the result follows from Lemma \ref{error} below.
816: \end{proof}
817:
818: \begin{lem} \label{error}
819: Fix $Z\subset(X,L)$ and $c\in(0,\epsilon(Z))\cap\Q$, or $c\in(0,\epsilon(Z)]
820: \cap\Q$ if $\epsilon(Z)\in\Q$ and the global sections of $L^k\otimes
821: \I_{\!Z}^{\epsilon(Z)k}$ saturate for $k\gg0$. Then
822: $$
823: \sum_{j=0}^{ck}h^{\ge1}_X(L^k\otimes\I_{\!Z}^j)=O(k^{n-1}).
824: $$
825: \end{lem}
826:
827: \begin{proof}
828: Fix $\varepsilon\in(0,c)\cap\Q$. Then $\L_{c-\varepsilon}$ is ample on
829: $\X$, while $\L_c$ is generated by its global sections and so nef. So we
830: can apply Fujita vanishing (\cite{La} Theorem
831: 1.4.35) to give $N\gg0$ such that $\L_{c-\varepsilon}^N\otimes\L_c^p$ has
832: no higher cohomology for any $p\ge0$. For $k>N$, setting $p=k-N$ shows
833: that $\L_{c-\varepsilon N/k}^k$ has no higher cohomology. So for $k\gg0$
834: we have
835: \begin{eqnarray} \nonumber
836: 0&=&H^i_\X((L-(c-\varepsilon N/k)P)^k)=H^i_{X\times\C}
837: (\pi_*(L^k-(ck-\varepsilon N)P)) \\
838: &=&\bigoplus_{j=0}^{ck-\varepsilon N-1}t^j H^i_X(L^k\otimes
839: \I_{\!Z}^{ck-\varepsilon N-j})\
840: \oplus\ t^{ck-\varepsilon N}\C[t]H^i_X(L^k), \label{dec}
841: \end{eqnarray}
842: for $i>0$. That is, $h^{\ge1}(L^k\otimes\I_{\!Z}^{ck-\varepsilon N-j})=0$ for
843: $j=0,\ldots,ck-\varepsilon N$. So the sum becomes
844: $$
845: \sum_{j=0}^{ck}h^{\ge1}_X(L^k\otimes\I_{\!Z}^j)=
846: \sum_{j=0}^{\varepsilon N-1}h^{\ge1}_X(L^k\otimes\I_{\!Z}^{ck-j})=
847: \sum_{j=0}^{\varepsilon N-1}h^{\ge1}_{\widehat X}(L^k-(ck-j)E),
848: $$
849: where we have taken $k$ sufficiently large that the pushdown of $\O_{\widehat
850: X}(-iE)$ is $\I_{\!Z}^i$ (and higher pushdowns are zero) for $i>ck-\varepsilon
851: N$. A corollary of Fujita vanishing is that $h^i(\curly F(kD))=O(k^{n-i})$
852: for any coherent sheaf $\curly F$ and nef divisor $D$ (\cite{La} Theorem
853: 1.4.40). Applying this on $\widehat X$ in turn to $\curly F=
854: \O,\O(E),\ldots,\O((\varepsilon N-1)E)$ and $D=L-cE$ shows that each
855: $h^{\ge1}_{\widehat X}(L^k-(ck-j+1)E)=O(k^{n-1})$. Since
856: $N$ is fixed, then, we get a similar bound on the whole sum.
857: \end{proof}
858:
859: \begin{cor} \label{weight}
860: For $c\in(0,\epsilon(Z)]\cap\Q$ define
861: $$
862: \tilde{w}_{r,k}(c)=r\chi(L^r)\sum_{j=1}^{ck}h^0(L^k\otimes\I_{\!Z}^j)
863: \,-k\chi(L^k)w(r)-ck\chi(L^k)r\chi(L^r),
864: $$
865: which, for $r$ sufficiently large \emph{for fixed $Z\subset X$,} is
866: $$
867: \tilde{w}_{r,k}(c)=r\chi(L^r)\sum_{j=1}^{ck}h^0(L^k\otimes
868: \I_{\!Z}^j) - k\chi(L^k)\sum_{j=1}^{cr}h^0(L^r\otimes\I_{\!Z}^j).
869: $$
870: For $cr\in\mathbb Z$, $L^r$ globally generated and $L^r\otimes\I_{\!Z}^{cr}$
871: saturated by its sections, $(X,L^r)$ is unstable if $\tilde{w}_{r,k}(c)\preceq0$,
872: where $\preceq$ is defined as in Theorem \ref{thm:instability}, depending
873: on the type of stability. We say that $(X,L^r)$ is destabilised by $Z$. Similarly
874: for strict instability and $\prec$.
875: \end{cor}
876:
877: \begin{proof}
878: The conditions on $r$ are just to ensure that $(\X,\L_c^r)$ (or $(\Y,\O_\Y(r))$
879: if $c=\epsilon$) is a genuine test configuration; for K-stability we are
880: free to twist by higher $r$ and use Proposition \ref{ample}
881: to remove these conditions.
882:
883: The result is a direct consequence of Theorems \ref{degen} and \ref{ss} and
884: the identity $\tilde{w}_{r,k}=w(k)r\chi(L^r)-w(r)k\chi(L^k)$ (\ref{normalised}).
885: \end{proof}
886:
887: Recall from the introduction the definition of the $a_i$ and $a_i(x)$ (\ref{HS}).
888: Choose $p$ so that $\mathbf R\pi_*\O(-jE)=\I_{\!Z}^j$ for $j\ge p$,
889: then for $k,xk\in\mathbb N,\,k\ge p/x$,
890: \begin{equation} \label{definitionofa_i}
891: \chi\_{\widehat X}(L^k-xkE)=\chi\_X(L^k\otimes\I_{\!Z}^{xk})=a_0(x)k^n
892: + a_1(x)k^{n-1} + \ldots +a_n(x).
893: \end{equation}
894: But by Riemann-Roch, $P(k,j):=\chi\_{\widehat{X}}(L^k-jE)$ is a polynomial
895: of total degree $n$. Writing $P=P_0+\ldots+P_n$ where $P_i$ is
896: homogeneous of degree $n-i,\ a_i(x)=P_i(1,x)$ is therefore a degree
897: $n-i$ polynomial in $x$ which can be extended to
898: all real $x$.
899:
900: \begin{prop} \label{integrals}
901: For $X$ irreducible the weights $\sum_{j=1}^{ck}\chi(L^k\otimes \I_{\!Z}^j)-
902: ckh^0(L^k)$ of Theorems \ref{ss} and \ref{degen} can be written
903: $$
904: w(k)=\left(\int_0^ca_0(x)dx-ca_0\right)k^{n+1}+\left(\int_0^c\!a_1(x)
905: +\frac{a_0'(x)}2\,dx-ca_1\right)k^n+O(k^{n-1}).
906: $$
907: \end{prop}
908:
909: \begin{proof} Using (\ref{definitionofa_i}) for $j>p$ we split up
910: $\sum_{j=1}^{ck}\chi(L^k\otimes \I_{\!Z}^j)$ as
911: \begin{eqnarray} \nonumber
912: && \sum_{j=1}^p\chi(L^k\otimes \I_{\!Z}^j)
913: +\sum_{j=p+1}^{ck}\chi(L^k\otimes \I_{\!Z}^j) \\
914: &=&p\,\chi(L^k)-\sum_{j=1}^p\chi(L^k\otimes\O_{jZ})+\sum_{j=p+1}^{ck}(a_0(j/k)k^n
915: + a_1(j/k)k^{n-1}+\ldots) \nonumber \\
916: &=&\int_0^{p/k}\!a_0(x)dx\,k^{n+1} + \sum_{j=p+1}^{ck}(a_0(j/k)k^n
917: + a_1(j/k)k^{n-1})+O(k^{n-1}), \label{prox}
918: \end{eqnarray}
919: using the fact that $jZ$ has dimension $\le n-1$ (since $X$ is irreducible)
920: and $a_0(x)$ is a polynomial with $a_0(0)=a_0$.
921:
922: For a smooth function $f$, as $k\to\infty$ with $ck\in\mathbb
923: N$, the trapezium rule gives \cite{Hi}
924: \begin{equation} \label{easyeuler}
925: \sum_{j=1}^{ck}f(j/k)=\int_0^c\left(kf(x)+\frac{f'(x)}2\right)dx+O(k^{-1}).
926: \end{equation}
927: This can be proved by Taylor's theorem, or directly for polynomials by noting
928: that for $f(x)=x^m$ we have the identity
929: $$
930: \sum_{j=1}^{ck} j^m = \frac{(ck)^{m+1}}{m+1} + \frac{(ck)^m}2+ O(k^{m-1}).
931: $$
932: Applying this to $f(x)=a_i(x),\ i=1,2$, (\ref{prox}) now approximates
933: $\sum_{j=1}^{ck}\chi(L^k\otimes \I_{\!Z}^j)$ by
934: \begin{multline*} \hspace{-4mm}
935: \int_0^{p/k}\!\!a_0(x)dx\,k^{n+1}
936: +\left(\int_{p/k}^cka_0(x)+\frac{a_0'(x)}2dx\right)k^n+
937: \left(\int_{p/k}^cka_1(x)dx\right)k^{n-1}+ O(k^{n-1}) \\
938: =\left(\int_0^c a_0(x) dx \right) k^{n+1} + \left(\int_0^c
939: a_1(x) + \frac{a_0'(x)}{2}dx \right) k^n + O(k^{n-1}),
940: \end{multline*}
941: since $\int_0^{p/k}\!a_1(x)+\frac{a_0'(x)}{2} dx\,k^n=O(k^{n-1})$.
942: Subtracting $ckh^0(L^k)=ca_0\,k^{n+1}+ca_1\,k^n+O(k^{n-1})$ gives the result.
943: \end{proof}
944:
945: Note that by Riemann-Roch on $\widehat X$,
946: $n!a_0(x)=(L-xE)^n>0$ for $x\in(0,\epsilon(Z))$ by the ampleness of $L-xE$,
947: so for $c\in(0,\epsilon(Z)]$, $\int_0^ca_0(x)dx>0$. Therefore we can define
948: the \emph{slope} (or K-slope) of $\I_Z$ (\ref{def:slope}) by
949: $$
950: \mu_c(\I_Z)=\mu_c(\I_Z,L) = \frac{\int_0^c a_1(x) + \frac{a'_0
951: (x)}{2} dx} {\int_0^c a_0(x)
952: dx}=\frac{\int_0^ca_1(x)dx}{\int_0^ca_0(x)dx}+
953: \frac{a_0(c)-a_0}{2\int_0^ca_0(x)dx}\,,
954: $$
955: and that of $X$ (\ref{def:X}),
956: $$
957: \mu(X) = \frac{a_1}{a_0} \qquad \left(=-\frac n2\cdot\frac{K_X.L^{n-1}}{L^n}
958: \ \text{ for $X$ smooth}\right).
959: $$
960: This gives the following obvious definition of K-slope stability, which,
961: like K-stability, is independent of replacing
962: $L$ by $L^r$ (on replacing $c,\,x$ and $\epsilon$ by $rc,\,rx$ and
963: $r\epsilon$).
964:
965: \begin{defn} \label{defstab}
966: We say that $(X,L)$ is slope (semi/poly)stable if for every proper
967: subscheme $Z$ of $X$, the following holds:
968: \begin{itemize}
969: \item \textbf{\emph{slope semistability}}: $\mu_c(\I_Z,L)\le\mu(X)$ for all
970: $c\in(0,\epsilon(Z)]$,
971: \item \textbf{\emph{slope stability}}: \hspace{6mm} $\mu_c(\I_Z,L)<\mu(X)$
972: for all $c\in(0,\epsilon(Z))$, and also for $c=\epsilon(Z)$ if
973: $\epsilon(Z)\in\Q$ and global sections of $L^k\otimes\I_{\!Z}^{\epsilon(Z)k}$
974: saturate for $k\gg0$,
975: \item \textbf{\emph{slope polystability}}: $\ (X,L)$ is slope
976: semistable, and if $(Z,c)$ is any pair such that
977: $\mu(\I_Z,c)=\mu(X)$, then $c=\epsilon(Z)$ is rational and, on the
978: deformation to the normal cone of $Z$, $\L_c$ is pulled back from a
979: product test configuration $(\Y,\O_\Y(1))\cong(X\times\C,L)$.
980: \end{itemize}
981: \end{defn}
982:
983: \begin{thm}
984: \label{thm:kstableslopestable}
985: If a polarised variety $(X,L)$ is K-stable then it is slope stable.
986: If it is K-polystable (respectively K-semistable) then it is slope polystable
987: (semistable).
988: \end{thm}
989:
990: \begin{proof}
991: From the deformation to the normal cone of $Z$ we get test configurations
992: $(\X,\L_c^k)$ (for $c<\epsilon(Z),\ k\gg0$), or $(\Y,\O_\Y(k))$ (for $c=\epsilon(Z)$
993: if $\epsilon(Z)\in\Q$ and the global sections of $L^k\otimes\I_{\!Z}^{\epsilon(Z)k}$
994: saturate). Its total weight $w(k)$ is given by Theorems \ref{degen}
995: and \ref{ss} respectively. Writing $w(k)=b_0k^{n+1}+b_1k^n + O(k^{n-1})$,
996: its Donaldson-Futaki invariant (\ref{Futaki}) is, by Proposition \ref{integrals},
997: \begin{eqnarray} \nonumber
998: F(Z) &=& b_0a_1-b_1a_0=\left(\int_0^c\!a_0(x)dx-ca_0\right)\!a_1-\left(
999: \int_0^c\!a_1(x)+\frac{a_0'(x)}2\,dx-ca_1\right)\!a_0 \\
1000: &=& a_0\Big(\mu(X)-\mu_c(\I_Z)\Big)\int_0^c\!a_0(x)dx. \label{F1}
1001: \end{eqnarray}
1002: This is a strictly positive multiple of $\mu(X)-\mu_c(\I_Z)$, and K-stability
1003: (K-semistability) implies it is strictly positive (nonnegative). This gives
1004: the result so long as, in the semistable/polystable case, the test configuration
1005: is not trivial. But the central fibres of both $(\X,\L_c)$ and $(\Y,\O_\Y(1))$
1006: have nontrivial $\C^\times$-actions.
1007: \end{proof}
1008:
1009: This result means that slope instability provides an obstruction to the existence
1010: of constant scalar curvature K\"ahler metrics (and so also K\"ahler-Einstein
1011: metrics); see \cite{RT} where there are also numerous examples.
1012:
1013: Letting $\tilde{a}_i(x) = a_i-a_i(x)$ be the coefficients of
1014: $\chi(L^k\otimes\O_{xkZ})=\chi(L^k/(L^k\otimes\I^{xk}_{\!Z}))$, we define
1015: the slope of $\O_Z$ (in slightly misleading notation) to be
1016: \begin{equation} \label{qslope}
1017: \mu_c(\O_Z)=\frac{\int_0^c \tilde{a}_1(x) + \frac{\tilde{a}'_0 (x)}{2}
1018: dx}{\int_0^c \tilde{a}_0(x)dx}\,.
1019: \end{equation}
1020: Notice that we can rephrase slope stability in the equivalent ways
1021: $$
1022: \mu_c(\I_Z)<\mu(X)\ \Longleftrightarrow\ \mu(X)<\mu_c(\O_Z)\ \Longleftrightarrow\
1023: \mu_c(\I_Z)<\mu_c(\O_Z),
1024: $$
1025: due to the implications
1026: $$
1027: \frac AB< \frac CD\ \Longleftrightarrow\
1028: \frac CD<\frac{C-A}{D-B}\ \Longleftrightarrow\ \frac AB<\frac{C-A}{D-B}
1029: $$
1030: for $0<B<D$, on setting $B=\int_0^c a_0(x) dx$ and $D=ca_0$ (so $D-B=\int_0^c
1031: \tilde{a}_0(x)dx>0$).
1032:
1033: \begin{rmks} \label{normal}
1034: On the blow up $\widehat X$ we have the formula
1035: \begin{equation} \label{a0}
1036: a_0(x)=\frac1{n!}(L-xE)^n,
1037: \end{equation}
1038: and so $a_0(0)=a_0$, since the intersection $L^n$ can be calculated
1039: equally on $X$ or $\widehat X$.
1040:
1041: The $a_i(0)$ are the coefficients of the polynomial
1042: $$
1043: \chi\_{\widehat{X}}(L^k)=\chi\_X ((\mathbf R\pi_*\O_{\widehat X})\otimes
1044: L^k)=\sum_{i=0}^n(-1)^ih^0_X(R^i\pi_*\O_{\widehat X}\otimes L^k)
1045: \quad\text{for }k\gg0.
1046: $$
1047: \emph{If $X$ is normal}, then $\pi_*\O_{\widehat X}=\O_X$ (\cite{Ha} proof
1048: of Corollary III.11.4). $E$ has dimension $n-1$ and $R^i\pi_*\O$
1049: is supported on points over which the fibre has dimension $\ge i$, so the
1050: support of $R^i\pi_* \O_{\widehat X}$ has codimension at least
1051: $i+1$. Hence $\chi(R^i\pi_* \O_{\widehat X}\otimes L^k) =
1052: O(k^{n-1-i})$, and $\chi\_{\widehat X}(\pi^*L^k)= \chi\_X(L^k) +
1053: O(k^{n-2})$ so $a_1(0)=a_1$ also.
1054:
1055: (The same argument shows that if $Z$ has dimension $j$, then $a_i(0)=a_i$
1056: for $i\le\max\{n-j-1,0\}$ (and $i\le\max\{n-j-1,1\}$ for $X$ normal).
1057: If $Z\subset X$ are both smooth in a neighbourhood
1058: of $Z$, then $\mathbf R\pi_* \O_{\widehat X}=\O_X$ so $a_i(0)=a_i$ for
1059: all $i$.)
1060:
1061: Since $H^0(L^r\otimes\I^{xr}_{\!Z})\subseteq H^0(L^r\otimes\I^{yr}_{\!Z})$
1062: for $x>y$, $a_0(x)$ is a decreasing function in $x$: $a_0'(x)\le0$. In fact
1063: from (\ref{a0}),
1064: \begin{equation} \label{a0'}
1065: a_0'(x)=-\frac{1}{(n-1)!}
1066: (L-xE)^{n-1}.E<0 \quad\text{for }x\in(0,\epsilon(Z)),
1067: \end{equation}
1068: by the ampleness of $L-xE$.
1069: In particular $a_0(x)<a_0$ for $x\in(0,\epsilon(Z))$, showing that $\mu_c(\O_Z)$
1070: (\ref{qslope}) is finite.
1071:
1072: For $X$ normal then,
1073: \begin{equation} \label{decreasing}
1074: \mu_c(\I_Z)=\mu(X)+\frac{a_0'(0)}{2a_0} + O(c)
1075: \end{equation}
1076: is strictly less than $\mu(X)$ for small $c$, and the slope inequality
1077: is automatically satisfied. In all of the examples we have
1078: considered \cite{Ro, RT}, one need only test
1079: the slope inequality at $c=\epsilon(Z)$.
1080: If this held in general it would
1081: simplify the definition of stability. Sz\'ekelyhidi \cite{Sz} has shown
1082: by example that for the modification of K-stability relevant to extremal
1083: metrics this is not the case.
1084: \end{rmks}
1085:
1086: \subsection*{\textbf{Simplifying destabilising subschemes}}
1087:
1088: \begin{prop}\label{simplify}
1089: Suppose that $Z\subset X$ is a strictly destabilising subscheme in the
1090: sense that it violates the slope inequality (\ref{defstab}). Then at least
1091: one of the connected
1092: components of $Z$ strictly destabilises. Similarly if $Z$ is a thickening
1093: $Z=mZ'$ of $Z'\subset X$ then $Z'$ strictly destabilises.
1094: \end{prop}
1095:
1096: \begin{proof}
1097: Suppose $Z=Z_1\cup Z_2$ with $Z_1$ and $Z_2$ disjoint. Then
1098: $\tilde{a}_i^{Z}(x)=\tilde{a}_i^{Z_1}(x) + \tilde{a}_i^{Z_2}(x)$
1099: (\ref{qslope}) since $\O_{xkZ}=\O_{xkZ_1}\oplus \O_{xkZ_2}$.
1100: Suppose $Z$ is strictly destabilises. Then there is a
1101: $c\in(0,\epsilon(Z)]\cap\Q$ such that
1102: \begin{equation} \label{simplify2}
1103: \mu_c(\O_Z) = \frac{\big(\int_0^c\tilde a_0^{Z_1}(x)dx\big)\mu_c(\O_{Z_1})+
1104: \big(\int_0^c\tilde a_0^{Z_2}(x)dx\big)\mu_c(\O_{Z_2})}{\int_0^c
1105: \tilde{a}_0^{Z_1}(x) dx + \int_0^c \tilde{a}_0^{Z_2}(x) dx} < \mu(X).
