-
Notifications
You must be signed in to change notification settings - Fork 0
/
greatestlower.tex
205 lines (197 loc) · 16.2 KB
/
greatestlower.tex
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
61
62
63
64
65
66
67
68
69
70
71
72
73
74
75
76
77
78
79
80
81
82
83
84
85
86
87
88
89
90
91
92
93
94
95
96
97
98
99
100
101
102
103
104
105
106
107
108
109
110
111
112
113
114
115
116
117
118
119
120
121
122
123
124
125
126
127
128
129
130
131
132
133
134
135
136
137
138
139
140
141
142
143
144
145
146
147
148
149
150
151
152
153
154
155
156
157
158
159
160
161
162
163
164
165
166
167
168
169
170
171
172
173
174
175
176
177
178
179
180
181
182
183
184
185
186
187
188
189
190
191
192
193
194
195
196
197
198
199
200
201
202
203
204
205
\chapter{$R(X)$ in complexity one} \label{chap:R(X)}
\chaptermark{The greatest lower bound on Ricci curvature}
Recall, as discussed in the introduction, that one approach to the existence of K\"ahler-Einstein metrics is the study of the continuity path, that is solutions \(\omega_t \in 2 \pi c_1(X) \) to the equation
\[
\Ric(\omega_t) = t\omega_t + (1-t) \omega.
\]
for \(t \in [0,1]\). By \cite{Yau1977} there is always a solution for \(t = 0\). However, Tian \cite{tian1992stability} showed that for some \(t\) sufficiently close to \(1\) there may not be a solution for certain Fano manifolds. It is natural to ask for the supremum of permissible \(t\), which turns out to be independent of the choice of \(\omega\).
\begin{definition}
Let \((X,\omega)\) be a K\"ahler manifold with \(\omega \in 2 \pi c_1(X)\). Define:
\[
R(X) := \sup ( t \in [0,1] : \exists \ \omega_t \in 2 \pi c_1(X) \ \Ric( \omega_t) = t \omega_t + (1-t) \omega ).
\]
\end{definition}
This invariant was first discussed, although not explicitly defined, by Tian in \cite{Tian87}. It was first explicitely defined by Rubenstein in \cite{rubinstein2008} and was further studied by Szekelyhidi in \cite{szekelyhidi2011}. It is sometimes referred to as the greatest lower bound on Ricci curvature.
In \cite{rubinstein2008} Rubenstein showed relation between \(R(X)\) and Tian's alpha invariant \(\alpha(X)\), and in \cite{rubinstein2009} conjectured that \(R(X)\) characterizes the \(K\)-semistability of \(X\). This conjecture was later verified by Li in \cite{li2017}.
In \cite{li2011} Li determined a simple formula for \(R(X_\Delta)\), where \(X_\Delta\) is the polarized toric Fano manifold determined by a reflexive lattice polytope \(\Delta\). This result was later recovered in \cite{datar2016kahler}, by Datar and Sz\'ekelyhidi, using notions of \(G\)-equivariant \(K\)-stability. Previously \(R(X)\) has been calculated for group compactifications by Delcroix \cite{delcroix2017} and for homogeneous toric bundles by Yao \cite{yao2017}. Let us briefly recall the toric formula.
\begin{theorem}[{\cite[Theorem 1?]{li2009greatest}}]
Suppose \(X\) is a smooth Fano toric variety. Let \(P\) be the corresponding Fano polytope. If \(\bc(P) = 0\) then \(X\) is K\"ahler-Einstein and \(R(X) = 1\). Otherwise let \(q\) be the intersection of the ray generated by \(-\bc(P)\) with the boundary \(\partial P\). We then have:
\[
R(X) = \frac{|q|}{|q-\bc(P)|}.
\]
\end{theorem}
\begin{example}
Consider the toric variety \(X = \Bl_z \PP^1 \times \PP^1\) from Example~\ref{toricexample}. It is then easy to calculate \(R(X)\) from the polytope \(P\) given in Example~\ref{ex:toricpolytope}. We have \(\bc(P) = (-2/21,-2/21)\) and \(q = (1/2,1/2)\) so \(R(X) = 21/25\).
\begin{figure}[h]
\centering
\begin{tikzpicture}[scale=1]
(1,-1) -- (1,0) -- (0,1) -- (-1,1) -- (-1,-1) -- (1,-1);
%\draw[dotted,step=1,gray,] (-2,-2) grid (2,2);
\draw[] (1,-1) -- (1,0) -- (0,1) -- (-1,1) -- (-1,-1) -- (1,-1);
\draw (0,0) node {\textbullet};
\draw (-2/21,-2/21) node[below] {\tiny{$\bc(P)$}};
\draw (-4/21,-4/21) node {\tiny{\textbullet}};
\draw (1/2,1/2) node {\tiny{\textbullet}};
\draw (1/2,1/2) node[right] {\tiny{$q$}};
\draw [->] (-4/21,-4/21) -- (1,1);
\foreach \i in {-2,...,2}{
\foreach \j in {-2,...,2}{
\draw[gray] (\i,\j) node {\tiny{$\circ$}};
}
}
\end{tikzpicture}
\caption{$R(X)$ calculation for a toric $X$}
\end{figure}
\end{example}
Using similar methods to \cite{datar2016kahler} we obtain an effective formula for manifolds with a torus action of complexity one, in terms its divisorial polytope. Let \(\Phi: \Box \to \wdiv_\QQ \PP^1\) be the Fano divisorial polytope corresponding to a smooth Fano complexity one \(T\)-variety \(X\). Let \(\{\Delta_y\}_{y \in \PP^1}\) be the finite set of degeneration polytopes corresponding to central fibres of the non-product test configurations of \(X\), as in Definition~\ref{def:configpoly}.
To state our result we must introduce a little more notation. Suppose we have \(\bc(\Delta_y) \neq 0\) for some i. Consider the halfspace \(H := N_\RR \times \RR^{\ge 0} \subset N'_\RR\). Let \(F_y\) be the face of \(\Delta_y\) in which \(q_y\) lies, and let \(S\) be the set of points \(y\) for which \(\bc(\Delta_y) \neq 0\) and all outer normals to \(F_y\) lie in \(H\). Note, by definition of \(\Delta_y\), we have:
\[
S = \{ y \in \PP^1 | \bc(\Delta_y) \not\in H \}
\]
Recall the definition of the Duistermatt Heckman measure \(\nu\), and associated weighted barycenter of \(\Box\) given in Definition~\ref{def:divpoly2}. Suppose \(\bc_\nu(\Box) \neq 0\). Let \(q\) be the intersection of the ray generated by \(-\bc_\nu(\Box)\) with \(\partial \Box\). Let \(q_y\) be the point of intersection of \(\partial \Delta_y\) with the ray generated by \(-\bc(\Delta_y)\).
Note, by the equation for the Donaldson Futaki invariants (\ref{eq:futaki-character}) and Theorem~\ref{thm:DS}, we know that \(R(X) = 1\) if \(\bc(\Delta_y) \in \{0\} \times \RR^+\) for each \(y\). We may now state our result:
\begin{theorem}[{\cite[Theorem 1.1]{cable2019greatest}}] \label{thm:R(X)} Let \(X\) be a complexity one Fano \(T\)-variety as above. If \(\bc(\Delta_y) \in \{0\} \times \RR^+\). Otherwise:
\begin{equation} \label{eq:