1106: \end{equation}
1107: This implies that for some $j\in\{0,1\}$, $\mu_c(\O_{Z_j})< \mu(X)$,
1108: and by Lemma \ref{seshadridisconnected} $c\le \epsilon(Z_j)$, so $Z_j$ is
1109: strictly destabilising.
1110:
1111: Finally if $(\I_{\!Z}^m,c)$ destabilises then so does $(\I_{\!Z},mc)$ since
1112: $\epsilon(mZ)=\frac1m\epsilon(Z)$ and
1113: \begin{equation*}
1114: \mu_c(\I_{\!Z}^m)=\mu_{mc}(\I_Z) + (m-1)\frac{\int_0^{mc}a_0'(x)dx}
1115: {2\int_0^{mc}a_0(x)dx}<\mu_{mc}(\I_Z),
1116: \end{equation*}
1117: as $a_0'(x)<0$ for $x\in(0,mc)$ (\ref{a0'}).
1118: \end{proof}
1119:
1120: \begin{lem}\label{seshadridisconnected}
1121: If $Z_1\cap Z_2=\emptyset$ then $\epsilon(Z_1\cup Z_2)\le
1122: \min(\epsilon(Z_1),\epsilon(Z_2))$.
1123: \end{lem}
1124:
1125: \begin{proof}
1126: Let $\pi\colon \widehat X\to X$ be the blowup of $X$ along $Z_1\cup
1127: Z_2$ with exceptional divisor $E=E_1\cup E_2$, where $E_i$ is the
1128: subset of $E$ sitting over $Z_i$. Let $\epsilon=\epsilon(Z_1\cup
1129: Z_2)$, so by definition $\pi^* L-\epsilon E$ is nef. If $C$ is an
1130: irreducible curve contained in $E_2$ then $(\pi^* L-\epsilon
1131: E_1).C\ge0$ by ampleness of $L$ and the fact that $E_1.C=0$. If $C$
1132: is not contained in $E_2$ then $(\pi^* L-\epsilon E_1).C=(\pi^*
1133: L-\epsilon E).C + \epsilon E_2.C \ge 0$. Hence by the Kleiman
1134: criterion, $\pi^* L-\epsilon E_1$ is nef. But this line bundle is the pullback
1135: of $L-\epsilon E_1$ from $\B_{Z_1}X$, so the latter line bundle is also nef
1136: (\cite{La} Example 1.4.4(ii)), proving that $\epsilon\le\epsilon(Z_1)$.
1137: \end{proof}
1138:
1139: \begin{lem} \label{pts}
1140: Let $(X,L)$ be a smooth polarised variety of dimension $n$ and
1141: $\epsilon=\epsilon(p,L)$ be the Seshadri constant of some point $p$
1142: in $X$. Then $p$ strictly destabilises if and only if
1143: $$
1144: \big((-K_X).L^{n-1}\big)\epsilon(p,L)>(n+1)L^n.
1145: $$
1146: \end{lem}
1147:
1148: \begin{proof}
1149: We have $\tilde{a}_0(x)=a_0-a_0(x)=\frac{1}{n!}(L^n-(L-xE)^n)=
1150: -\frac{(-x)^n}{n!}E^n=\frac{x^n}{n!}$ since $c_1(L)^{\cap j},\,j>0$ can be
1151: represented by a cycle on $X$ missing $p$, so by a cycle on $\widehat X$
1152: missing $E$. Similarly using $K_{\widehat X}=K_X+(n-1)E$,
1153: $$
1154: \tilde{a}_1(x)=a_1-a_1(x)=-\frac{K_XL^{n-1}-(K_X+(n-1)E)(L-xE)^{n-1}}{2(n-1)!}=
1155: \frac{(n-1)x^{n-1}}{2(n-1)!}\,,
1156: $$
1157: yielding (\ref{qslope}),
1158: $$
1159: \mu_c(\O_p,L) = \frac{\int_0^c\frac{(n-1)x^{n-1}}{2(n-1)!}+\frac{x^{n-1}}{2(n-1)!}dx}
1160: {\int_0^c\frac{x^n}{n!}dx}=\frac{n(n+1)}{2c}\,.
1161: $$
1162: So $p$ strictly destabilises if and only if $n(n+1)a_0<2ca_1$. As
1163: this is linear in $c$ it holds for some $c\le\epsilon$ if and only if
1164: it holds for $c=\epsilon$. Substituting $a_0=\frac1{n!}L^n$ and $2a_1=
1165: -\frac1{(n-1)!}K_X.L^{n-1}$ gives the result.
1166: \end{proof}
1167:
1168: One can of course also calculate the slope of a smooth point more directly
1169: by working locally with
1170: $H^0(L^k\otimes\O_{ckp})\cong\bigoplus_{j=0}^{ck-1}S^jT^*_pX$. We now get
1171:
1172: \begin{thm} \label{pts2}
1173: If $X$ is smooth then no point strictly destabilises.
1174: \end{thm}
1175:
1176: \begin{proof}
1177: For any line bundle $A$ we say that $H^0(A)$ \emph{generates $s$-jets at
1178: $p$} if
1179: $$
1180: H^0(A) \rightarrow A \otimes\big(\O_p/\I_p^{s+1}\big)=A|_{(s+1)\{p\}}
1181: $$
1182: is a surjection, where $\I_p$ denotes the maximal ideal at $p$. We define
1183: $s(A)$ to be the maximum integer $s$ such that $H^0(A)$ generates $s$-jets
1184: at $p$ (and set $s(A)=-\infty$ if no such $s$ exists).
1185:
1186: Demailly (\cite{De} Lemma 8.6) proves that given any $s\ge0$, if $H^0(kA)$
1187: generates $(k(n+s)+1)$-jets at $p$, then $H^0(A+K_X)$ generates $s$-jets
1188: at $p$. Setting
1189: $s=(n+1)(m-1)$ (and noting that $k(n+1)m\ge k(n+s)+1$) we see that for $m\ge
1190: 1$ and any $k\ge1$,
1191: $$
1192: s(kA) \ge k(n+1)m \quad\Rightarrow\quad s(A+K_X)\ge (n+1)(m-1).
1193: $$
1194: Applying this again to $A+K_X$ (with $k=1$) implies that $s(A+2K_X)\ge(n+1)(m-2)$
1195: and so inductively $s(A+rK_X)\ge(n+1)(m-r)$ for $r\le m$. Setting $r=m$,
1196: \begin{equation}\label{generationofadjointbundles}
1197: s(kA) \ge k(n+1)m \quad\Rightarrow\quad s(A+mK_X)\ge0.
1198: \end{equation}
1199:
1200: Fix any rational number $m/M\in(0,\epsilon(p,L))$, so that
1201: $\epsilon(p,(n+1)ML)>(n+1)m$. Then by the identity (\cite{De} Lemma 7.6)
1202: $$
1203: \epsilon(p,A) = \lim_{k\to\infty}\frac{s(kA)}k\,,
1204: $$
1205: applied to $A=(n+1)ML$,
1206: we can find a $k\gg0$ such that $s(k(n+1)ML)>k(n+1)m$. Thus by
1207: (\ref{generationofadjointbundles}), $(n+1)ML+mK_X$ has a section (not vanishing
1208: at $p$), and so $((n+1)ML+mK_X).L^{n-1}\ge0$. Therefore
1209: $((n+1)L+\epsilon(p,L)K_X).L^{n-1}\ge0$.
1210: \end{proof}
1211:
1212: \begin{rmk}
1213: If $X=\PP^n$ we may as well assume that $L$ is the
1214: hyperplane bundle as slope stability is invariant under rescaling. So
1215: $\epsilon(p)=1$ for any point $p$ and $-K=(n+1)L$, which gives
1216: equality in (\ref{pts}), so $\PP^n$ is at best K-semistable (since
1217: $L^k\otimes\I_p^k$ is generated by global sections).
1218: In fact $\PP^n$ is K-polystable in the sense of Definition \ref{defstab},
1219: and the deformation to the normal cone of $\{p\}$ has semi-ample
1220: $\L_c$ for $c=1$, pulled back from a blow down to $\PP^n\times\C$.
1221: So the degeneration is a product family with a nontrivial
1222: $\C^\times$-action, and the semistability we are seeing comes from a
1223: $\C^\times$-action on $\PP^n$ with Donaldson-Futaki invariant zero.
1224: \end{rmk}
1225:
1226:
1227: \subsection*{Asymptotic slope Chow stability}
1228:
1229: If instead of K-instability we are interested in asymptotic stability,
1230: then the relevant slope function is more complicated.
1231:
1232: Let $(X,L)$ be a polarised manifold, $r$ be a positive integer, and
1233: $Z$ a subscheme of $X$ such that on the blow up $p\colon\widehat X\to X$
1234: of $X$ along $Z$,
1235: \begin{equation} \label{condition}
1236: \mathbf Rp_*\O(-jE)=p_*\O(-jE)=\I_{\!Z}^j\quad\forall j\ge0.
1237: \end{equation}
1238: (E.g. if $Z\subset X$ are both smooth in a neighbourhood of $Z$.)
1239: Then for \emph{all} $k$ and $x$ (with $k,\,xk\in\mathbb N$) we have the exact
1240: formula
1241: $$
1242: \chi(L^k\otimes\I_{\!Z}^{xk}) = a_0(x)k^n + a_1(x)k^{n-1} +\ldots+a_n(x).
1243: $$
1244: Letting $B_i$
1245: denote the Bernoulli numbers, define $\beta_0=1,\,\beta_1=\frac12$ and
1246: $\beta_i=\frac{B_i}{i!}$ for $i\ge 2$. Then we define the \emph{asymptotic
1247: Chow slope} $\eta_c(\I_Z)$ of $\I_Z$ for $c\le\epsilon(Z)$ to be
1248: \begin{eqnarray*}
1249: \eta_c(\I_Z)&=&\eta_c(\I_Z,L,r) = \sum_{j=0}^{n+1} c_j r^{n+1-j} \\
1250: &=& r^{n+1} + \mu_c(\I_Z)r^n + \ldots + c_{n+1},
1251: \end{eqnarray*}
1252: where
1253: $$
1254: c_j = \frac{\sum_{i=0}^j
1255: \int_0^c\beta_ia_{j-i}^{(i)}(x) dx}{\int_0^c a_0(x)dx}\,.
1256: $$
1257: I.e. $c_0=1,\ c_1=\frac{\int_0^c a_1(x)+\frac{a_0'(x)}2dx}{\int_0^c a_0(x)dx}
1258: =\mu_c(\I_Z)$,
1259: $$
1260: c_2=\frac{1}{\int_0^c a_0(x) dx} \int_0^c
1261: a_2(x) + \frac{a_1'(x)}{2} + \frac{a_0^{''}(x)}{12}\,dx, \quad\mathrm{etc.}
1262: $$
1263: Defining the asymptotic Chow slope of $X$ to be the slope of the empty
1264: subscheme,
1265: $$
1266: \eta\_X(r) = \eta(\O_X,L,r) = \frac{r\chi(L^r)}{a_0} = r^{n+1} + \mu(X)
1267: r^n + \frac{a_2}{a_0}r^{n-1} + \ldots + \frac{a_n}{a_0}r,
1268: $$
1269: we say (with great difficulty) that $X$ is asymptotically
1270: Chow slope strictly destabilised by $Z$ if, for all $r\gg0$,
1271: $$
1272: \eta_c(\I_Z,r)>\eta\_X(r).
1273: $$
1274:
1275: \begin{thm} \label{chow}
1276: If a polarised variety is asymptotically Chow slope strictly
1277: destabilised by $Z$ satisfying (\ref{condition}) then it is
1278: asymptotically Chow strictly unstable.
1279: \end{thm}
1280:
1281: \begin{proof}
1282: The proof is almost the same as for Theorem
1283: \ref{thm:kstableslopestable}. We calculate the coefficient
1284: $e_{n+1}(r)$ of $k^{n+1}$ in $\tilde{w}_{r,k}$
1285: (\ref{normalised}),
1286: $$
1287: e_{n+1}(r)\ =\ b_0r\chi(L^r)-a_0w(r)\ =\ (b_0+ca_0)r\chi(L^r)-
1288: a_0\big(w(r)+cr\chi(L^r)\big)
1289: $$
1290: (instead of its $r^n$ coefficient the Donaldson-Futaki invariant) for the
1291: deformation to the normal cone. By Theorem \ref{thm:instability}, if
1292: $e_{n+1}(r)<0$ then $(X,L)$ is Chow unstable with respect to $r$. By the
1293: continuity of $\eta_c$ we may take $c<\epsilon(Z)$,
1294: so that $w(r)$ is calculated by Theorem \ref{degen} and $b_0$ by Proposition
1295: \ref{integrals}, giving
1296: $$
1297: e_{n+1}(r)<0 \quad\Longrightarrow\quad
1298: \int_0^ca_0(x)dx\,r\chi(L^r)<a_0\sum_{j=1}^{cr}\chi(L^r\otimes\I_{\!Z}^j).
1299: $$
1300: Instead of the trapezium rule (\ref{easyeuler}) we use the fact \cite{Hi}
1301: that for any polynomial $f$,
1302: $$
1303: \sum_{j=1}^{cr} f(j/r) = \int_0^c \sum_{i=0}^n
1304: \beta_i \frac{f^{(i)}(x)}{r^{i-1}}dx.
1305: $$
1306: (The case $f(x)=x^m$ follows from the definition of $\beta_i$, implying the
1307: general case by linearity.) The theorem follows
1308: by applying this to the polynomial $f(x)= a_0(x)r^n + a_1(x)r^{n-1}
1309: +\ldots+a_n(x)=\chi(L^r\otimes\I_{\!Z}^{xr})$.
1310: \end{proof}
1311:
1312:
1313: \section{Simplifying arbitrary test configurations} \label{testconfigs}
1314:
1315: We begin with an important technical result allowing us to calculate weights
1316: of a test configuration $(\Y,\O_\Y(k))$ in terms of a semi test configuration
1317: $(\X,\L^k)$ that dominates it.
1318:
1319: So let $(\Y,\O_\Y(1))$ be an equivariantly polarised flat $\C^\times$-family
1320: with general fibre $(X,L)$, and fix another flat $\C^\times$-family $(\X,\L)\to\C$
1321: with a birational $\C^\times$-equivariant map
1322: $$
1323: p\colon(\X,\L)\to(\Y,\O_\Y(1)) \qquad \text{such that }\ \L=p^*\O_\Y(1)
1324: \ (\text{equivariantly}).
1325: $$
1326:
1327: \begin{prop} \label{stein}
1328: If $X$ is normal then there exists an $a\ge0$ such that
1329: $$
1330: w(H^0_{\Y_0}\!(\O_\Y(k)))
1331: =w\left(H^0_\X(\L^k)\big/tH^0_\X(\L^k)\right)-ak^n+O(k^{n-1}).
1332: $$
1333: \end{prop}
1334:
1335: \begin{proof}
1336: $p$ factors through the its Stein factorisation (\cite{Ha} Corollary
1337: III.11.5) as
1338: $$
1339: (\X,\L)\Rt{q}(\X'=\Proj\bigoplus_kH^0_\X(\L^k),\O\_{\!\X'}(1))\Rt{p'}(\Y,\O_\Y(1)).
1340: $$
1341: $q_*$ induces an equivariant isomorphism between $H^0_\X(\L^k)$ and
1342: $H^0_{\X'}(\O\_{\!\X'}(k))$ for $k\gg0$, and $\X'$ is flat over $\C$ since
1343: $H^0_\X(\L^k)$ has no $t$-torsion by the flatness of $\X$.
1344: Thus replacing $\X$ by $\X'$ and $p$ by $p'$ if necessary, we may assume
1345: that $p$ is finite and so $\L=p^*\O_\Y(1)$ is ample. Moreover, the general
1346: fibre of $\X'$ is $(X,L)$ by the normality of $X$, so
1347: $(\X,\L^k)$ is now a genuine test configuration for $(X,L^k)$.
1348:
1349: So we have a $\C^\times$-equivariant morphism of polarised families
1350: $(\X\rt{\Pi\,}\C,\L)\Rt{p}(\Y\rt{\pi}\C,\O_\Y(1))$, and we wish to relate
1351: the total weights of the $\C^\times$-actions on
1352: $H^0_{\Y_0}(\O_\Y(k))$ and $H^0_\X(\L^k)\big/tH^0_\X(\L^k)$; but the latter
1353: is now isomorphic to $H^0_{\X_0}(\L^k)$ for $k\gg0$ by ampleness and flatness.
1354: $p^*$ induces an exact sequence
1355: $$
1356: 0\to\O_\Y\to p_*\O_\X\to Q\to0,
1357: $$
1358: for some cokernel $Q$ supported on $\Y_0$ (since $p$ is an isomorphism away
1359: from the central fibres).
1360: So $t^sQ=0$ for some $s\ge0$. We first give the argument for $s=1$
1361: to illustrate the more technical general case.
1362:
1363: Tensoring with $\O_\Y(k)$ and pushing down to $\C$ gives an exact sequence on
1364: $\C$,
1365: $$
1366: 0\to\pi_*(\O_\Y(k))\to\Pi_*(\L^k)\to Q_k\to0,
1367: $$
1368: where $Q_k=\pi_*(Q\otimes \O_\Y(k))$ satisfies $tQ_k=0$. Therefore its restriction to $0\in\C$
1369: is also isomorphic to $Q_k$, and Tor$_1(Q_k,\O_0)=\ker\big(Q_k\Rt{\times t}Q_k\big)\cong
1370: Q_k$, giving the exact diagram
1371: $$
1372: \begin{array}{ccccccccc}
1373: & 0 & \to & \pi_*(\O_\Y(k)) & \Rt{p^*} & \Pi_*(\L^k) & \to & \!\!Q_k & \to0
1374: \\ & \downarrow && \downarrow && \downarrow && \downarrow\wr \\
1375: 0\to & Q_k\! & \to & H^0_{\Y_0}(\O_\Y(k)) & \Rt{p^*} & H^0_{\X_0}(\L^k)
1376: & \to & Q_k|\_0\! & \to0 \\
1377: &&& \downarrow && \downarrow \\ &&& 0 && 0
1378: \end{array}
1379: $$
1380: where the vertical arrows are restriction to $0\in\C$, and the flatness of
1381: $\pi,\,\Pi$ and ampleness of $\O_\Y(1),\,\L$ give the two
1382: central terms and ensure that Tor$_1(\Pi_*(\L^k),\O_0)=0$.
1383:
1384: These are maps of $\C^\times$-modules, except that the left hand $Q_k$
1385: has weight shifted by $-1$ (i.e.\ is isomorphic to $Q_k\otimes\langle t\rangle$
1386: as a $\C^\times$-module). We see this by exhibiting an explicit weight-$(-1)$
1387: isomorphism $\delta$ from $Q_k$ to the kernel of the lower $p^*$. Given $q\in
1388: Q_k$, choose a lift $\hat q\in\Pi_*(\L^k)$. $t\hat q$ has zero image in $Q_k$
1389: so is in the image of some $f\in\pi_*\O_\Y(k)$. $f|\_0\in H^0_{\Y_0}(\O_\Y(k))$
1390: maps under $p^*$ to $t\hat q|\_0\in H^0_{\X_0}(\L^k)$, i.e.\ zero, so is
1391: an element $\delta(q)\in\ker p^*=\,$Tor$_1(Q_k,\O_0)$. Since this map involved
1392: multiplication by $t$ it has weight $-1$.
1393:
1394: So in this case of $s=1$, $w(H^0_{\Y_0}\!(\O_\Y(k)))=w(H^0_{\X_0}(\L^k))-\dim Q_k$, and
1395: $\dim Q_k=ak^n+O(k^{n-1})$, where $a=$\,rank$(Q_k)\O_\Y(1)^n/n!\ge0$ since $Q_k$ is
1396: supported on the $n$-dimensional central fibre.
1397:
1398: For general $s$ we still have the exact sequence
1399: $$
1400: 0\to\mathrm{Tor}_1(Q_k,\O_0)\to H^0_{\Y_0}(\O_\Y(k))\Rt{p^*} H^0_{\X_0}(\L^k)
1401: \to Q_k|\_0\to0,
1402: $$
1403: where again the first term $\ker\big(Q_k\Rt{\times t}Q_k\big)$ is isomorphic
1404: to the last $Q_k|\_0$. For instance if we pick an isomorphism $Q_k\cong\bigoplus_i
1405: \C[t]/(t^{j_i})$, each $j_i\le s$, then $Q_k|\_0\cong\bigoplus_i\C[t]/(t)$
1406: maps isomorphically via $\oplus_it^{j_i-1}$ to $\bigoplus_i(t^{j_i-1}\C[t])
1407: /(t^{j_i})=\ker\big(Q_k\Rt{\times t}Q_k\big)$. More invariantly, filter
1408: $Q^k$ by $\ker t^j\subseteq\ker t^{j+1}\subseteq\ldots\subseteq\ker t^s=Q^k$,
1409: and so $Q_k|\_0$ by $\ker t^j/(\im t\cap\ker t^j)$ with graded pieces $V^j=
1410: \ker t^j/(\ker t^{j-1}+\im t\cap\ker t^j)$. Then there exists a weight-$(-j)$
1411: ``multiplication
1412: by $t^j$" map taking $V^j$ into $H^0_{\Y_0}(\O_\Y(k))$ whose sum over $j\le
1413: s$ gives an isomorphism to Tor$_1(Q_k,\O_0)=\ker p^*$.
1414:
1415: The map is defined roughly as before. Given $q\in V^j$, lift to $\ker(Q_k\rt{t^j}Q_k)$
1416: and then to $\hat q\in\Pi_*(\L^k)$. The image of $t^j\hat q$ in $Q_k$
1417: is zero by construction, so is in the image of $f\in\pi_*\O_\Y(k)$. $f|\_0\in
1418: H^0_{\Y_0}(\O_\Y(k))$ maps to $t^j\hat q|\_0\in H^0_{\X_0}(\L^k)$, i.e.\ to
1419: zero since $j\ge1$. Thus $f|\_0=\delta(q)$ for some $\delta(q)\in\ker p^*$.
1420: $\ker t^{j-1}$ and $\im t\cap\ker t^j$ clearly map to zero via this construction,
1421: so $\delta$ is well defined. The upshot is that
1422: $$
1423: w(H^0_{\Y_0}\!(\O_\Y(k)))=w(H^0_{\X_0}(\L^k))-\sum_{j=0}^sj\dim V_j.
1424: $$
1425: $\sum_{j=0}^sj\dim V_j$ is a polynomial in $k$ (since both weights $w$
1426: are), so is $ak^n+O(k^{n-1})$ for some $a\ge0$ because $0\le\sum\dim V_j=\dim
1427: Q_k=O(k^n)$ as before.
1428: \end{proof}
1429:
1430: \begin{rmk} \label{assumenormal}
1431: In particular if $X$ is normal then for K-stability we need only
1432: consider \emph{normal} test configurations. This is because
1433: any test configuration $(\Y,\O_\Y(1))$ is dominated by its normalisation,
1434: which also has general fibre $X$ if $X$ is normal. The pullback of $\O_\Y(1)$
1435: is ample, so some twist gives another test configuration which is \emph{less
1436: stable} than $(\Y,\O_\Y(1))$, in the sense that it has the same weight to
1437: leading order and a smaller Donaldson-Futaki invariant, by Proposition
1438: \ref{stein} above.
1439:
1440: Similarly for Chow stability we may compute the Chow weight of any test
1441: configuration on its normalisation. Of course most of our test configurations
1442: have nonnormal central fibre, however.
1443: \end{rmk}
1444:
1445: Given any test configuration $(\Y,\O_\Y(1))$ (\ref{any}) we now build
1446: inductively the semi test configuration $(\B_{I_r}(X\times\C),L-E)$
1447: (\ref{ideals}) that dominates it, starting from the deformation to the normal
1448: cone $(\B_{I_1}(X\times\C),L-P)$ of $Z_0$.
1449:
1450: So let $\X^1\rt{\pi^1}X\times\C$ denote $\B_{I_1}(X\times\C)$, i.e.\ the blow
1451: up in $Z_0\times\{0\}$:
1452: $$
1453: \X^1=\Proj\bigoplus_k\,\S_k, \qquad\qquad \S_k:=(\I_0+(t))^k.
1454: $$
1455: Recall that the central fibre of $\X^1$ is $\widehat X\cup_eP$, where $\widehat
1456: X$ is the blow up of $X$ along $Z_0$ with exceptional divisor $e$, and $P$
1457: is the exceptional divisor of $\pi^1$.
1458: $P$ is a projective cone over $Z_0$ (Proj of the graded algebra over $Z_0$
1459: with $k$th graded piece $\bigoplus_{i=0}^{k-1}\I_0^i/\I_0^{i+1}$) -- the projective
1460: completion of the normal cone to $Z_0\subset X$. Its zero
1461: section $Z_0'$ is a copy of $Z_0$ which fits into a flat
1462: family with the $Z_0\times\{t\}$ in each fibre $X_t$, which we see as follows.
1463:
1464: The proper transform $\overline{Z_0\times\C}$ is defined
1465: by the graded sheaf of ideals generated by $\I_0\subset\S_1=\I_0+(t)$
1466: in the graded sheaf of algebras $\bigoplus_k\S_k$. That is, $\O(-P)\otimes
1467: \I_{\overline{\!Z_0\times\C}}$ is generated by the sections of $\I_0\subset\S_1$.
1468: It is abstractly isomorphic
1469: to the blow up of $Z_0\times\C$ along its intersection with $Z_0\times\{0\}$,
1470: but this is a divisor in $Z_0\times\C$, so $\overline{Z_0\times\C}\cong
1471: Z_0\times\C$. The central fibre $Z_0'\cong Z_0$ is
1472: defined by the graded sheaf of ideals generated by $\I_0+t\S_1=\I_0+(t^2)
1473: \subset\S_1$.
1474:
1475: \begin{figure}[h]
1476: \center{\input{normalcone.pstex_t}}
1477: \end{figure}
1478:
1479: Similarly the proper transform of $Z_1\times\C$ is the blow up of $Z_1\times\C$
1480: along its intersection $Z_1\times\{0\}$ with $Z_0\times\{0\}$; that is
1481: $\overline{Z_1\times\C}\cong Z_1\times\C$. It is defined by the graded sheaf
1482: of ideals generated by $\I_0+\I_1\S_1=\I_0+t\I_1\subset\S_1$,
1483: with central fibre $Z_1'\subseteq Z_0'\subset P$ isomorphic to $Z_1$ and
1484: defined by
1485: \begin{equation} \label{I1'}
1486: \I_0+(\I_1+(t))\S_1=\I_0+t\I_1+(t^2)\ \subset\ \S_1.
1487: \end{equation}
1488:
1489: We now form $\X^2$ by blowing up $\X^1$ in $Z_1'$. Since $\I_0+t\I_1+(t^2)$
1490: is just $I_2$ (\ref{ideals}), we have basically shown that $\X^2$ dominates
1491: $\B_{I_2}(X\times\C)$. Precisely, we have maps $\X^2\rt{\pi^2}
1492: \X^1\rt{\pi^1}X\times\C$, and set $E_1:=(\pi^1)^*P,\ E_2$ to be the exceptional
1493: divisor of $\pi^2$, and $E$ to be the exceptional divisor of $\B_{I_2}(X\times\C)$.
1494: While $\O(-cE_1-eE_2)$ is relatively
1495: ample for $0<e<c$, it is only semi-ample for $e=c$\,:
1496:
1497: \begin{prop} \label{p}
1498: $\X^2\to X\times\C$ factors through a map $p^2\colon\X^2\to\B_{I_2}(X\times\C)$.
1499: Under this map, $(p^2)^*(\O(-E))=\O(-E_2-E_1)$.
1500: \end{prop}
1501:
1502: \begin{proof}
1503: For $k$ sufficiently large, $\pi^1_*\O(-kE_1)=(\I_0+(t))^k=\S_k$, in which
1504: the ideal $(\I_0+t\I_1+(t^2))^k$ defines the $k$th power of the ideal of
1505: $Z_1'$ (\ref{I1'}). That is, $(\pi^1\comp\pi^2)^*(\I_0+t\I_1+(t^2))^k\cong
1506: (\pi^2)^*\I_{\!Z_1'}^k(-kP)=\O(-kE_1-kE_2)$.
1507: Therefore the sections $(\I_0+t\I_1+(t^2))^k\subset(\pi^1\comp\pi^2)_*\O(-kE_1-kE_2)$
1508: generate $\O(-kE_1-kE_2)$ and so define a regular map from $\X^2$ to
1509: $\Proj\bigoplus_{k\gg0}(\I_0+t\I_1+(t^2))^k$, i.e.\ to
1510: $$
1511: \Proj\bigoplus_{k\gg0}I^k_2=\B_{I_2}(X\times\C). \vspace{-9mm}
1512: $$ \vspace{2mm}
1513: \end{proof}
1514:
1515: In fact the contraction $p^2$ just collapses the restriction $P\res{Z_1}$
1516: of the cone $P\to Z_0$ down to $Z_1$, but we will not need this.
1517:
1518: Similarly in $\X^2$ there is a copy $Z_1''$ of $Z_1$, sitting in a flat
1519: family with $Z_1\subset X$ as the central fibre of the proper transform
1520: $\overline{Z_1\times\C}$. In the coordinate ring
1521: $\bigoplus_k\S_k=\bigoplus_k(\I_0+t\I_1+(t^2))^k$ (we are recycling the symbol
1522: $\S_k$ and shall do so again below) pulled back from $\X$ by the above map
1523: $p^2$ (\ref{p}), $\I_{\overline{Z_1\times\C}}$ is generated by $\I_0+t\I_1\subset
1524: (\I_0+t\I_1+(t^2))$, since this is the largest ideal that localises to $\I_1$
1525: when we invert $t$ and work on $X\times\C^\times$. Thus $\I_{Z_1''}$ is generated
1526: by $\I_0+t\I_1+t(\I_0+t\I_1+(t^2))=\I_0+t\I_1+(t^3)$.
1527:
1528: So there is also a $Z_2''\subset Z_1''$, isomorphic to $Z_2$, inside it.
1529: $\overline{Z_2\times\C}$ has ideal generated by $\I_0+t\I_1+t^2\I_2\subset
1530: \S_1=\I_0+t\I_1+(t^2)$, since this is the largest ideal that localises to $\I_2$
1531: when we invert $t$ and work on $X\times\C^\times$. Thus its central fibre
1532: $Z_2''$ is defined by the ideal
1533: \begin{equation} \label{I3}
1534: \I_0+t\I_1+t^2\I_2+t\S_1=\I_0+t\I_1+t^2\I_2+(t^3).
1535: \end{equation}
1536:
1537: Blowing it up gives $\X^3$, with exceptional divisor $E_3$ (and we denote
1538: the pullbacks to $\X^3$ of $E_1,\,E_2$ by the same notation). Then, just
1539: as in Proposition \ref{p}, the pushdown of $\O_{\X^3}(-E_3-E_2-E_1)$ to $X\times\C$
1540: is generated by $I_3=\I_0+t\I_1+t^2\I_2+(t^3)$ by (\ref{I3}), so $\X^3\to X\times\C$
1541: factors through $\B_{I_3}(X\times\C)$.
1542:
1543: Inductively we obtain $\X^s\to X\times\C$
1544: as the blow up $\pi^s$ along $Z_{s-1}^{(s-1)}\subset\X^{s-1}$,
1545: the central fibre of the proper transform of $Z_{s-1}
1546: \times\C$. The coordinate ring of $\X^{s-1}$ over $X\times\C$ has $k$th graded
1547: piece $\S_k=(\I_0+t\I_1+\ldots+t^{s-2}\I_{s-2}+(t^{s-1}))^k$, and the ideal
1548: of the proper transform of $\overline{Z_{s-1}\times\C}$ is generated by
1549: $\I_0+\ldots+t^{s-2}\I_{s-2}+t^{s-1}\I_{s-1}\subset\S_1=\I_0+\ldots+t^{s-2}\I_{s-2}
1550: +(t^{s-1})$, since this is the largest ideal that localises to $\I_{s-1}$
1551: when we invert $t$ and work on $X\times\C^\times$. Therefore the ideal of
1552: its central fibre $Z_{s-1}^{(s-1)}$ is generated by
1553: $$
1554: \I_0+\ldots+t^{s-2}\I_{s-2}+t^{s-1}\I_{s-1}+t\S_1\ =\
1555: \I_0+\ldots+t^{s-2}\I_{s-2}+t^{s-1}\I_{s-1}+(t^s).
1556: $$
1557: Thus the pushdown of $\O\_{\!\X^s}(-E_s-\ldots-E_1)$ contains $\I_0+t\I_1+
1558: \ldots+t^{s-1}\I_{s-1}+(t^s)$, giving the following, as in Proposition \ref{p}.
1559:
1560: \begin{thm} \label{pr}
1561: $\X^s\to X\times\C$ factors through a map $p^s\colon\X^s\to\B_{I_s}(X\times\C)$,
1562: where $I_s=\I_0+t\I_1+\ldots+t^s\I_{s-1}+(t^s)$ (\ref{ideals}).
1563: Under this map, $(p^s)^*(\O(-E))=\O(-E_s-\ldots-E_1)$. \hfill$\square$
1564: \end{thm}
1565:
1566: In turn any test configuration $(\Y,\O_\Y(1))$ is dominated by a map $\phi$
1567: from some $\B_{I_r}(X\times\C)$ (\ref{any}), giving $\rho^r:=\phi\comp p^r\colon
1568: \X^r\to\Y$. Denote by $L_r$ the semi-ample line bundle $(\pi^1\comp\ldots\comp
1569: \pi^r)^*L-E_r-\ldots-E_1$ on $\X^r$, so that $(\rho^r)^*\O_\Y(1)=L_r$.
1570: Then by the above and Proposition \ref{stein} we have
1571:
1572: \begin{cor} \label{ref}
1573: The total weight on (the $k$th twist of) an arbitrary test configuration
1574: $(\Y,\O_\Y(1))$ can be calculated on $(\X^r,L_r)$ by
1575: $$
1576: w(H^0_{\Y_0}(\O_\Y(k))=
1577: w\left(H^0_{\X^r}(L_r^k)\big/tH^0_{\X^r}(L_r^k)\right)-ak^n+O(k^{n-1}).
1578: $$
1579: \end{cor}
1580:
1581: \begin{cor} \label{seshadriIr}
1582: The Seshadri constant $\epsilon(I_r)$ of the ideal $I_r$ (\ref{ideals})
1583: is $\min\{\epsilon(Z_i)\}_{i=1}^{r-1}$.
1584: \end{cor}
1585:
1586: \begin{proof}
1587: If $c\in(0,\min\{\epsilon(Z\_i)\}_{i=1}^{r-1})\cap\Q$ then for $k\gg0$, the sections
1588: $$
1589: H^0_X(L^k\otimes\I_0^{ck})\,\oplus\,t^{ck}H^0_X(L^k\otimes\I_1^{ck})\,\oplus\,
1590: \ldots\,\oplus\,t^{(r-1)ck}H^0_X(L^k\otimes\I_{r-1}^{ck})\,\oplus\,t^{rck}H^0_X(L^k)
1591: $$
1592: saturate $I^{ck}_r$ (\ref{ideals}). Therefore the Seshadri constant of $I_r$
1593: (which is $\ge1$ by the semi ampleness of $L-E$ in (\ref{any})) is at least
1594: $\min\{\epsilon(Z\_i)\}_{i=1}^{r-1}$.
1595:
1596: We prove the opposite inequality inductively. The induction begins for
1597: $r=0$ by Proposition \ref{ample}; suppose it is true for $r$. That is,
1598: $L-xE$ is nef on $\B_{I_r}(X\times\C))$ precisely when
1599: $x\in[0,\min\{\epsilon(Z_i)\}_{i=1}^{r-1}]$. Equivalently, pulling back by
1600: $p^r$, $L-x(E_1+\ldots+E_r)$ is nef on $\X^r$ if and only if $x\in[0,
1601: \min\{\epsilon(Z_i)\}_{i=1}^{r-1}]$. Pick $c$ such that
1602: $L-c(E_1+\ldots+E_{r+1})$ is nef on $\X^{r+1}$; we must show that
1603: $c\le\min\{\epsilon(Z_i)\}_{i=1}^{r-1}$ and $c\le\epsilon(Z_r)$.
1604:
1605: To show the former, we claim that since $L-c(E_1+\ldots+E_{r+1})$
1606: is nef on $\X^{r+1}$, $L-c(E_1+\ldots+E_r)$
1607: is nef on $\X^r$. Given any irreducible proper curve $C\subset\X^r$ not entirely
1608: contained in $Z_r^{(r)}$, let $\overline C$ denote its proper transform in
1609: $\X^{r+1}$. Then $[L-c(E_1+\ldots+E_r)].C\ge[L-c(E_1+\ldots+E_{r+1})].\overline
1610: C\ge0$ since $\overline C.E_{r+1}\ge0$. On the other hand if $C\subset Z_r^{(r)}$,
1611: then there is an isomorphic copy $C'\subset Z_r^{(r+1)}\subset\X^{r+1}$ such
1612: that $[L-c(E_1+\ldots+E_r)].C=[L-c(E_1+\ldots+E_{r+1})].C'\ge0$. So indeed
1613: $L-c(E_1+\ldots+E_r)$ is nef on $\X^r$ and so $c\le\min\{\epsilon(Z_i)\}_{i=1}^{r-1}$
1614: by induction.
1615:
1616: Secondly, fix $c$ such that $L-c(E_1+\ldots+E_r)$
1617: is nef on $\X^r$. $Z_r^{(r)}\cong Z_r$ lies in the central fibre $(\X^r)\_0$,
1618: fitting into a flat family with
1619: $Z_r\times\C^\times$ away from the central fibre. Seshadri constants are
1620: lower semicontinuous in polarised families, so the Seshadri constant (with
1621: respect to $L-c(E_1+\ldots+E_r)$) of $Z_r^{(r)}$ \emph{inside
1622: the central fibre $(\X^r)\_0$} is at most $\epsilon(Z_r)$ (with
1623: respect to $L$, since this is $L-c(E_1+\ldots+E_r)$ restricted to a general
1624: fibre). The Seshadri constant of $Z_r^{(r)}$ inside the whole of
1625: $\X^r$ can only be smaller still; therefore if $L-c(E_1+\ldots+E_r)-cE_{r+1}$
1626: is nef then $c\le\epsilon(Z_r)$.
1627: \end{proof}
1628:
1629: We could contract each $\X^s$ using $L_s$; this would give an
1630: isomorphism in a neighbourhood of $Z_s^{(s)}$ by (\cite{Ha}
1631: Proposition II.7.3) since $L_s|_{Z_s^{(s)}}
1632: \cong L|\_{Z_s}$ is ample \emph{and} $L_s^k\otimes\I_{\!Z_s^{(s)}}$
1633: is globally generated (by sections of $L^k\otimes(\I_0+t\I_1+\ldots+
1634: t^s\I_s+(t^{s+1}))$ on $X$). Thus we could proceed inductively with these
1635: contracted $\X^s$s with \emph{ample} line bundles on them, but since we
1636: cannot currently seem to get significantly better estimates by working with
1637: ample bundles we proceed with the semi-ample $(\X^s,L_s)$ and contract at
1638: the last, $r$th stage. By Corollary \ref{ref} we lose nothing by ignoring
1639: this contraction and simply calculating the weight on $\X^r$.
1640:
1641: So, modulo the contraction, we have exhibited any test
1642: configuration (\ref{any}) as a finite number of blow ups (starting with
1643: $X\times\C$) \emph{in subschemes $Z^{(i)}_i$ supported in the scheme theoretic
1644: central fibre} that themselves sit in flat families with the $Z_i\subset X$. We
1645: calculate the weight on such a blow up in Theorem \ref{big}, for which
1646: we need two preliminary results.
1647:
1648: The following Proposition is the appropriate generalisation to general
1649: test configurations of the case $Z\times\C\subset X\times\C$ used in (\ref{St}),
1650: (\ref{qt}) and (\ref{error}). We will apply it to the flat families
1651: $\Z=\overline{Z_s\times\C}\,\subset\X^s\to\C$
1652: and their thickenings $k\Z$, when these thickenings are also flat.
1653:
1654: \begin{prop} \label{central}
1655: Fix flat families $\Z\subset\X\to\C$ with central fibres
1656: $Z'\subset\X_0$, such that the thickenings $k\Z\subset\X\to\C$
1657: are also flat over $\C$. Let $\I\_{\!Z'\subset\X}$ (respectively
1658: $\I_{Z'}$) denote the ideal sheaf of $Z'\subset\X$ ($Z'\subset\X_0$).
1659: Then
1660: $$
1661: \frac{\I_{\!Z'\subset\X}^k}{t\I_{\!Z'\subset\X}^k}\ \cong\
1662: \I_{\!Z'}^k\ \oplus\ \bigoplus_{j=1}^k\,t^j
1663: \frac{\I_{\!Z'}^{k-j}}{\I_{\!Z'}^{k-j+1}}\,.
1664: $$
1665: \end{prop}
1666:
1667: \begin{proof}
1668: Flatness of $\X\to\C$ and $j\Z\to\C$ imply the exactness of the
1669: bottom two rows of the following, from which follows the exactness of the
1670: top row.
1671: $$
1672: \begin{array}{ccccccc}
1673: &0&&0&&0 \\
1674: & \downarrow && \downarrow && \downarrow \\
1675: 0\to & \I_\Z^j & \rt{t} & \I_\Z^j & \to & \I_{\!Z'}^j & \to0 \\
1676: & \downarrow && \downarrow && \downarrow \\
1677: 0\to & \O_\X & \rt{t} & \O_\X & \to & \O_{\X_0} & \to0 \\
1678: & \downarrow && \downarrow && \downarrow \\
1679: 0\to & \O_{j\Z} & \rt{t} & \O_{j\Z} & \to & \O_{jZ'} & \to0 \\
1680: & \downarrow && \downarrow && \downarrow \\
1681: &0&&0&&0
1682: \end{array}
1683: $$
1684: Chasing through either the top two rows or the first two columns then shows
1685: that in $\O_\X$, $(t)\cap\I_\Z^j=t\I_\Z^j$. Also
1686: by flatness there is a similar diagram with $t$ replaced by $t^i$ (and the
1687: right hand column suitably modified) showing that in fact
1688: \begin{equation} \label{one}
1689: (t^i)\cap\I_\Z^j=t^i\I_\Z^j.
1690: \end{equation}
1691: Applying $H^0_\X(L\otimes\,\cdot\,)$ gives a similar exact diagram without
1692: the right hand and lower zeros, so the same argument shows that
1693: \begin{equation} \label{one'}
1694: (t^i)\cap H^0_\X(L\otimes\I_\Z^j)=t^iH^0_\X(L\otimes\I_\Z^j).
1695: \end{equation}
1696: The top row also gives $\I_\Z^j/t\I_\Z^j=\I_{\!Z'}^j$, which implies that
1697: \begin{equation} \label{four}
1698: \frac{\I_\Z^j}{\I_\Z^{j+1}+t\I_\Z^j}=\frac{\I_{\!Z'}^j}{\I_{\!Z'}^{j+1}}\,.
1699: \end{equation}
1700: $\I\_{Z'\subset\X}=\I_\Z+(t)$, so $\I_{\!Z'\subset\X}^k=
1701: \sum_{j=0}^k t^j\I_\Z^{k-j}$ and
1702: \begin{equation} \label{five}
1703: \frac{\I_{\!Z'\subset\X}^k}{t\I_{\!Z'\subset\X}^k}=
1704: \frac{\sum_{j=0}^k t^j\I_\Z^{k-j}}{\sum_{j=0}^k t^{j+1}\I_\Z^{k-j}}=
1705: \sum_{j=0}^k\frac{t^j\I_\Z^{k-j}}{t^j\I_\Z^{k-j}\ \cap\ \sum_{i=0}^k
1706: t^{i+1}\I_\Z^{k-i}}\,,
1707: \end{equation}
1708: where we use the fact that in an abelian category, for $A,B,C\subset V$
1709: with $C\subset A+B$, we have $(A+B)/C=A/(A\cap C)+B/(B\cap C)$ in $V/C$.
1710: We claim that
1711: \begin{equation} \label{three}
1712: t^j\I_\Z^{k-j}\ \cap\ \sum_{i=0}^kt^{i+1}\I_\Z^{k-i}=
1713: t^j\I_\Z^{k-j+1}+t^{j+1}\I_\Z^{k-j},
1714: \end{equation}
1715: except for $j=0$ when the right hand side becomes $t\I_\Z^k$.
1716: The inclusion $\supseteq$ is clear. For $\subseteq$, consider the left
1717: hand side:
1718: \begin{eqnarray*}
1719: t^j\I_\Z^{k-j} \!\!&\cap&\!\! \left((t\I_\Z^k+\ldots+t^j\I_\Z^{k-j+1})+
1720: (t^{j+1}\I_\Z^{k-j}+\ldots+(t^{k+1}))\right) \\
1721: \subseteq t^j\I_\Z^{k-j} \!\!&\cap&\!\! \left(t\I_\Z^{k-j+1}+(t^{j+1})\right).
1722: \end{eqnarray*}
1723: An element of this can be written as $t^jf=tg+t^{j+1}h$, that is
1724: $t^{j-1}f-g=t^jh$, where $f\in\I_\Z^{k-j},\ g\in\I_\Z^{k-j+1}$
1725: and so $t^{j-1}f-g\in\I_\Z^{k-j}$.
1726: So $t^jh\in(t^j)\cap\I_\Z^{k-j}$, which by (\ref{one}) is $t^j\I_\Z^{k-j}$.
1727: Thus $h$ may be taken to lie in $\I_\Z^{k-j}$. Similarly
1728: $g\in(t^{j-1})\cap\I_\Z^{k-j+1}=t^{j-1}\I_\Z^{k-j+1}$.
1729: Therefore $t^jf=tg+t^{j+1}h\in t^j\I_\Z^{k-j+1}+t^{j+1}\I_\Z^{k-j}$, proving
1730: the inclusion.
1731:
1732: So by (\ref{three}) and (\ref{four}), equation (\ref{five}) has become
1733: $$
1734: \frac{\I_{\!Z'\subset\X}^k}{t\I_{\!Z'\subset\X}^k}=\frac{\I_\Z^k}{t\I_\Z^k}+
1735: \sum_{j=0}^k t^j \frac{\I_\Z^{k-j}}{\I_\Z^{k-j+1}+t\I_\Z^{k-j}}
1736: =\I_{\!Z'}^k+\sum_{j=0}^k t^j\frac{\I_{\!Z'}^{k-j}}{\I_{\!Z'}^{k-j+1}}\,.
1737: $$
1738: To check that the sum is direct, intersect the $j$th numerator with the others:
1739: $$
1740: t^j\I_\Z^{k-j}\cap\sum_{p\ne j}t^p\I_\Z^{k-p}\subseteq t^j\I_\Z^{k-j}\cap
1741: \left(\I_\Z^{k-j+1}+(t^{j+1})\right).
1742: $$
1743: By the same methods as before this lies in $t^j\I_\Z^{k-j+1}+t^{j+1}\I_\Z^{k-j}$,
1744: which is the $j$th denominator, as required.
1745: \end{proof}
1746:
1747: To apply the above result inductively to the $\X^i$ requires the flatness
1748: of the
1749: thickenings $k(\overline{Z_i\times\C})$ in $\X^i$ of the proper transforms
1750: of the $Z_i\times\C\subset X\times\C$. This is only automatic for $k=1$,
1751: but also holds for arbitrary $k$ if the $Z_i$ and $X$ are all \emph{smooth},
1752: or, we shall show, if:
1753: \begin{eqnarray} \nonumber
1754: \text{\emph{$X$ is reduced, each $Z_{r-1}\subseteq\ldots\subseteq Z_0$ is
1755: a Cartier divisor in $X$, and any irreducible}} \hspace{-1cm} \\ \label{reduced}
1756: \text{\emph{component common to any pair $Z_i,\,Z_j$ has the same
1757: multiplicity in each.}} \quad
1758: \end{eqnarray}
1759: This odd looking condition is clearly satisfied if, for instance, each $Z_i$
1760: is reduced.
1761:
1762: \begin{prop} \label{flat}
1763: Suppose that $X$ and the $Z_i$ satisfy (\ref{reduced}).
1764: Then $j(\overline{Z_i\times\C})\subset \X^i$ is flat over $\C$ for each
1765: $j\in\mathbb N$.
1766: \end{prop}
1767:
1768: \begin{proof}
1769: Firstly, consider $\overline{j(Z_i\times\C)}$. Its ideal is defined by
1770: those functions which, on restriction to $t\ne0$, lie in $\I_{\!Z_i\times
1771: \C^\times}^j$. Since $t$ is invertible there, this implies that if $tf\in
1772: \I\_{\overline{j(Z_i\times\C)}}$ then $f\in\I\_{\overline{j(Z_i\times\C)}}$\,.
1773: Therefore the structure sheaf of
1774: $\overline{j(Z_i\times\C)}$ has no $t$-torsion. It is also flat away from
1775: $t=0$ (where it is $j(Z_i\times\C^\times)$); thus $\overline{j(Z_i\times\C)}$
1776: is automatically flat over $\C$.
1777:
1778: $\I\_{\overline{j(Z_i\times\C)}}\supseteq\I^j_{\overline{Z_i\times\C}}$;
1779: therefore to prove that $j(\overline{Z_i\times\C})$ is flat over $\C$ it
1780: is sufficient to prove the opposite inclusion
1781: \begin{equation} \label{inclusion}
1782: \I\_{\overline{j(Z_i\times\C)}}\ \subseteq\ \I^j_{\overline{Z_i\times\C}}\,.
1783: \end{equation}
1784:
1785: $\X^i$ is obtained by blowing up $\X^{i-1}$ in $Z_{i-1}^{(i-1)}$; we claim
1786: the complement
1787: $\X^i_\text{aff}$ of the proper transform $\overline{(\X^{i-1})\_0}$ of the
1788: central fibre in $\X^i$ is affine over $X\times\C$ with coordinate ring
1789: \vspace{3mm} \begin{equation} \vspace{-11mm} \label{coord} \end{equation}
1790: \begin{multline*}
1791: \!\!\sum_{a\_0,a\_1,\ldots,a\_{i-1}\ge0}\left(\!\frac{\I_0}{t^i}\!\right)^{\!a_0}
1792: \!.\!\left(\!\frac{\I_1}{t^{i-1}}\!\right)^{\!a_1}
1793: \!\ldots\!\left(\!\frac{\I_{i-1}}t\!\right)^{\!a\_{i-1}}\ = \\
1794: \O_X\otimes\C[t]\ +\ \frac{\I_{\!i-1}}t\ +\ \frac{\I_{\!i-2}+\I_{\!i-1}^2}{t^2}
1795: \ +\ \frac{\I_{\!i-3}+\I_{\!i-2}\I_{\!i-1}+\I_{\!i-1}^3}{t^3}\ +\ \ldots,
1796: \end{multline*}
1797: with its obvious ring structure. Each $\I_i$ is an $\O_X$-module, so the
1798: whole ring inherits an $\O_X\otimes\C[t]$-module structure, corresponding
1799: to the projection to $X\times\C$.
1800:
1801: There are many ways to see this. One is to note that, away from $\overline
1802: {(\X^{i-1})\_0}$\,, the map to $\B_{I_i}(X\times\C)$ of Theorem
1803: \ref{pr} is an isomorphism. This is because, in the notation of that section,
1804: sections of $\O(-E_1-\ldots-E_i)$ over $X\times\C$ do not contract the
1805: exceptional divisor $E_i$ of the blow up of $\X^{i-1}$ in $Z_{i-1}^{(i-1)}$
1806: (as noted in the remarks following the proof of Corollary \ref{seshadriIr}).
1807: Let $E$ denote the exceptional divisor of
1808: $\B_{I_i}(X\times\C)$, and $s_E\in H^0(\O(E))$ the canonical section
1809: vanishing on $E$. For $k\gg0$, the sections of $\O(-kE)$ are the sections
1810: of $I_i^k/s_E^k$ (that is, pull back sections of $I_i^k$ from
1811: $X\times\C$ to $\B_{I_i}(X\times\C)$ and divide by $s_E^k$ to get a
1812: regular section of $\O(-kE)$). $t^i\in I_i=\I_0+t\I_1+\ldots+
1813: t^{i-1}\I_{i-1}+(t^i)$
1814: defines the section $t^i/s_E$ of $\O(-E)$ which trivialises $\O(-E)$ over
1815: $\X^i_\text{aff}$ -- the complement of its zero locus $\overline{(\X^{i-1})\_0}$.
1816: Using this trivialisation identifies the functions on $\X^i_\text{aff}$ which
1817: have poles of order $\le k$ on $\overline{(\X^{i-1})\_0}$ with
1818: $$
1819: \frac{I_i^k}{t^{ik}}\ =\
1820: \left(\O+\frac{\I_{i-1}}{t}+\ldots+\frac{\I_0}{t^i}\right)^k;
1821: $$
1822: taking the limit as $k\to\infty$ gives the regular functions (\ref{coord}).
1823:
1824: Alternatively, we can work inductively with the $\X^j$. A similar analysis
1825: as above shows the coordinate ring of the complement of $\overline{\X_0}$
1826: in the blow up of a family $\X$ over $\C$ in an ideal $I+(t)$ is
1827: \begin{equation} \label{BI}
1828: \O_\X+\frac It+\frac{I^2}{t^2}+\frac{I^3}{t^3}+\ldots
1829: \end{equation}
1830: Thus we find the coordinate ring of $\X^1_\text{aff}=\X^1\backslash\,
1831: \overline{X\times\{0\}}$ is
1832: $$
1833: \O_X\otimes\C[t]\,+\,\frac{\I_0}t\,+\,\frac{\I_0^2}{t^2}\,+\,\frac{\I_0^3}{t^3}\,+\ldots
1834: $$
1835: Inside this the ideal of $\overline{Z_1\times\C}$ is $\I_1+\frac{\I_0}t+
1836: \frac{\I_0^2}{t^2}+\ldots$ (as this is the largest ideal that localises
1837: to $\I_1\otimes\C[t,t^{-1}]$ on $t\ne0$); applying (\ref{BI}) to this ideal
1838: $I$ gives the coordinate ring of $\X^2_\text{aff}=\X^2\backslash\,
1839: \overline{(\X^1)_0}$ as
1840: $$
1841: \O_X\otimes\C[t]\ +\ \frac{\I_1}t\ +\ \frac{\I_1^2+\I_0}{t^2}\ +\
1842: \frac{\I_1^3+\I_0\I_1}{t^3}\ +\ \ldots
1843: $$
1844: But this is the $i=2$ case of (\ref{coord}), and inductively we recover it
1845: for all $i$. In (\ref{coord}) we have the ideal
1846: \begin{equation} \label{kZ}
1847: \I\_{\overline{j(Z_i\times\C)}}\ =\
1848: \sum_{a\_0,a\_1,\ldots,a\_{i-1}\ge0}\frac{\I_i^j\cap(\I_0^{a_0}.\I_1^{a_1}
1849: \ldots\I_{i-1}^{a\_{i-1}})}{t^{ia_0}.t^{(i-1)a_1}\ldots t^{a\_{i-1}}}\,,
1850: \end{equation}
1851: as this is the largest ideal that localises to $\I_i^j\otimes\C[t,t^{-1}]$
1852: on $t\ne0$. In the $j=1$ case, $\I_i\subseteq\I_0^{a_0}.\I_1^{a_1}
1853: \ldots\I_{i-1}^{a\_{i-1}}$ unless $a_j=0\ \forall j$,
1854: so $\I\_{\overline{Z_i\times\C}}$ differs from (\ref{coord}) only in the
1855: first term $\I_i\subset\O_X\otimes\C[t]$:
1856: \begin{equation} \label{Z}
1857: \I\_{\overline{Z_i\times\C}}\ =\ \I_i\ +\!\sum_{a\_j\ge0,\ \sum_{j=0}^{i-1}a_j\ge1}
1858: \frac{\I_0^{a_0}.\I_1^{a_1}
1859: \ldots\I_{i-1}^{a\_{i-1}}}{t^{ia_0}.t^{(i-1)a_1}\ldots t^{a\_{i-1}}}\,.
1860: \end{equation}
1861: By (\ref{inclusion}) we are left with showing that each term of (\ref{kZ}) is
1862: contained in the $j$th power of (\ref{Z}):
1863: \begin{equation} \label{Zk}
1864: \I^j_{\overline{Z_i\times\C}}\ =\sum_{p=0}^j\left(\sum_{a\_j\ge0,\
1865: \sum_{j=0}^{i-1}a_j\ge p}\!\!\I_i^{j-p}\cdot\frac{\I_0^{a_0}.\I_1^{a_1}
1866: \ldots\I_{i-1}^{a\_{i-1}}}{t^{ia_0}.t^{(i-1)a_1}\ldots t^{a\_{i-1}}}\right).
1867: \end{equation}
1868:
1869: We now work locally, where the conditions (\ref{reduced}) on $Z_i\subseteq
1870: Z_{i-1}\subseteq\ldots\subseteq Z_0$ imply that
1871: $\I_i=(f_i)$ and $\I_j=(g_jf_i),\ j\le i-1$ for some $g_j$
1872: \emph{which do not divide} $f_i$. Therefore,
1873: for any $a_j\ge0$ with $p:=\sum_{j=0}^{i-1}a_j\le j$, we have
1874: \begin{multline*}
1875: \I_i^j\cap(\I_0^{a_0}\ldots\I_{i-1}^{a\_{i-1}})=
1876: (f_i^j)\cap(f_i^p.g_0^{a_0}\ldots g_{i-1}^{a\_{i-1}})
1877: =(f_i^j.g_0^{a_0}\ldots g_{i-1}^{a\_{i-1}}) \\
1878: =(f_i^{j-p}.f_i^p.g_0^{a_0}\ldots g_{i-1}^{a\_{i-1}})
1879: =\I_i^{p-a}.\I_0^{a_0}\ldots\I_{i-1}^{a\_{i-1}},
1880: \end{multline*}
1881: using the fact the $\O_X$ is
1882: torsion-free. This gives the desired inclusion.
1883: \end{proof}
1884:
1885: To apply Proposition \ref{flat} will involve replacing $X$ by a blow up on
1886: which the pullbacks of the $Z_i$ are divisors. Pulling back the polarisation,
1887: we find we are forced to work with a semi-ample line bundle. To this end,
1888: for any $Z\subset X$ and \emph{semi-ample} $L\to X$, we define
1889: \begin{equation} \label{wkZ}
1890: w_k(Z):=\sum_{j=1}^{ck}h^0_X(L^k\otimes\I_{\!Z}^j)-ckh^0_X(L^k)
1891: \end{equation}
1892: (cf. (\ref{degen}) and (\ref{ss})) for any $c\in\Q$ such that $L^k\otimes\I_{\!Z}^{ck}$
1893: is saturated by its global sections. This equals
1894: \begin{equation} \label{wkchi}
1895: w_k(Z)=\sum_{j=1}^{ck}\chi\_X(L^k\otimes\I_{\!Z}^j)-ck\chi\_X(L^k)+O(k^n)
1896: \qquad (O(k^{n-1})\text{ for $L$ ample}),
1897: \end{equation}
1898: by the following generalisation of Lemma \ref{error} applied to $Z$ and
1899: $Z=\emptyset$.
1900:
1901: \begin{lem} \label{wkerror}
1902: For $Z\subset X,\ L\to X$ semi-ample, and $L^k\otimes\I_{\!Z}^{ck}$ saturated
1903: by global sections for $k\gg0$,
1904: $\sum_{j=1}^{ck}h^{\ge1}(L^k\otimes\I_{\!Z}^j)=O(k^n)$.
1905: \end{lem}
1906:
1907: \begin{proof}
1908: Consider the
1909: blow up of $X\times\PP^1$ in $Z\times\{0\}$ with exceptional divisor $P$
1910: and line bundle $\O_{\PP^1}(ck)\boxtimes L^k-ckP$. This is semi-ample, generated
1911: by $t^{ck}\boxtimes H^0_X(L^k)\,+\,s^{ck}\boxtimes H^0_X(L^k\otimes\I_{\!Z}^{ck})$,
1912: where $s,\,t\in H^0(\O_{\PP^1}(1))$ are the sections vanishing at $\infty,
1913: \,0\in\PP^1$ respectively. So its higher
1914: cohomology has total dimension bounded by $O(k^n)$. But by pushing down
1915: first to $X\times\PP^1$, then to $X$, this higher cohomology can be computed
1916: as $\bigoplus_{j=0}^{ck}s^{ck-j}t^j\boxtimes H^{\ge1}_X(L^k\otimes\I_{\!Z}^{ck-j})$,
1917: of total dimension $\sum_{j=1}^{ck}h^{\ge1}(L^k\otimes\I_{\!Z}^j)$.
1918: \end{proof}
1919:
1920: Fix $(\X,L)\to\C$ such that for $k\gg0$, $(\X,L^k)$ is a semi test configuration
1921: for $(X,L^k)$. Given $\Z\subset\X$ a $\C^\times$-invariant
1922: subscheme, denote its general fibre by $Z\subset X$ and central fibre $Z'\subset
1923: \X_0$. Let $(\B_{Z'}(\X),\L_c)\rt{\pi}(\X,L)$
1924: denote the blow up of $\X$ along $Z'$, with exceptional divisor $E$ and
1925: line bundle $\L_c=\pi^*L-cE$. (As usual $c\in(0,\epsilon(Z'))$ or $c=\epsilon(Z')$
1926: if $L^k\otimes\I_{\!Z'}^{\epsilon(Z')k}$ is saturated by
1927: global sections for $k\gg0$.) Thus $\L_c$ is semi-ample by Proposition
1928: \ref{ample}, making $(\B_{Z'}(\X),\L_c^k)$ a semi test configuration for
1929: $(X,L^k)$ for $k\gg0$.
1930:
1931: Then we have the following generalisation of Theorems \ref{degen} and \ref{ss}.
1932:
1933: \begin{thm} \label{big}
1934: In the above situation, suppose that the thickenings $j\Z\subset\X$ are flat
1935: over $\C$ for all $j\in\mathbb N$, and the $\C^\times$-action on
1936: $H^0_{\X_0}\!(L^k)$ has only weights
1937: which lie between $-Ck$ and $0$, for some $C>0$. Then
1938: $$
1939: w\big(H^0_{\X_0}(L^k)\big)=w\big(H^0_\X(L^k)\big/tH^0_\X(L^k)\big)+O(k^n),
1940: $$
1941: $H^0_{(\B_{Z'}(\X))\_0}\!(\L_c^k)$ has only weights which lie between $-(C+c)k$
1942: and $0$, and
1943: $$
1944: w\big(H^0_{(\B_{Z'}(\X))\_0}\!(\L_c^k)\big)=w\big(H^0_{\X_0}(L^k)\big)+w_k(Z)
1945: +O(k^n).
1946: $$
1947: If $L$ is ample then the first $O(k^n)$ correction vanishes, and if in
1948: addition either $c<\epsilon(Z')$ or the $\C^\times$-action on
1949: $H^0_{\X_0}(L^k)$ is trivial then the second correction is $O(k^{n-1})$.
1950: \end{thm}
1951:
1952: \begin{proof}
1953: $(\Y=\Proj\bigoplus_kH^0_\X(L^k),\O_\Y(1))$ is a polarised
1954: family over $\C$ with general fibre $(\Proj\bigoplus_kH^0_X(L^k),\O(1))$.
1955: It is flat because $H^0_\X(L^k)$ has no $t$-torsion, by the flatness of
1956: $\X$. For $k\gg0$, by the flatness of $\Y$ and ampleness of
1957: $\O_\Y(1)$, the central fibre has sections $H^0_{\Y_0}(\O(k))=
1958: H^0_\Y(\O(k))\big/tH^0_\Y(\O(k))=H^0_\X(L^k)\big/tH^0_\X(L^k)$. Again by
1959: flatness and ampleness, this has the same dimension
1960: as for the general fibre, which is $h^0_X(L^k)$, which equals
1961: $\chi\_X(L^k)+O(k^{n-1})$ by semi-ampleness. In turn by flatness and
1962: semi-ampleness, this equals $\chi\_{\X_0}(L^k)+O(k^{n-1})=
1963: h^0_{\X_0}(L^k)+O(k^{n-1})$. Therefore the inclusion
1964: $$
1965: \frac{H^0_\X(L^k)}{tH^0_\X(L^k)}\ \subseteq\ H^0_{\X_0}(L^k)
1966: $$
1967: has codimension $O(k^{n-1})$. This inclusion is $\C^\times$-equivariant,
1968: and all weights on the right hand side lie between $-Ck$ and $0$ by assumption.
1969: Therefore the total weights on the two vector spaces differ by at most
1970: $O(k^n)$, as claimed. If $L$ is ample then by cohomology vanishing
1971: $H^0_{\X_0}(L^k)=H^0_\X(L^k)\big/tH^0_\X(L^k)$ and the correction vanishes.
1972: \\
1973:
1974: To streamline notation we fix the convention in the proof of the second
1975: result that
1976: $ck+1=\infty$ so that, for instance, $\I_{\!Z'}^{ck}/\I_{\!Z'}^{ck+1}$ means
1977: $\I_{\!Z'}^{ck}$ and we can deal with all of the terms in Proposition \ref{central}
1978: uniformly. For $k\gg0$, this gives
1979: \begin{equation} \label{HBl}
1980: H^0_{(\B_{Z'}(\X))\_0}\!(\L_c^k)=H^0_{\X_0}\!\bigg(L^k\otimes
1981: \frac{\I_{\!Z'\subset\X}^{ck}}{t\I_{\!Z'\subset\X}^{ck}}\bigg)=
1982: \bigoplus_{i=0}^{ck}\ t^i\,H^0_{\X_0}\!\left(\!L^k\otimes
1983: \frac{\I_{\!Z'}^{ck-i}}{\I_{\!Z'}^{ck-i+1}}\right).
1984: \end{equation}
1985: Since the weight on $t^i$ is $-i$ and lies between $-ck$ and $0$, this
1986: shows the weights on $H^0_{(\B_{Z'}(\X))\_0}\!(\L_c^k)$ indeed lie between
1987: $-(C+c)k$ and $0$. Included in this $\C^\times$-module is
1988: \begin{equation} \label{split2}
1989: \bigoplus_{i=0}^{ck}\ t^i\frac{H^0_{\X_0}\!(L^k\otimes
1990: \I_{\!Z'}^{ck-i})}{H^0_{\X_0}\!(L^k\otimes\I_{\!Z'}^{ck-i+1})}\,,
1991: \end{equation}
1992: of dimension $h^0_{\X_0}\!(L^k)$, which we have already noted above
1993: is the same as $h^0_X(L^k)$ to $O(k^{n-1})$. The same working (applied
1994: to $(\B_{Z'}(\X),\L_c)$ instead of $(\X,L)$) shows that
1995: $h^0_{(\B_{Z'}(\X))\_0}\!(\L_c^k)$ also equals $h^0_X(L^k)+O(k^{n-1})$.
1996: Thus the $-(C+c)k$-bound on weights means we can instead calculate the
1997: weight on (\ref{split2}) at the expense of an $O(k^n)$ error.
1998: If $L$ is ample and $c<\epsilon(Z')$ then $\L_c$ is also ample, so
1999: $h^0_{(\B_{Z'}(\X))\_0}\!(\L_c^k)=h^0_X(L^k)=h^0_{\X_0}\!(L^k)$ and we
2000: can calculate the weight on (\ref{split2}) without error.
2001: Finally if $c=\epsilon(Z')$ and $L$ is ample
2002: then as in Proposition \ref{error}, (\ref{HBl}) and (\ref{split2}) agree
2003: for all but $i=0,\ldots eN$ (independent of $k$) so if the $\C^\times$-action
2004: on $H^0_{\X_0}(L^k)$ is trivial then their weights differ by
2005: $\le\sum_{i=0}^{eN}ih^1(L^k\otimes\I_{\!Z'}^{ck-i+1})\le eN.O(k^{n-1})=O(k^{n-1})$.
2006:
2007: Define $V^0:=H^0_{\X_0}\!(L^k)$, and
2008: \begin{equation} \label{filt}
2009: V^p:=H^0_{\X_0}\!(L^k\otimes\I_{\!Z'}^p)\subseteq V^{p-1}\subseteq\ldots
2010: \subseteq V^0.
2011: \end{equation}
2012: Let $V^0=\bigoplus_{j=-Ck}^{\,0}V^{0,j}$ be its weight space decomposition.
2013: Given such a splitting $\bigoplus_jV^{0,j}$ of a vector space
2014: $V^0$, the generic subspace $V^p\subset V^0$ is not generated
2015: by the pieces $V^{p,j}:=V^p\cap V^{0,j}$; i.e.\ $V^p\supsetneq
2016: \bigoplus_jV^{p,j}$. But if $V^0$ has a $\C^\times$-action whose weight space
2017: decomposition is $\bigoplus_jV^{0,j}$, and each $V^p\subset V^0$ is
2018: $\C^\times$-invariant, then indeed
2019: \begin{equation} \label{filt1}
2020: V^p=\bigoplus_jt^jV^{p,j} \qquad\text{and}\qquad \frac{V^p}{V^{p+1}}\cong
2021: \bigoplus_jt^j\frac{V^{p,j}}{V^{p+1,j}}\,.
2022: \end{equation}
2023: This holds here since $\I_{\!Z'}^p$ is $\C^\times$-invariant.
2024:
2025: So the splitting $\bigoplus_{i=0}^{ck}t^iV^{k-i}\big/V^{k-i+1}$ (\ref{split2})
2026: further splits as
2027: \begin{equation} \label{split9}
2028: \bigoplus_{i,j}t^i\frac{V^{ck-i,j}}{V^{ck-i+1,j}}\,.
2029: \end{equation}
2030: Defining $h^{a,b}:=\dim V^{ck-a,b}/V^{ck-a+1,b}$, the weight on the determinant
2031: of (\ref{split9}) is $\sum_{i,j}(-i+j)h^{i,j}$.
2032:
2033: By (\ref{filt1}), $\sum_jh^{i,j}=\dim V^{ck-i}/\dim V^{ck-i+1}$, which by (\ref{filt})
2034: is $h^0_{\X_0}\!(L^k\otimes\I_{\!Z'}^{ck-i})-h^0_{\X_0}\!(L^k\otimes\I_{\!Z'}^{ck-i+1})$.
2035: So as in (\ref{noco}), $-\sum_{i=0}^{ck}i\sum_jh^{i,j}=\sum_{i=1}^{ck}h^0_{\X_0}
2036: (L^k\otimes\I_{\!Z'}^i)\,-ckh^0_{\X_0}\!(L^k)$.
2037:
2038: Similarly, the fact that $V^{i+1,j}\subset V^{i,j}\subset\ldots\subset V^{0,j}$
2039: filters $V^{0,j}$ means that $\sum_ih^{i,j}=\dim V^{0,j}$. Therefore
2040: $$
2041: w\big(H^0_{(\B_{Z'}(\X))\_0}\!(\L_c^k)\big)=\sum_{i,j}(-i+j)h^{i,j}=
2042: -\sum_{i=0}^{ck}i\bigg(\!\sum_jh^{i,j}\bigg)
2043: +\!\sum_{j=-Ck}^0j\bigg(\!\sum_ih^{i,j}\bigg)+O(k^n)
2044: $$
2045: equals
2046: $$
2047: \sum_{i=1}^{ck}h^0_{\X_0}\!(L^k\otimes\I_{\!Z'}^{ck-i})-ckh^0_{\X_0}\!(L^k)\
2048: +\sum_{j=-Ck}^0j\dim V^{0,j}+O(k^n).
2049: $$
2050: We can replace $h^0$ by $\chi$ at the expense of $O(k^n)$ by Lemma \ref{wkerror}
2051: (and at the expense of $O(k^{n-1})$, by Proposition \ref{error}, if $L$
2052: is ample). But $\I_{\!Z'}^j\subset\O_{\X_0}$ sits in a flat family with
2053: $\I_{\!Z}^j\subset\O_X$, in which Euler characteristics are preserved, so the above
2054: weight is
2055: \begin{multline*}
2056: \sum_{j=1}^{ck}\chi\_X(L^k\otimes\I_{\!Z}^j)-ckh^0_X(L^k)+w\big(
2057: H^0_{\X_0}(L^k)\big)+O(k^n) \\
2058: =w_k(Z)+w\big(H^0_{\X_0}(L^k)\big)+O(k^n),
2059: \end{multline*}
2060: by (\ref{wkchi}). And if $L$ is ample and if either $c<\epsilon(Z')$ or the
2061: $\C^\times$-action on $H^0_{\X_0}(L^k)$ is trivial then the correction
2062: is $O(k^{n-1})$.
2063: \end{proof}
2064:
2065:
2066: \section{Towards a converse} \label{converse}
2067:
2068: To apply Theorem \ref{big} to Corollary \ref{ref} to express weights of test
2069: configurations (\ref{any}) as sums of $w_k(Z)$s requires flatness of the
2070: thickenings $j(\overline{Z_i\times\C})$ in $\X^i$. To help achieve this by
2071: applying
2072: Proposition \ref{flat} we assume that $X$ is reduced and pass to a blow up
2073: $p\colon\widehat X\to X$ on which the (pullbacks of the) $Z_i$ are Cartier
2074: divisors $D_i$: $p^*\I_{Z_i}=\O(-D_i)$. In fact
2075: by Hironaka's resolution of singularities we may take the $D_i$ to have simple
2076: normal crossing (snc) support. That
2077: is there are smooth reduced divisors $\{F_j\}\subset\widehat X$ such that
2078: $F=\cup_jF_j$ has simple normal crossings and $D_i=\sum_jm_{ij}F_j$ for each
2079: $i$ and some nonnegative integers $m_{ij}$. We pull $L$ back
2080: to $\widehat X$, and construct the families $(\XX^s,L_s)\to\C$
2081: as before, using $D_s\subset\widehat X$ in place
2082: of $Z_s\subset X$. There are equivariant surjective maps $\XX^s\to\X^s$
2083: which identify (by pullback) the $L_s$
2084: line bundles, defined by pulling back sections of $L^k\otimes\I_i^j$ on $X$
2085: to sections of $L^k\otimes\O(-jm_iD_i)$ on $\widehat X$.
2086:
2087: \begin{thm} \label{bigger}
2088: Suppose that $X$ is normal, and fix an arbitrary test configuration $(\Y,\O_\Y(1))$
2089: (\ref{any}) with associated subschemes $Z_0\subseteq\ldots\subseteq Z_{r-1}\subset
2090: X$ (\ref{ideals}). Suppose that there is a resolution $D_0\subseteq\ldots\subseteq
2091: D_{r-1}\subset\widehat X$ with divisors $D_i$ satisfying condition
2092: (\ref{reduced}). (For instance if the connected components of the snc divisors
2093: $D_i$ have the same multiplicities locally, i.e.\ $m_{ij}\in\{0,m\}$ for some
2094: locally constant $m$ and all $i$. E.g. if the $D_i$ are all reduced, which
2095: is the $m=1$ case.) Then
2096: $$
2097: w(H^0_{\Y_0}\!(\O_\Y(k)))=\sum_{i=0}^{r-1}w_k(Z_i)+O(k^n).
2098: $$
2099: \end{thm}
2100:
2101: \begin{proof}
2102: $(\XX^r,L_r)$ dominates $(\X^r,L_r)$ which in turn dominates
2103: $(\B_{I_r}(X\times\C),\L_1)$ and so $(\Y,\O_\Y(1))$,
2104: and the space of sections of the $k$th power of the line bundle on the
2105: general fibre is $h^0_X(L^k)$ for all three, by the normality of $X$.
2106: Therefore by Proposition \ref{stein}
2107: and the first part of Theorem \ref{big} we may calculate $w(H^0_{(\widehat
2108: \X^r)\_0}\!(L_r^k))$ at the expense of an $O(k^n)$ error.
2109:
2110: We apply Theorem \ref{big} inductively to $\X=\XX^i$ with $c=1$ (which is
2111: $\le\epsilon(Z_i)=\epsilon(D_i)=\epsilon(D^{(i)}_i)$ by Corollary
2112: \ref{seshadriIr}) and $\Z=\overline{D_i\times\C}$ with central fibre
2113: $D_i^{(i)}$. The condition (\ref{reduced}) on the $D_i$ guarantees that each
2114: $j(\overline{D_i\times\C})$ is flat over $\C$ by Proposition \ref{flat}.
2115:
2116: The induction starts with $\widehat X\times\C$, for which all weights are
2117: zero so trivially satisfy the $-Ck$ bound. Theorem \ref{big} then ensures
2118: the induction continues to compute the weight as
2119: $$
2120: \sum_{i=0}^{r-1}w_k(D_i)+O(k^n).
2121: $$
2122: Since $X$ is normal, all sections of $p^*L$ on $\widehat X$ are pullbacks
2123: from $X:\ H^0_{\widehat X}(L^k)\cong H^0_X(L^k)$. Thus the same is true of
2124: those sections vanishing on the pullback of $Z_i:\ H^0_{\widehat X}
2125: (L^k\otimes\I_{\!D_i}^j)\cong H^0_X(L^k\otimes\I_{\!Z_i}^j)$. Thus $w_k(D_i)
2126: =w_k(Z_i)$ by their definition (\ref{wkZ}), as required.
2127: \end{proof}
2128:
2129: In particular we can now calculate the leading order term of the weight if
2130: the snc divisors
2131: $D_i$ in a resolution of singularities $\widehat X$ of $(X,Z_i)$ are
2132: reduced. The next case we consider is when they have multiplicities which
2133: can vary with $i$ but are still locally constant over their snc support.
2134: Here the relevant flatness result does not hold, but we will find that it
2135: does after performing a basechange and normalisation.
2136:
2137: So we consider the case when $I_r$ (\ref{ideals}) is locally of the form
2138: \begin{equation} \label{ideal}
2139: \I_D^{p_1}+t\I_D^{p_2}+\ldots+(t^r),
2140: \end{equation}
2141: for some reduced snc divisor $D$. These $p_i$ may
2142: vary over the different connected components of $D$, but since the total
2143: weight is a sum over contributions from each connected component we can
2144: calculate the weight of each separately.
2145:
2146: Pick a local function $z$ generating the ideal $\I_D$, so that (\ref{ideal})
2147: is $(z^{p_1})+t(z^{p_2})+\ldots+(t^r)$. In the concave hull of the points
2148: $(p_i,i),\,i=1,\ldots,r$ in the $(z,t)$-plane, we choose extremal vertices
2149: $(k_i,\rho_i),\,i=1,\ldots,l$ so that they form a
2150: concave set with the same concave hull (and $(k_1,\rho_1)=(p_1,0),\ (k_l,\rho_l)
2151: =(0,r)$). This defines a new ideal
2152: \begin{equation} \label{ideal2}
2153: (z^{k_1})+t^{\rho\_2}(z^{k_2})+\ldots+(t^{\rho\_l}),
2154: \end{equation}
2155: with the same integral closure as (\ref{ideal}) since in this situation
2156: taking integral closures corresponds to taking concave hulls, and an ideal
2157: saturates its integral closure. In the next theorem we decorate $w_k(D)$
2158: (\ref{wkZ}) as $w_k(D,c)$ with the value $c$ that determines the line bundle
2159: $\L_c=L-cE$ on the blow up.
2160:
2161: \begin{thm} \label{divs}
2162: Set $m_i=\frac{\rho_{i+1}-\rho_i}{k_i-k_{i+1}},\ i=1,\ldots l$, and $m_0=0$.
2163: If $L$ is ample, then the total weight of the blow up in the ideal (\ref{ideal})
2164: is the sum over the connected components $D$ of
2165: $$
2166: -\sum_{i=1}^{l-1}(m_i-m_{i-1})w_k(D,k_i)-ak^n+O(k^{n-1}),
2167: $$
2168: for some $a\ge0$. If $L$ is semi-ample, the above expression is correct
2169: to $O(k^n)$.
2170: \end{thm}
2171:
2172: \begin{proof}
2173: We start by proving the weaker estimate for $L$ semi-ample.
2174: First blow up $X\times\C$ in $D\times\{0\}$ (i.e.\ locally in the ideal $(z)+(t)$),
2175: then in $D'$ (recall this is
2176: the central fibre of $\overline{D\times\C}$ in the blow up), then $D''=D^{(2)}$,
2177: etc. up to $D^{(j-1)}$. As in Theorem \ref{pr} we denote by $E_i$ the
2178: pullback of the exceptional divisor of the $i$th blow up, and by $s_i$
2179: the canonical section of $\O(E_i)$ vanishing on $E_i$. Then we claim
2180: that the pushdown of $\O(-p_1E_1-\ldots-p_jE_j),\ p_1\ge p_2\ge\ldots\ge
2181: p_j$, to $X\times\C$ is the ideal
2182: \begin{multline} \label{induction}
2183: (z^{p_1})+\ldots+t^{p_1+p_2-2p_2}(z^{p_2})+\ldots+t^{p_1+p_2+p_3-3p_3}(z^{p_3})
2184: \\ +\ldots+\qquad\cdots\qquad+\ldots+
2185: t^{p_1+\ldots+p_j-jp_j}(z^{p_j})+(t^{p_1+\ldots+p_j}).
2186: \end{multline}
2187: Here ``$\ldots$" means ``all convex combinations in between", i.e.\ the integral
2188: closure of the ideal generated by the named terms.
2189: (So $(f)+\ldots+(g)$ includes all monomials $h$ such that there exist $\lambda,\mu
2190: \in\mathbb N$ with $h^{\lambda+\mu}=f^\lambda g^\mu$.) That is, we claim
2191: the sections of $\O(-p_1E_1-\ldots-p_jE_j)$ over $X\times\C$ are $s_1^{-p_1}\ldots
2192: s_j^{-p_j}$ times by the above ideal. This is standard but
2193: fiddly to prove, and is best done by Newton diagram. We give
2194: an unenlightening proof; the reader is advised to skip straight to the
2195: Newton diagram (Figure \ref{diag}) for the special case below.
2196:
2197: We prove (\ref{induction}) inductively alongside the claim that the complement
2198: of the divisor $t/s_j=0$ (which is the proper transform, in the $j$th blow
2199: up, of the central fibre of the $(j-1)$th blow up) is affine over $X\times\C$
2200: with coordinate ring
2201: \begin{equation} \label{affcoord}
2202: \O\_{\!X\times\C}\left[\frac z{t^j}\right].
2203: \end{equation}
2204: (In particular, at a smooth point of $D$ where $z$ is a local coordinate
2205: in an analytic coordinate system $(y_1,\ldots y_{n-1},z)$ for $X$, we see
2206: that this part of the $j$th blow up is $\Spec\C[y_1,\ldots,y_{n-1},z,t,z/t^j]
2207: =\Spec\C[y_1,\ldots,y_{n-1},Z,t]$, where $Z=z/t^j$, and so is locally
2208: isomorphic to $X\times\C$ with the proper transform of $(z^k=0)$ being $(Z^k=0)$.)
2209:
2210: For the first blow up, the sections of $O(-kE_1)$ are $(z,t)^ks_1^{-k}$ (there
2211: are no more because the exceptional divisor is a $\PP^1$-bundle over $D$).
2212: This is in agreement with (\ref{induction}),
2213: and the proper transform of the central fibre is $t/s_1=0$. On its complement,
2214: $t^k/s_1^k$ trivialises $O(-kE_1)$; dividing by it identifies the sections
2215: $((z)+(t))^ks_1^{-k}$ with $\O\_{\!X\times\C}.((z/t)^k+(z/t)^{k-1}+\ldots+(z/t)+1)$.
2216: Taking the limit as $k\to\infty$ gives the coordinate ring $\O\_{\!X\times\C}
2217: \left[\frac zt\right]$ claimed (\ref{affcoord}). So the induction starts
2218: at $j=1$.
2219:
2220: At the $j$th stage the coordinate ring (\ref{affcoord}) shows that $D^{(j)}$
2221: has ideal $(Z,t)$, where $Z=z/t^j$. Let $\pi$ denote the $(j+1)$th blow
2222: up. By the first step of the induction, $\{t/s_{j+1}\ne0\}$ has coordinate
2223: ring augmented by $Z/t=z/t^{j+1}$;
2224: i.e.\ by induction $\O\_{\!X\times\C}\left[\frac z{t^j}\right]\left[\frac
2225: z{t^{j+1}}\right]=\O\_{\!X\times\C}\left[\frac z{t^{j+1}}\right]$.
2226:
2227: $\pi_*\O(-p\_{j+1}E_{j+1})$ is the ideal $((Z)+(t))^{p\_{j+1}}=
2228: t^{-jp\_{j+1}}(z^{p\_{j+1}})+\ldots+(t^{p\_{j+1}})$; i.e.\ the sections of
2229: $\O(-p\_{j+1}E_{j+1})$ over $\X^j$ are $s_{j+1}^{-p\_{j+1}}$ times by sections
2230: of this ideal. Multiplying by the trivialising section
2231: $t^{p_1}s_1^{-p_1}\ldots t^{p_j}s_j^{-p_j}$ of $\O(-p_1E_1-\ldots-p_jE_j)$
2232: shows that over our affine piece,
2233: $$
2234: \pi_*\O(-p_1E_1-\ldots-p\_{j+1}E_{j+1})=\big(\pi_*\O(-p\_{j+1}E_{j+1})\big)\otimes
2235: \O(-p_1E_1-\ldots-p_jE_j)
2236: $$
2237: is the ideal
2238: $$
2239: \big[t^{p_1+\ldots+p_j-jp\_{j+1}}(z^{p\_{j+1}})+\ldots+(t^{p_1+\ldots+p_j+p\_{j+1}})
2240: \big].s_1^{-p_1}\!\ldots s_j^{-p_j}\subseteq\O(-p_1E_1-\ldots-p_jE_j).
2241: $$
2242: Thus the sections of $\O(-p_1E_1-\ldots-p\_{j+1}E_{j+1})$ are the sections
2243: of $\O(-p_1E_1-\ldots-p_jE_j)$ which lie in the above ideal, i.e.\ the intersection
2244: of (\ref{induction}) with the ideal $t^{p_1+\ldots+p_j-jp\_{j+1}}(z^{p\_{j+1}})
2245: +\ldots+(t^{p_1+\ldots+p_j+p\_{j+1}})$. But this is (\ref{induction}) with
2246: $j$ replaced by $j+1$, completing the induction.
2247:
2248: We first assume that the $m_i$ are all integers. Then the ideal (\ref{ideal2})
2249: has integral closure of the form (\ref{induction}), on taking $p_1,\ldots,p_{m_1}$
2250: all equal
2251: \begin{figure}[h]
2252: \center{\input{newton.pstex_t} \caption{Newton diagram for the blow up in
2253: the ideal (\ref{ideal2}) \label{diag}}}
2254: \end{figure}
2255: to $k_1$, then $p_{m_1+1},\ldots,p_{m_2}$ all equal to $k_2$, and so
2256: on, up to $p_{m\_{N-1}+1},\ldots,p_{m\_N}$ all equal to $k_N$.
2257: This is illustrated in Figure \ref{diag}, the Newton diagram of the $(z,t)$
2258: plane, with the $m_i$s being (minus) the gradients of the bold lines.
2259: Taking the integral closure, i.e.\ including the monomials ``$\ldots$" in
2260: (\ref{induction}), corresponds to including all integral points
2261: both on and above the line to lie in the ideal. Replacing $z$ by $Z=z/t^j$
2262: and multiplying by the trivialising section $t^{p_1+\ldots+p_j}$ in the
2263: above working corresponds to the integral affine transformation that locally
2264: takes one corner of the bold line into the $(z,t)$-axes.
2265:
2266: So we may calculate the weight of this sequence of blow ups, $m_1$ times
2267: in the central fibre of $\overline{D\times\C}$ with weight $c=k_1$, then
2268: $(m_2-m_1)$ times in the central fibre of $\overline{D\times\C}$ with weight
2269: $c=k_2$, etc. (The weight here just means the coefficient of the exceptional
2270: divisor in the line bundle $L-cE$ we use, as usual.) The flatness criterion
2271: (\ref{reduced}) is trivially satisfied so that by Theorem \ref{big} we
2272: may calculate the weight to be that claimed. This differs from the weight
2273: of the blow up in (\ref{ideal}) by a $-ak^n+O(k^{n-1})$ correction by Proposition
2274: \ref{stein} since the blow up in the integral closure (\ref{ideal2})
2275: is the normalisation of the blow up in (\ref{ideal}).
2276:
2277: If $L$ is ample then we can improve the estimate. In the first blow up,
2278: $\X=X\times\C$ is the trivial product configuration, so the
2279: $\C^\times$-action on $H^0_{\X_0}(L^k)$ is trivial and we can use the better
2280: estimate of Theorem \ref{big}. Next we group the first $m_1$ blow ups together
2281: as one blow up in $((z)+(t^{m_1}))^{k_1}$; this has the advantage
2282: that we do no blowing down (in fact the previous $m_1$ blow ups blow down
2283: to this). Since this blow up is the $t\mapsto t^{m_1}$ basechange of the
2284: blow up in the ideal $((z)+(t))$ (with $c=k_1$), it has weight $m_1w_k(D,k_1)$,
2285: the same as the sum of the $m_1$ blow ups we performed above. Similarly we
2286: group the next
2287: $m_2$ blow ups together, blowing up with $c=k_2$ in the ideal generated
2288: by $t^{m_2}$ and the ideal of $\overline{D\times\C}$, and calculate its
2289: weight as the $t\mapsto t^{m_2}$ basechange of a blow up we already know.
2290: Inductively we end up with the same formula for the weights, but with
2291: the added $O(k^{n-1})$ accuracy of Theorem \ref{big} coming from the fact
2292: that (after the first blow up) there are no blow downs, so $c$ is less
2293: than the Seshadri constant of the relevant $D^{(i)}$ at each stage.
2294:
2295: Finally, if the $m_i$ are not integers, we simple replace $t$ by $t^M$
2296: in (\ref{ideal}), i.e.\ we basechange our test configuration, where $M$
2297: clears the denominators of all the $m_i$. This replaces all the $m_i$ by
2298: the integers $Mm_i$ while multiplying the $\C^\times$-weight by $M$.
2299: Substituting the $Mm_i$ into our formula for the weights gives the weight
2300: of this new test configuration with a $-ak^n+O(k^{n-1})$ correction (coming
2301: from taking
2302: the integral closure of this new ideal; replacing the test configuration
2303: with its normalisation and using Proposition \ref{stein}); dividing by $M$
2304: gives the weight of the original test configuration for any $m_i$.
2305: \end{proof}
2306:
2307: This just leaves the case of where the divisors $F_j$ in the snc divisors
2308: $D_i$ in a resolution $\widehat X$ intersect with differing multiplicities
2309: (the simplest example
2310: being $\I_{D_0}=(x^2y)$ and $\I_{D_1}=(x)$ locally). This of course cannot
2311: happen for curves, so we have
2312:
2313: \begin{cor} \label{Kslopecurve}
2314: A smooth curve $(X,L)$ is K-(semi/poly)stable if and only if it is slope
2315: (semi/poly)stable.
2316: \end{cor}
2317:
2318: \begin{proof}
2319: Smooth curves are normal, and for the resolution of singularities $\widehat
2320: X$ we of course take $X$ itself, so we can apply the stronger form of Theorem
2321: \ref{divs}. Since the Donaldson-Futaki invariant only uses the coefficients
2322: $b_0$ and $b_1$ of $w(k)=b_0k^{n+1}+b_1k^n+O(k^{n-1})$, this implies that
2323: the Futaki
2324: invariant of an arbitrary test configuration (\ref{any}) is $\ge$ a positive
2325: linear combination
2326: of Futaki invariants of (the deformation to the normal cone of) subschemes.
2327: Slope stability implies that these are all positive.
2328: \end{proof}
2329:
2330: This result makes it trivial to understand K-stability for smooth curves;
2331: see Theorem \ref{smoothcurvesareKstable}.
2332:
2333:
2334: \section{Chow stability} \label{sec:chow}
2335:
2336: Mumford's notion of Chow (semi)stability \cite{Mu} of $(X,\O_X(1))\subseteq\PP^N$,
2337: for \emph{fixed} $N$, is the simplest form of stability to calculate (as
2338: opposed to \emph{asymptotic} Chow stability (\ref{chow}), which is second
2339: only to asymptotic Hilbert stability in difficulty). It is also
2340: useful in algebro-geometric applications since it is a genuine GIT
2341: notion giving projective (and so proper and separated) moduli spaces.
2342:
2343: For any subscheme $Z\subseteq X$, we define the polynomial
2344: $a_0(x)$ by $\chi\_X(\I_{\!Z}^{xk}(k))=a_0(x)k^n+a_1(x)k^{n-1}+\ldots$ for
2345: $k\gg x^{-1}>0$ as before. We also define $a_0$ by $h^0(\O_X(k))=a_0k^n+
2346: a_1k^{n-1}+\ldots$\,; by (\ref{a0}), $a_0=a_0(0)$.
2347:
2348: For any subscheme $Z\subseteq X\subseteq\PP^N$ and \emph{integer}
2349: $0<c\le\epsilon(Z)$, we define the \emph{Chow slope} of $\I_Z$ to be
2350: $$
2351: Ch_c(\I_Z):=\frac{\sum_{i=1}^ch^0_{\PP^N}(\I_{\!Z}^i(1))}{\int_0^ca_0(x)dx}\,,\qquad
2352: Ch(\I_Z)=\max_{\mathbb N\ni c\le\epsilon(Z)}\big(Ch_c(\I_Z)\big)\ \in\,[-\infty,\infty].
2353: $$
2354: (Here we define max of the empty set to be $-\infty$, and division by 0
2355: to give $\infty$.) Setting $Z=\emptyset$ gives $Ch(X)=Ch(\O_X)$ as
2356: $$
2357: Ch(X):=\frac{h^0_{\PP^N}(\O(1))}{a_0}=\frac{N+1}{a_0}\,.
2358: $$
2359: If $X\into\PP^N$ is the Kodaira embedding of $X$ in $\PP(H^0(\O_X(1))^*)$,
2360: i.e.\ if $H^0_{\PP^N}(\O(1))\to H^0_X(\O(1))$ is an isomorphism, then
2361: $h^0_{\PP^N}(\I_Z(1))=h^0_X(\I_Z(1))$ for all $Z\subseteq X$. This can
2362: be arranged, for fixed $(X,L)$, by taking a sufficiently large multiple
2363: $L^k=:\O_X(1)$ of the polarisation and setting $\PP^N=\PP(H^0_X(L^k)^*)$.
2364: In this situation the above slopes can be written intrinsically in terms
2365: of $(X,\O_X(1))$ as
2366: $$
2367: Ch_c(\I_Z)=\frac{\sum_{i=1}^ch^0_X(\I^i_{\!Z}(1))}{\int_0^ca_0(x)dx}
2368: \qquad\text{and}\qquad Ch(X)=\frac{h^0_X(\O(1))}{a_0}\,.
2369: $$
2370:
2371: We say that $X\subset\PP^N$ is \emph{Chow slope stable} if $Ch(\I_Z)<Ch(X)$
2372: for all nonempty $Z\subseteq X$, and Chow slope semistable if $Ch(\I_Z)\le
2373: Ch(X)$. We have
2374: \begin{equation} \label{sss}
2375: Ch_c(\I_Z)<Ch(X) \quad\Longleftrightarrow\quad
2376: Ch(X)<Ch_c(\O_Z):=\frac{\sum_{i=1}^c\big(N+1-h^0(\I_{\!Z}^i(1))\big)}
2377: {\int_0^c\!\tilde a_0(x)dx}\,.
2378: \end{equation}
2379:
2380: \begin{thm} \label{thmchow}
2381: If $X\subseteq\PP^N$ is Chow (semi)stable then it is Chow slope (semi)stable.
2382: \end{thm}
2383:
2384: \begin{proof}
2385: Choose a basis of $H^0_{\PP^N}(\O(1))$ compatible with the filtration
2386: $H^0_{\PP^N}(\I_{\!Z}^c(1))\subseteq H^0_{\PP^N}(\I_{\!Z}^{c-1}(1))\subseteq\ldots
2387: \subseteq H^0_{\PP^N}(\I_Z(1))\subseteq H^0_{\PP^N}(\O(1))$, so that the
2388: first $p_i:=h^0_{\PP^N}(\I_{\!Z}^i(1))$ elements are contained in
2389: $H^0_{\PP^N}(\I_{\!Z}^i(1))$. The corresponding hyperplanes $H_1,H_2,\ldots,$
2390: $H_{N+1}$ and subschemes $Z_i:=
2391: X\cap H_1\cap\ldots\cap H_i$ therefore satisfy $Z_{p_i}\supseteq iZ$ (i.e.\
2392: $\I_{Z_{p_i}}\subseteq\I_{\!Z}^i$).
2393:
2394: Choose weights $\rho_1=0=\ldots=\rho_{p_c},\ \rho\_{p\_
2395: c+1}=1=\ldots=\rho_{p_{c-1}},\
2396: \ldots,\ \rho\_{p\_1+1}=c=\ldots=\rho_{p_0}$ so that ideal on $X\times\C$,
2397: \begin{equation} \label{IN}
2398: t^{\rho_1}\I_{Z_1}+t^{\rho_2}\I_{Z_2}+\ldots+t^{\rho\_{\!N}}\I_{Z_N}
2399: +(t^{\rho\_{\!N+1}})
2400: \end{equation}
2401: equals
2402: \begin{equation} \label{wobbly}
2403: I=\I_{Z_{p_c}}+t\I_{Z_{p_{c-1}}}+\ldots+t^{c-1}\I_{Z_{p_1}}+(t^c),
2404: \end{equation}
2405: which by construction is contained in the ideal
2406: \begin{equation} \label{straight}
2407: \I_{\!Z}^c+t\I_{\!Z}^{c-1}+\ldots+t^{c-1}\I_Z+(t^c)=(\I_Z+(t))^c.
2408: \end{equation}
2409:
2410: $X\subset\PP^N$ is Chow stable for the $\C^\times$-action which has
2411: weight $\rho_i$ on the $i$th vector of our basis of $H^0_{\PP^N}(\O(1))$.
2412: Mumford (\cite{Mu} Theorem 2.9) shows that this is equivalent to the inequality
2413: \begin{equation} \label{mummy}
2414: -a<\frac{a_0}{N+1}\sum_{i=1}^{N+1}\rho_i,
2415: \end{equation}
2416: where $a$ is the Chow weight of the blow up of $X\times\C$ in
2417: the ideal $I$ (\ref{wobbly}), i.e.\ $ak^{n+1}+O(k^n)$ is the total weight
2418: of the induced $\C^\times$-action on $H^0_{(\B_I(X\times\C))\_0}\!(L^k(-kE))$.
2419: (This is only one half of Mumford's result, the
2420: harder part being his computation of $a$ in terms of
2421: ideals on $X\times\C$, which we do not use. The reader wishing
2422: to compare our conventions with Mumford's should rewrite the above as
2423: $-(n+1)!\,a<(n!\,a_0)\frac{n+1}{N+1}\sum\rho_i$,
2424: replace $n$ with $r$, $N$ with $n$, and $\rho_i$ with $\rho\_{N-i}$.
2425: Finally the sign arises from our convention (\ref{convention}) that if
2426: $g$ acts on $V$ then the induced action on its functions $S^KV^*$ is by
2427: $(S^Kg^*)^{-1}$; Mumford calculates the weight of $S^Kg^*$.) Notice this
2428: is Theorem \ref{thm:instability} applied to (\ref{normalised}) on setting
2429: $r=1,\,h^0(L^r)=N+1$ and $w(1)=\sum_i\rho_i$.
2430:
2431: Since $c\ge\epsilon(Z)$, we know the Chow weight on the blow up in $(\I_Z+(t))^c$
2432: (Proposition \ref{integrals}). But $I$ is
2433: contained in $(\I_Z+(t))^c$ (\ref{straight}), so the weight on the latter
2434: is more negative (for instance Mumford's formula (\cite{Mu} Theorem 2.9)
2435: for the Chow weight shows this), giving the inequality
2436: $$
2437: a\le\int_0^ca_0(x)dx-ca_0.
2438: $$
2439: Thus (\ref{mummy}) gives
2440: $$
2441: \frac{N+1}{a_0}\!\left(\!ca_0-\int_0^ca_0(x)dx\!\right)\!<
2442: \sum_{i=1}^{N+1}\rho_i=1(p_{c-1}-p_c)+2(p_{c-2}-p_{c-1})+\ldots+c(p_0-p_1)
2443: \vspace{-10pt} $$$$ \hspace{3cm}
2444: =-p_c-p_{c-1}-\ldots-p_1+cp_0=-\sum_{i=1}^ch^0_{\PP^N}(\I_{\!Z}^i(1))+c(N+1),
2445: $$
2446: which is
2447: \begin{equation} \label{r1}
2448: \frac{\sum_{i=1}^ch^0_{\PP^N}(\I^i_{\!Z}(1))}{\int_0^ca_0(x)dx}
2449: <\frac{N+1}{a_0}\,.
2450: \end{equation}
2451: Semistability is similar, replacing (\ref{mummy}) by the non strict inequality.
2452: \end{proof}
2453:
2454: \begin{rmks}
2455: Setting $c=1$ gives the inequality $\frac{N+1}{a_0}\int_0^1a_0(x)dx>h^0(\I_Z(1))$,
2456: and replacing $Z$ by $kZ$ then gives $\frac{N+1}{a_0}\frac1k\int_0^ka_0(x)dx>
2457: h^0(\I_{\!Z}^k(1))$. We could have gotten this alternative slope-type inequality
2458: directly from stability for the $\C^\times$-action that had all of the above
2459: $\rho_i$ equal to zero (for $i\le p_k$) or one ($i>p_k$). However, it is
2460: strictly weaker than our slope inequality $\frac{N+1}{a_0}\int_0^ka_0(x)dx>
2461: \sum_{i=1}^kh^0(\I_{\!Z}^i(1))$ (\ref{thmchow}) since this last sum is clearly
2462: $\ge kh^0(\I_{\!Z}^k(1))$. \smallskip
2463:
2464: For $X$ semistable, taking $Z=X,\,c=1$ (so $a_0(x)\equiv0$) shows that
2465: $h^0_{\PP^n}(\I_X(1))=0$, i.e.\ $X$ is not contained in any hyperplane and
2466: $H^0_{\PP^N}(\O(1))$ injects into $H^0_X(\O(1))$. To make this injection
2467: an isomorphism requires the assumption that $X\subseteq\PP^N$ is a Kodaira
2468: embedding. \smallskip
2469:
2470: Notice that (\ref{r1}) is more-or-less the $k^{n+1}$
2471: coefficient of (\ref{weight}) (with $r=1$ and rearranged). There can be
2472: higher cohomology corrections since $I$ (\ref{wobbly}) is not the same as
2473: $(\I_Z+(t))^c$ (\ref{straight}); these disappear for $r\gg0$.
2474: \end{rmks}
2475:
2476: We now want to understand to what extent slope Chow stability should
2477: imply Chow stability. This involves
2478: demonstrating the inequality (\ref{mummy}) for \emph{all} linearly independent
2479: sequences of hyperplanes $H_1,H_2,\ldots$ and \emph{all} choices of weights
2480: $0=\rho_1\le\rho_2\le\ldots\le\rho\_{N+1}$. As before we set $Z_i=X\cap
2481: H_1\cap\ldots H_i,\ \I_i:=\I_{Z_i}$.
2482:
2483: The idea is to relate the weight of the associated $\C^\times$-action (with
2484: weight $\rho_i$ on the $i$th vector of our basis of $H^0_{\PP^N}(\O(1))$)
2485: to a sum of weights of $\C^\times$-actions of the standard form considered
2486: in Theorem \ref{thmchow}. The problem is that we have noted
2487: that we can only express weights as a sum of weights of the deformation
2488: to the normal cone of subschemes if either
2489: condition (\ref{reduced}) holds or the multiplicities of $\I_i$ are locally
2490: constant (at least on some resolution where they are ncds). But in either
2491: of these good cases we can demonstrate the general procedure of passing
2492: from slope stability to stability.
2493:
2494: So let us assume for illustration that each $\I_i$ has the local form $O(-k_iD)$
2495: for some reduced Cartier divisor $D$ and number $k_i$ constant on $D$.
2496: (We have seen, as in Theorem \ref{bigger}, one can also pass to a resolution
2497: $\widehat X$ if necessary to make the $Z_i$ sncds to calculate Chow weights
2498: (the $k^{n+1}$-coefficient of Hilbert weights), so the assumption is not
2499: so restrictive, and is sufficient to deal with curves.)
2500: That is, $I=t^{\rho_1}\I_1+t^{\rho_2}\I_2+\ldots+
2501: (t^{\rho\_{\!N+1}})$ has the local form
2502: \begin{equation} \label{deali}
2503: I=\I_D^{k_1}+t^{\rho_2}\I_D^{k_2}+\ldots+(t^{\rho\_{l}}),
2504: \end{equation}
2505: with $0=\rho_1\le\rho_2\le\ldots\le\rho\_{\!N+1}$,
2506: and $k_1\ge k_2\ge\ldots>k_l=0=k_{l+1}=\ldots=k_{N+1}$ (so $l\le N+1$ is
2507: the smallest number with $k_l=0$). These $k_i$ and $l$ may vary as $D$ ranges
2508: over the connected components of $Z_1$.
2509:
2510: \begin{thm} \label{uniform}
2511: If $(X,\O(1))$ is slope Chow stable then the weight $a$ (\ref{mummy})
2512: satisfies $-\frac{N+1}{a_0}\,a<\sum_{i=1}^{N+1}\rho_i+\delta$,
2513: where $\delta=\sum_{i=1}^N(\rho_{i+1}-\rho_i)h^1(\I_i(1))$.
2514: \end{thm}
2515:
2516: This is the inequality $-\frac{N+1}{a_0}\,a<
2517: \sum_{i=1}^{N+1}\rho_i$ required for Chow stability (\ref{mummy}), modulo
2518: some $h^1$ corrections (which we estimated away in the K-stability
2519: analogue). In (\ref{uniformcurve}) the result is strengthened slightly and
2520: the correction estimated on curves to prove their asymptotic
2521: Chow stability.
2522:
2523: \begin{proof}
2524: The $k^{n+1}$-coefficient $a$ of
2525: the weight of our $\C^\times$-action is the same as that on the blow
2526: up of $\widehat X\times\C$ in $I$ (\ref{deali}). In Theorem \ref{divs} we
2527: calculated this to be a sum $a=\sum_Da_D$ over
2528: the connected components $D$ of its support, where
2529: \begin{equation} \label{a}
2530: a_D=-\sum_{i=1}^{l-1}{}'\left(\!\frac{\tilde\rho_{i+1}-\tilde\rho_i}{k_i-k_{i+1}}-
2531: \frac{\tilde\rho_i-\tilde\rho_{i-1}}{k_{i-1}-k_i}\right)\!
2532: \int_0^{k_i}\!\tilde a_0^D(x)dx.
2533: \end{equation}
2534: Here $\tilde\rho_i$ is defined uniquely by requiring that $(k_i,\tilde\rho_i)$
2535: lies on the boundary of the concave hull of the set of points $(k_i,\rho_i)_{i=1}^l$
2536: in the $(k,\rho)$-plane. Thus the $\tilde\rho_i$ need not be integers but,
2537: for instance, $\tilde\rho_1=\rho_1$ and $\tilde\rho_i=\rho_l$ for all $i\ge
2538: l$. More generally, $\tilde\rho_i\le\rho_i$ for all $i$. Concavity of the
2539: $(k_i,\tilde\rho_i)$ ensures that any term in the above sum with a zero
2540: in the denominator also has zero in the numerator; the prime $'$ on the summation
2541: sign signifies that we ignore these $\frac00$ terms (the terms with $k_i=k_{i+1}$)
2542: in the sum; equivalently
2543: we set $\frac00:=0$. In the first term we set $\tilde\rho_0:=0$.
2544:
2545: The Seshadri constant of $D$ is $\ge k_i$ for
2546: all $i$ since $\I_D^{k_i}$ is locally the intersection of
2547: a sequence of hyperplanes, so $\I_D^{k_i}(1)$ is globally generated
2548: near $D$, so $\I_D^{rk_i}(r)$ is too. Therefore Chow slope stability
2549: for $D\subset X$ (\ref{sss}) gives the inequalities
2550: $$
2551: \frac{N+1}{a_0}\int_0^{k_i}\!\tilde a_0^D(x)dx<\sum_{i=1}^{k_i}\big(
2552: N+1-h^0(\I^i_D(1))\big).
2553: $$
2554: Since all of the integrals in (\ref{a}) have coefficients which are $\ge0$
2555: by the concavity of the $(k_i,\tilde\rho_i)$, we obtain
2556: \begin{eqnarray} \nonumber
2557: -\frac{N+1}{a_0}\,a_D &<& \sum_{i=1}^{l-1}{}'\left(\!\frac{\tilde\rho_{i+1}-
2558: \tilde\rho_i}{k_i-k_{i+1}}-\frac{\tilde\rho_i-\tilde\rho_{i-1}}
2559: {k_{i-1}-k_i}\right)\sum_{j=1}^{k_i}\big(N+1-h^0(\I^j_D(1))\big) \\
2560: &=& \label{diverge} \sum_{i=1}^{l-1}{}'\ \frac{\tilde\rho_{i+1}-\tilde\rho_i}
2561: {k_i-k_{i+1}}\sum_{j=k\_{i+1}+1}^{k_i}\big(N+1-h^0(\I^j_D(1))\big) \\ \nonumber
2562: &\le& \sum_{i=1}^N(\tilde\rho_{i+1}-\tilde\rho_i)\big(N+1-h^0(\I_D^{k_i}(1))\big),
2563: \end{eqnarray}
2564: where in the last line we have added back in the $k_i=k_{i+1}$ terms since
2565: they are positive.
2566: Since the last sum is also $\sum_{i=2}^{N+1}\tilde\rho_i(h^0(\I_D^{k_i}(1))-
2567: h^0(\I_D^{k_{i+1}}(1)))$, with a positive coefficient for each $\tilde\rho_i$,
2568: we can replace $\tilde\rho_i$ by $\rho_i\ge\tilde\rho_i$ to give
2569: $$
2570: -\frac{N+1}{a_0}\,a_D<\sum_{i=1}^N(\rho_{i+1}-\rho_i)h^0(\O_{k_iD}(1)).
2571: $$
2572: Sum over the connected components $D$ and use $h^0(\I_i(1))\ge
2573: i$ to give
2574: \begin{eqnarray*}
2575: -\frac{N+1}{a_0}\,a &<& \sum_{i=1}^N(\rho_{i+1}-\rho_i)h^0(\O_{Z_i}(1)) \\
2576: &\le& \sum_{i=1}^N(\rho_{i+1}-\rho_i)\big(N+1-h^0(\I_i(1))+h^1(\I_i(1))\big) \\
2577: &\le& \sum_{i=1}^N(\rho_{i+1}-\rho_i)\big(N+1-i\big)
2578: +\sum_{i=1}^N(\rho_{i+1}-\rho_i)h^1(\I_i(1)) \\
2579: &\le& \sum_{i=1}^{N+1}\rho_i+\delta.
2580: \end{eqnarray*}
2581: (In the above, any sum from $i$ to $j$ with $j<i$ is to be interpreted as
2582: zero, and in passing from the first line to the second we have used the fact
2583: that $\sum_{j=1}^{k_l}=0$ since $k_l=0$.)
2584: \end{proof}
2585:
2586: The Chow slope inequality (\ref{sss}) can be rewritten
2587: $$
2588: \epsilon+\frac{N+1}{a_0}=\epsilon+Ch(X)<Ch_c(\O_Z)=
2589: \frac{\sum_{i=1}^c\big(N+1-h^0(\I_{\!Z}^i(1))\big)}{\int_0^c\!\tilde a_0(x)dx}
2590: $$
2591: for some small $\epsilon>0$. This implies
2592: the weaker inequality (which is all that we shall require for curves)
2593: \begin{equation} \label{css}
2594: \left(\frac{N+1}{a_0}+\epsilon\right)\int_0^c\!\tilde a_0(x)dx<
2595: \sum_{i=1}^ch^0(\O_{iZ}(1)).
2596: \end{equation}
2597: If $\epsilon$ can be chosen uniformly in (\ref{css}) for all $Z\subset X$
2598: we call $X$ \emph{uniformly Chow slope stable with constant $\epsilon$}.
2599: This will help us to deal with the correction $\delta$.
2600:
2601: For $X$ a curve we can improve the estimates of Theorem \ref{uniform} slightly
2602: and use Clifford's theorem to bound the $h^1$ terms:
2603:
2604: \begin{thm} \label{uniformcurve}
2605: If a smooth curve $(X,L)$ of genus $g$ and $d=\deg L>2g-2$ is
2606: uniformly Chow slope stable (\ref{css}) with constant $\epsilon\ge
2607: \big(1+\frac1{d-g}\big)\big(g-\frac12\big).\frac1d$ then it is Chow stable.
2608: \end{thm}
2609:
2610: \begin{proof}
2611: We follow the same proof as above with $\frac{N+1}{a_0}$ replaced by
2612: $\frac{N+1}{a_0}+\epsilon$ throughout, up to (\ref{diverge}), at which
2613: point we use estimates specific to curves to better bound the $h^1$ terms.
2614: That is, $D$ is locally a smooth point $\{p\}$ in $X$, and for those $i$
2615: with $k_i\ne k_{i+1}$ (i.e.\ those involved in the sum $\sum{}'$),
2616: $$
2617: \frac1{k_i-k_{i+1}}\sum_{j=k\_{i+1}+1}^{k_i}h^0(\O_{j\{p\}}(1))=
2618: \frac12\big(h^0(\O_{k_i\{p\}}(1))+h^0(\O_{k_{i+1}\{p\}}(1))+1\big).
2619: $$
2620: Therefore (\ref{diverge}) becomes
2621: $$
2622: -\left(\frac{N+1}{a_0}+\epsilon\right)a_p<\sum_{i=1}^{l-1}{}'
2623: (\tilde\rho_{i+1}-\tilde\rho_i)\frac12\big(h^0(\O_{k_i\{p\}}(1))+
2624: h^0(\O_{k_{i+1}\{p\}}(1))+1\big).
2625: $$
2626: We can now add back in those $i$ with $k_i=k_{i+1}$, as all terms are positive.
2627: Each $\tilde\rho_i$ appears with positive coefficient in the result, since
2628: it can be rearranged as $\sum_{i=2}^N\tilde\rho_i\big(
2629: h^0(\O_{k_{i-1}\{p\}}(1))-h^0(\O_{k_{i+1}\{p\}}(1))\big)+\rho\_{\!N+1}\big(
2630: h^0(\O_{k_N\{p\}}(1))+1\big)$.
2631: So replacing $\tilde\rho_i$ by $\rho_i\ge\tilde\rho_i$, summing over
2632: $p$ in the support
2633: of $Z_1$, and using $h^0(\I_i(1))\ge i$, gives
2634: \enlargethispage*{1ex}
2635: \begin{eqnarray*}
2636: -\left(\frac{N+1}{a_0}+\epsilon\right)a \hspace{-2cm} & \\ \qquad\quad &<&
2637: \sum_{i=1}^N(\rho_{i+1}-\rho_i)\frac12
2638: \big(h^0(\O_{Z_i}(1))+h^0(\O_{Z_{i+1}}(1))+1\big) \\
2639: &\le& \sum_{i=1}^N(\rho_{i+1}-\rho_i)\frac12\Big[N+1-h^0(\I_i(1))
2640: +h^1(\I_i(1))+ \\
2641: && \hspace{45mm} N+1-h^0(\I_{i+1}(1))+h^1(\I_{i+1}(1))+1\Big] \\
2642: &\le& \sum_{i=1}^N(\rho_{i+1}-\rho_i)\big[N+1-i\big]
2643: +\frac12\sum_{i=1}^N(\rho_{i+1}-\rho_i)\big(h^1(\I_i(1))+h^1(\I_{i+1}(1))\big)
2644: \\ &=& \sum_{i=1}^{N+1}\rho_i+\delta, \hspace{20mm}
2645: \delta:=\frac12\sum_{i=1}^N(\rho_{i+1}-\rho_i)
2646: \big(h^1(\I_i(1))+h^1(\I_{i+1}(1))\big)
2647: \end{eqnarray*}
2648: i.e.
2649: \begin{equation}
2650: -\left(1+\epsilon\frac{a_0}{N+1}\right)\frac{N+1}{a_0}\,a
2651: <\left(1+\frac{\delta}{\sum_{i=1}^{N+1}\rho_i}\right)
2652: \sum_{i=1}^{N+1}\rho_i\,. \label{d}
2653: \end{equation}
2654: This $\delta$ is a tiny improvement over the one in Theorem \ref{uniform},
2655: but can be bounded using Clifford's theorem, which for our purposes says
2656: that $h^1(L)\le\max\{1+g-h^0(L),0\}$. Thus $h^1(\I_i(1))\le1+g-i$ for $i\le
2657: g$ and vanishes for $i>g$. This yields
2658: \begin{multline*}
2659: \delta\le\sum_{i=1}^g(\rho_{i+1}-\rho_i)\frac12\big(1+g-i+1+g-(i+1)\big)=
2660: \\ \sum_{i=1}^g(\rho_{i+1}-\rho_i)\Big(g-i+\frac12\Big)=\sum_{i=2}^g\rho_i+
2661: \frac12\rho_{g+1},
2662: \end{multline*}
2663: since $\rho_1=0$. Since the $\rho_i$ are monotonic in $i$,
2664: $$
2665: \sum_{i=2}^g\rho_i+\frac12\rho_{g+1}\le\Big(g-\frac12\Big)\rho_{g+1}
2666: \le\Big(g-\frac12\Big)\frac{\frac12\rho_{g+1}+\sum_{i=g+2}^{N+1}\rho_i}
2667: {N-g+\frac12}\,.
2668: $$
2669: Adding $\frac{g-\frac12}{N-g+\frac12}\big(\sum_{i=2}^g\rho_i+
2670: \frac12\rho_{g+1}\big)$ to both sides gives
2671: $$
2672: \frac N{N-g+\frac12}\left(\sum_{i=2}^g\rho_i+\frac12\rho_{g+1}\right)
2673: \le\frac{g-\frac12}{N-g+\frac12}\sum_{i=2}^{N+1}\rho_i,
2674: $$
2675: and so
2676: $$
2677: \delta\le\sum_{i=2}^g\rho_i+\frac12\rho_{g+1}\le
2678: \frac{g-\frac12}N\sum_{i=2}^{N+1}\rho_i.
2679: $$
2680: Combined with (\ref{d}) we find that uniform slope stability implies that
2681: $$
2682: -\left(1+\epsilon\frac{a_0}{N+1}\right)\frac{N+1}{a_0}\,a<
2683: \left(1+\frac{g-\frac12}N\right)\sum_{i=1}^{N+1}\rho_i,
2684: $$
2685: which implies the inequality $-\frac{N+1}{a_0}\,a<\sum_{i=1}^{N+1}\!\rho_i\,$
2686: required (\ref{mummy}) if $\epsilon\ge\frac{N+1}{a_0}\frac{g-\frac12}N$.
2687: The condition $d>2g-2$ implies that $h^1(\O(1))=0$ so $N+1=d+1-g$ and $a_0=d$.
2688: Therefore the inequality is $\epsilon\ge\big(1+\frac1{d-g}\big)
2689: \big(g-\frac12\big)\frac1d$.
2690: \end{proof}
2691:
2692: \begin{thm}
2693: Smooth curves $(X,\O(1))$ of genus $g\ge1$ are uniformly slope stable (\ref{css})
2694: with any constant $\epsilon<\frac gd$, and so are asymptotically Chow stable.
2695: \end{thm}
2696:
2697: \begin{proof}
2698: We need only demonstrate the inequality
2699: (\ref{css}) in the case of $Z=\{p\}$ a single reduced point, as the inequality
2700: for a multiple of $\{p\}$ is weaker and the inequality for arbitrary $Z$
2701: follows from adding the inequalities for its connected components.
2702: By Riemann-Roch, $\tilde a_0(x)=x$, so that $\int_0^c\tilde a_0(x)=c^2/2$
2703: while $\sum_{i=1}^ch^0(\O_{iZ}(1))=\sum_{i=1}^ci=c(c+1)/2$. Therefore
2704: $$
2705: \left(\frac{d+1-g}d+\epsilon\right)\frac{c^2}2=
2706: \left(\frac{N+1}{a_0}+\epsilon\right)\int_0^c\!\tilde a_0(x)dx\ <\
2707: \sum_{i=1}^ch^0(\O_{iZ}(1))=\frac{c^2+c}2\,,
2708: $$
2709: so long as $\epsilon<\frac1c+\frac{g-1}d$. Of course $c\le d$, so $(X,L)$
2710: is uniformly Chow slope stable with any constant $\epsilon<\frac gd$.
2711:
2712: Theorem \ref{uniformcurve} then gives Chow stability, so long as
2713: $\frac gd>\big(1+\frac1{d-g}\big)\big(g-\frac12\big).\frac1d$ which is true
2714: for $g\ge1$ and sufficiently large $d$.
2715: \end{proof}
2716:
2717: \begin{rmk}
2718: Mumford \cite{Mu} proves the sharper result that $(X,\O(1))$ is Chow stable
2719: for $\deg\O(1)>2g$ using a combinatorial argument.
2720: \end{rmk}
2721:
2722:
2723: \section{Examples} \label{egs}
2724:
2725: Our remaining examples all deal with K-slope stability. Many more examples,
2726: calculations and applications are given in \cite{RT}.
2727:
2728: \subsection{Varieties with nonnegative canonical bundle}\ \vskip 5pt
2729: \noindent
2730: Suppose that $X$ has at worst canonical singularities.
2731: That is $X$ is normal, there is an integer $m$ such that $mK_X$ is
2732: Cartier, and given any resolution of singularities
2733: $\pi_1\colon\overline{\!X}\,\rightarrow X$ we have
2734: \begin{equation}\label{canonical}
2735: mK_{\overline{\!X}}=\pi_1^*(mK_X)+\sum_i\alpha_iF_i \quad\text{with}\ \ \alpha_i\ge 0,
2736: \end{equation}
2737: where the $F_i$ are the irreducible components of the
2738: exceptional set of $\pi_1$. We can define intersection with the canonical
2739: class of $X$ on $\overline{\!X}\,$ by $K_X.(\,\cdot\,):=\frac1m(mK_X).(\,\cdot\,)$.
2740:
2741: For any subscheme $Z\subset X$ let
2742: $$
2743: \overline{\!X}\Rt{\pi_2\ }\widehat X\rt{\pi}X
2744: $$
2745: be a resolution of singularities of $\widehat X$, the blow up of $X$ along
2746: $Z$. $\pi_1=\pi\comp\pi_2\colon\overline{\!X}\rightarrow
2747: X$ is a resolution of singularities of $X$, so (\ref{canonical}) holds.
2748: Letting $F=\pi_2^*E$, $L-xF=\pi_2^*(L-xE)$ is nef on $\overline{\!X}$
2749: for $0\le x\le\epsilon(Z)$.
2750:
2751: We wish to compute $a_i(x)$ on $\overline{\!X}$ instead of $\widehat
2752: X$. Since $\O_{\widehat X}\subseteq\pi_{2*}\O_{\,\overline{\!X}}$ with quotient
2753: supported in codimension one, we have an inclusion
2754: $$
2755: H^0_{\widehat X}((L-xE)^k)\subseteq H^0_{\overline{\!X}}((L-xF)^k),
2756: $$
2757: with cokernel of dimension $\le O(k^{n-1})$ by Fujita vanishing, for
2758: $0\le x<\epsilon(Z)$. As in (\ref{normal}), for $k\gg0$, $h_{\,\overline{\!X}}^i
2759: ((L-xF)^k)=h^0_{\widehat X}((L-xE)^k\otimes R^i\pi_{2*}\O)=O(k^{n-1-i})$
2760: since the support of $R^i\pi_{2*}\O$ has codimension at least $i+1$.
2761: Therefore $\chi\_{\widehat X}((L-xE)^k)=\chi\_{\overline{\!X}}((L-xF)^k)
2762: -ak^{n-1}+O(k^{n-2})$ for some $a\ge0$. That is, for $0\le x<\epsilon(Z)$,
2763: we have
2764: $$
2765: a_0(x) = \frac{1}{n!}(L-xF)^n, \vspace{-6pt}
2766: $$
2767: and \vspace{-6pt}
2768: \begin{equation}\label{a_idefinedbyaresolution}
2769: \qquad\qquad a_1(x)\le-\frac{1}{2(n-1)!}K_{\overline{\!X}}.(L-xF)^{n-1}
2770: \le-\frac{1}{2(n-1)!}K_{X}.(L-xF)^{n-1},
2771: \end{equation}
2772: where the second inequality follows from (\ref{canonical}) and the
2773: fact that $L-xF$ is nef. Again as in (\ref{normal}), the first two
2774: terms of the Euler characteristic of $L$ are the same as those of
2775: $\pi_1^* L$, so equality holds in (\ref{a_idefinedbyaresolution}) when
2776: $x=0$. That is $a_0=\frac1{n!}L^n$ and
2777: \begin{equation} \label{a_1definedbyaresolution}
2778: a_1 = -\frac{1}{2(n-1)!} K_{\overline{\!X}}.L^{n-1}=- \frac{1}{2(n-1)!} K_X.L^{n-1},
2779: \end{equation}
2780: since $L$ is trivial along the $F_i$. With these preliminaries, it becomes
2781: easy to prove the following.
2782:
2783: \begin{thm} \label{CY} \textbf{Calabi-Yaus and canonical models}. \\
2784: Let $X$ be an irreducible variety with at worst canonical singularities.
2785: \newline
2786: ${}\quad\bullet$ If $K_X$ is numerically trivial then $(X,L)$ is slope stable for all polarisations $L$. \newline
2787: ${}\quad\bullet$ If $K_X$ is ample then the canonical polarisation $(X,K_X)$
2788: is slope stable.
2789: \end{thm}
2790:
2791: \begin{proof}
2792: In both cases, $K_X\sim\alpha L$ is numerically equivalent to a nonnegative
2793: multiple $\alpha\ge0$ of the polarisation. So by (\ref{a_1definedbyaresolution}),
2794: $\mu(X)=a_1/a_0=-n\alpha/2$. By (\ref{a_idefinedbyaresolution}),
2795: $$
2796: -\mu(X)a_0(x)+a_1(x)\le\frac{\alpha}{2(n-1)!}(L-xF)^n-\frac1{2(n-1)!}
2797: (\alpha L).(L-xF)^{n-1},
2798: $$
2799: which equals $-\frac{\alpha x}{2(n-1)!}F.(L-xF)^{n-1}\le0$ since $L-xF$ is
2800: nef. Since $a_0'(x)<0$ for $x\in(0,\epsilon(Z))$ (\ref{a0'}), integration
2801: gives
2802: $$
2803: -\mu(X)\int_0^ca_0(x)dx+\int_0^ca_1(x)+\frac{a_0'(x)}2dx<0 \quad\text{for}\
2804: c\in(0,\epsilon(Z)],
2805: $$
2806: which rearranges to give slope stability: $\mu_c(\I_Z)<\mu(X)$.
2807: \end{proof}
2808:
2809: With more work, this can be generalised as follows.
2810:
2811: \begin{thm}
2812: Suppose that $(X,L)$ has at worst canonical singularities, and
2813: $K_X$ is nef and big. Then $(X,L)$ is slope stable
2814: for $L$ ample and sufficiently close to $K$. More precisely,
2815: \begin{itemize}
2816: \item For any divisor $G$ there is a $\delta_0>0$ such
2817: that if $0\le \delta<\delta_0$ and $L=K_X+\delta G$ is ample then
2818: $(X,L)$ is slope stable.
2819: \item If $2\mu(X,L)L + nK_X$ is nef then $(X,L)$ is slope stable.
2820: \item If $-2\mu(X,L)L - nK$ is nef then $(X,L)$ is slope stable.
2821: \end{itemize}
2822: \end{thm}
2823:
2824: In \cite{RT} we prove that no smooth $Z$ can slope destabilise a smooth $(X,L)$
2825: with these properties, and the proof extends to general $Z$ and $X$ with
2826: canonical singularities using the preliminaries
2827: (\ref{a_idefinedbyaresolution}) and (\ref{a_1definedbyaresolution}).
2828:
2829: These results are to be expected due to the deep and difficult related
2830: results of Viehweg \cite{V}, and the expectation that the minimal
2831: model programme can be carried out in all dimensions (and so can be
2832: done in families) \cite{Ka}. In fact Theorem \ref{CY} can be proved in a
2833: round about way for smooth varieties with no holomorphic vector fields by
2834: the stability results of \cite{Do1, Zh} applied to the K\"ahler-Einstein
2835: metrics of \cite{Au, Y} on such
2836: varieties. Similarly, if $X$ has no holomorphic vector fields and $L$ is
2837: ample and sufficiently close to $K_X$ then an implicit function theorem
2838: argument applied to the K\"ahler-Einstein metric provides a constant scalar
2839: curvature K\"ahler metric in $[c_1(L)]$. But the quick proofs
2840: above demonstrate that using slope stability it could become much easier
2841: to produce and compactify moduli of varieties with semi-ample canonical
2842: bundle \cite{V}.
2843:
2844:
2845: \subsection{Irreducible Curves} \label{curves}\ \vskip 5pt
2846:
2847: \begin{prop}\label{prop:stabilityofpolarisedcurves}
2848: Let $(\Sigma,L)$ be an irreducible polarised curve with arithmetic genus
2849: $g\ge 2$. The Hilbert-Samuel polynomial of a subscheme $Z\subset\Sigma$
2850: can be written
2851: \begin{equation}\label{hilbertsamuelofsubschemesofcurves}
2852: h^0(L^k\big/(L^k\otimes\I_{\!Z}^{xk})) = e(Z)xk-\rho(Z)\quad \text{for}\ k\gg0,
2853: \end{equation}
2854: and $Z$ destabilises $\Sigma$ if and only if it strictly
2855: destabilises, if and only if $2\rho(Z)> e(Z)$.
2856: \end{prop}
2857:
2858: \begin{proof}
2859: The assumption on the genus implies that $\mu(\Sigma)<0$. As $\dim
2860: \Sigma=1$,
2861: $\tilde{a}_0(x)$
2862: is a degree 1 polynomial vanishing at the origin, while $\tilde a_1(x)$
2863: has degree 0. So writing $\tilde a_0(x)=e(Z)x$ and $\tilde{a}_1(x)=-\rho(Z)$,
2864: $$
2865: \mu_c(\O_Z) = \frac{\int_0^c \tilde{a}_1(x) +
2866: \frac{\tilde{a}_0'(x)}{2} dx}{\int_0^c \tilde{a}_0(x) dx} =
2867: \frac{e(Z)-2\rho(Z)}{e(Z)c}\,.
2868: $$
2869: This has the same sign as $e(Z)-2\rho(Z)$ for all $c>0$, and tends to $\pm\infty$
2870: as $c\to0$. Thus $p$ destabilises if and only if it strictly destabilises
2871: if and only if this sign is negative.
2872: \end{proof}
2873:
2874: \begin{thm} \label{multiplicityofdestabilisingsingularpoints}
2875: Let $(\Sigma,L)$ be a polarised irreducible curve of arithmetic
2876: genus $g\ge 2$.
2877: \begin{itemize}
2878: \item If $\Sigma$ has a point of multiplicity $e\ge 3$ then
2879: $(\Sigma,L)$ is not slope stable.
2880: \item If $\Sigma$ has at worst ordinary double points then
2881: $(\Sigma,L)$ is slope stable.
2882: \end{itemize}
2883: \end{thm}
2884:
2885: \begin{proof}
2886: For the first statement suppose $Z=\{p\}$ is a point of $\Sigma$
2887: with multiplicity $e\ge 3$. Northcott has shown (\cite{No}
2888: Lemma 1) that $\rho\ge e-1$. Hence $2\rho\ge 2e-2>e$ and
2889: $Z$ strictly destabilises.
2890:
2891: For the second statement suppose for a contradiction that $Z$ is a
2892: destabilising subscheme of $\Sigma$. From Proposition
2893: \ref{simplify} (especially (\ref{simplify2})) and Lemma
2894: \ref{seshadridisconnected} we see that
2895: $\mu_c(\I_{Z_0})\ge\mu(\Sigma)$ for some connected component $Z_0$ of $Z$
2896: and $c\le\epsilon(Z_0)$ (we are not saying that $Z_0$ destabilises since
2897: we may not be allowed to take $c=\epsilon(Z_0)$). If the support $p\in\Sigma$
2898: of $Z_0$ is a smooth point then $Z_0=m\{p\}$ must by a thickened point,
2899: which
2900: by Proposition \ref{simplify} shows that $\mu_{mc}(\I_{\{p\}})\ge\mu(X)$.
2901: But in the notation of (\ref{hilbertsamuelofsubschemesofcurves}),
2902: $e(\{p\})=1$, which by Theorem 3.2 of \cite{KM} implies that
2903: $\rho(\{p\})=0$, so by Proposition \ref{prop:stabilityofpolarisedcurves}
2904: $\mu_{mc}(\I_{\{p\}})$ is in fact $<\mu(X)$.
2905:
2906: Hence $Z$ must be supported at one of the singular points, and we
2907: can reduce to looking at the local analytic model $\Spec R,\
2908: R=\C[X,Y]/(XY)$. Let $I$ be an ideal of $R$
2909: which is supported at $(X,Y)$. (By abuse of notation we shall not
2910: distinguish between a polynomial in two variables and its class in
2911: $R$.) We can pick a finite number of generators of $I$ of the form
2912: $f_i=a_iX^{p_i}+b_iY^{q_i}$, with $a_i,b_i\in\mathbb C$.
2913:
2914: Let $p=\min\{p_j\colon a_j\neq 0\},\ q=\min\{q_j\colon a_j\neq 0\}$;$p,q\ge1$ since $I$ is supported at the origin.
2915: Pick an $i$ such that $p_i=p$ and $a_i\ne0$;
2916: then $X^{pk+1}=\frac{X}{a_i^k}(a_iX^{p_i} + b_i Y^{q_i})^k\in I^k$; similarly
2917: $Y^{qk+1}\in I^k$. Therefore
2918: $R/I^k$ is spanned by $\{1,X,\ldots,X^{pk},Y,\ldots
2919: Y^{qk}\}$. By the definition of $p$ and $q$, $I^k$ is spanned by
2920: $\{X^i, Y^j \colon i\ge pk, j\ge qk\}$, so the vectors $\{1,X,\ldots
2921: X^{pk-1}, Y,\ldots, Y^{qk-1}\}$ in $R/I^k$ are linearly independent.
2922: Thus
2923: $$
2924: (p+q)k-1\le \dim R/I^k \le (p+q)k+1.
2925: $$
2926: Writing $\dim R/I^k=ek-\rho$ we have $-1\le\rho\le 1$.
2927: Hence $2\rho \le 2\le p+q=e$, so by Proposition
2928: \ref{prop:stabilityofpolarisedcurves} $I$ does not destabilise.
2929: \end{proof}
2930:
2931: \begin{rmk}
2932: Eisenbud and Mumford \cite{Mu} analyse the effect of singular points
2933: on Chow stability of higher dimensional varieties. It would be interesting
2934: to know if their results can be seen using slope stability.
2935: \end{rmk}
2936:
2937: These results combined with Theorem \ref{thm:kstableslopestable} imply
2938: that curves with singularities of multiplicity greater than two are strictly
2939: K-unstable. We cannot deduce positive results about K-stability from the
2940: results of Section \ref{converse}, however, since $\Sigma$ need not
2941: be normal. Unless, that is, $\Sigma$ is smooth:
2942:
2943: \begin{thm}\label{smoothcurvesareKstable} Any smooth polarised curve $(\Sigma,L)$
2944: of genus $g$ is K-stable if $g\ge1$ and strictly K-polystable if $g=0$.
2945: \end{thm}
2946:
2947: \begin{proof}
2948: By Corollary \ref{Kslopecurve} it is equivalent to prove the results for
2949: slope (poly)stability. Instead of using previous results it is now easier
2950: to proceed directly. Any nonempty subscheme $Z$ is a divisor
2951: of degree $d>0$, so
2952: $$
2953: \chi(L^k\otimes\I_{\!Z}^{xk})=k\deg L-xdk+1-g
2954: $$
2955: shows that $\tilde a_0(x)=xd$ and $\tilde a_1(x)=0$. Thus $\mu_c(\O_Z)=
2956: \frac{cd}{c^2d}=\frac1{c}>0\ge\frac{1-g}{\deg L}=\mu(X)$ for $g\ge1$, proving
2957: slope stability.
2958:
2959: For $g=0$, $c$ may take values up to and \emph{including} $\epsilon(Z)=\deg
2960: L/d$, since $L^d\otimes\I_Z^{\deg L}=\O_{\PP^1}(d\deg L-d\deg L)=\O_{\PP^1}$
2961: is globally generated.
2962: Thus $\mu_c(\O_Z)\ge\frac{d}{\deg L}\ge\frac1{\deg L}=\mu(X)$ with equality
2963: (strict semistability) only for $d=1$, \emph{i.e.}\ $Z$ a single point, and
2964: $c=\epsilon(Z)$. Since the deformation to
2965: the normal cone of a single point on $\PP^1$ blows down to $\PP^1\times\C$
2966: (with a nontrivial $\C^\times$-action) from which the relevant line bundle
2967: $\L_c$ pulls back, we find $\PP^1$ is in fact slope polystable.
2968: \end{proof}
2969:
2970: This can also be proved using the constant curvature
2971: metric on $\Sigma$ and analysis of the Mabuchi functional, but this
2972: seems to be the first direct algebraic proof.
2973:
2974:
2975: \begin{thebibliography}{GIT}
2976:
2977: \bibitem[Au]{Au}
2978: Aubin, T. (1976).
2979: \emph{\'Equations du type Monge-Amp\`ere sur les vari\'et\'es k\"ahleriennes
2980: compactes.} C. R. Acad. Sci. Paris S\'r. A-B \textbf{283}, A119--A121.
2981:
2982: \bibitem[De]{De}
2983: Demailly, J.-P (1994).
2984: \emph{$L\sp 2$ vanishing theorems for positive line bundles and
2985: adjunction theory}, Transcendental methods in algebraic geometry
2986: (Cetraro), 1--97.
2987:
2988: \bibitem[Do1]{Do1}
2989: Donaldson, S. K. (2001).
2990: \emph{Scalar curvature and projective embeddings, I.}
2991: Jour. Diff. Geom. \textbf{59}, 479--522.
2992:
2993: \bibitem[Do2]{Do2}
2994: Donaldson, S. K. (2002).
2995: \emph{Scalar curvature and stability of toric varieties}. Jour. Diff. Geom.
2996: \textbf{62}, 289--349.
2997:
2998: \bibitem[Fu]{Fu}
2999: Fulton, W. (1984).
3000: \emph{Intersection theory}. Springer-Verlag, Berlin.
3001:
3002: \bibitem[Gi]{Gi}
3003: Gieseker, D. (1977).
3004: \emph{Global moduli for surfaces of general type.}
3005: Invent. Math. \textbf{43}, 233--282.
3006:
3007: \bibitem[Gr]{Gr}
3008: Grothendieck, A. (1960/61).
3009: \emph{Techniques de construction et th\'eor\`emes d'existence en g\'eom\'etrie
3010: alg\'ebrique. IV. Les sch\'emas de Hilbert.} S\'eminaire Bourbaki, No. 221.
3011:
3012: \bibitem[Ha]{Ha}
3013: Hartshorne, R. (1977).
3014: \emph{Algebraic Geometry.} Graduate Texts in Mathematics 52, Springer-Verlag.
3015:
3016: \bibitem[Hi]{Hi}
3017: Hildebrand, F. B. (1974).
3018: \emph{Introduction to numerical analysis}, 2nd Ed., McGraw-Hill,
3019: New York.
3020:
3021: \bibitem[HL]{HL}
3022: Huybrechts, D. and Lehn, M. (1997).
3023: \emph{Geometry of moduli spaces of shaves.}
3024: Aspects in Mathematics Vol. E31, Vieweg.
3025:
3026: \bibitem[Ka]{Ka}
3027: Karu, K. (2000).
3028: \emph{Minimal models and boundedness of stable varieties}. Jour. Alg. Geom.
3029: \textbf{9}, 93--109.
3030:
3031: \bibitem[KM]{KM}
3032: Kirby, D. and Mehran, A. (1982)
3033: \emph{A note on the
3034: coefficients of the Hilbert-Samuel polynomial for a Cohen-Macaulay
3035: module.} Jour. London Math. Soc. (2) \textbf{no. 3}, 449--457.
3036:
3037: \bibitem[Kl]{Kl}
3038: Kleiman, S. L. (1966).
3039: \emph{Toward a numerical theory of ampleness.} Ann. of Math. \textbf{84},
3040: 293--344.
3041:
3042: \bibitem[Ko]{Ko}
3043: Koll\'ar, J. (1994).
3044: \emph{Projectivity of complete moduli.}
3045: Jour. Diff. Geom \textbf{32}, 235--268.
3046:
3047: \bibitem[La]{La}
3048: Lazarsfeld, R. (2004).
3049: \emph{Positivity in algebraic geometry. I. Classical setting: line bundles
3050: and linear series.} Ergeb. Math. Grenzgeb. (3), Springer-Verlag.
3051:
3052: \bibitem[Li]{Li}
3053: Li, J. (1993).
3054: \emph{Algebraic geometric interpretation of Donaldson's polynomial invariants.}
3055: Jour. Diff. Geom. \textbf{37}, 417--466.
3056:
3057: \bibitem[Ma]{Ma}
3058: Matlis, E. (1973)
3059: \emph{The multiplicity and reduction number of a one-dimensional local
3060: ring.} Proc. London Math. Soc. \textbf{26}, 273--288.
3061:
3062: \bibitem[Mat]{Mat}
3063: Matsusaka, T. (1972).
3064: \emph{Polarized varieties with a given Hilbert polynomial.}
3065: Amer. J. Math. \textbf{94}, 1027--1077.
3066:
3067: \bibitem[Mo]{Mo}
3068: Morrison, I. (1980).
3069: \emph{Projective stability of ruled surfaces.}
3070: Invent. Math. \textbf{56}, 269--304.
3071:
3072: \bibitem[Mor]{Mor}
3073: Mori, S. (1982).
3074: \emph{Threefolds whose canonical bundles are not numerically effective.}
3075: Ann. of Math. \textbf{116}, 133--176
3076:
3077: \bibitem[Mu]{Mu}
3078: Mumford, D. (1977).
3079: \emph{Stability of projective varieties.}
3080: Enseignement Math. (2) \textbf{23}, 39--110.
3081:
3082: \bibitem[GIT]{GIT}
3083: Mumford, D., Fogarty, J. and Kirwan, F. (1994).
3084: \emph{Geometric Invariant Theory.}
3085: Third edition, Erg. Math. \textbf{34}, Springer-Verlag, Berlin.
3086:
3087: \bibitem[No]{No}
3088: Northcott, D. (1960).
3089: \emph{A note on the coefficients of the abstract Hilbert function.} J. London Math. Soc. \textbf{35}, 209--214.
3090:
3091: \bibitem[PT]{PT}
3092: Paul, S. and Tian, G. (2004).
3093: \emph{Algebraic and Analytic K-Stability.} Preprint math.DG/0405530.
3094:
3095: \bibitem[Ro]{Ro}
3096: Ross, J. (2003).
3097: \emph{Instability of polarised algebraic varieties.} PhD thesis, Imperial
3098: College.
3099:
3100: \bibitem[RT]{RT}
3101: Ross, J. and Thomas, R. P. (2004).
3102: \emph{An obstruction to the existence of constant scalar
3103: curvature K\"{a}hler metrics.} Preprint math.DG/0412518.
3104:
3105: \bibitem[Sz]{Sz}
3106: Sz\'ekelyhidi, G. (2004).
3107: \emph{Extremal metrics and K-stability.} Preprint math.AG/0410401.
3108:
3109: \bibitem[Ti1]{Ti1}
3110: Tian, G. (1994).
3111: The $K$-energy on hypersurfaces and stability.
3112: Comm. Anal. Geom. \textbf{2}, 239--265.
3113:
3114: \bibitem[Ti2]{Ti2}
3115: Tian, G. (1997).
3116: \emph{K\"ahler-Einstein metrics with positive scalar curvature.}
3117: Invent. Math. \textbf{130}, 1--37.
3118:
3119: \bibitem[V]{V}
3120: Viehweg, E. (1995).
3121: \emph{Quasi-projective moduli for polarized manifolds.}
3122: Erg. Math. (3) \textbf{30}. Springer-Verlag, Berlin.
3123:
3124: \bibitem[Y]{Y}
3125: Yau, S.-T. (1978).
3126: \emph{On the Ricci curvature of a compact K\"ahler manifold and the complex
3127: Monge-Amp\`ere equation. I.} Comm. Pure Appl. Math. \textbf{31}, 339--411.
3128: \bibitem[Zh]{Zh}
3129: Zhang, S. (1996).
3130: \emph{Heights and reductions of semi-stable varieties.}
3131: Compositio Math. \textbf{104}, 77--105.
3132:
3133: \end{thebibliography}
3134:
3135: \vskip 4mm
3136:
3137: {\small \noindent {\tt jaross@math.columbia.edu} \\
3138: \noindent \small{\tt richard.thomas@imperial.ac.uk}} \newline
3139: \noindent Department of Mathematics, Columbia University, New York, NY 10027.
3140: USA. \\
3141: Department of Mathematics, Imperial College, London SW7 2AZ. UK.
3142:
3143: \end{document}
3144:
3145:
3146: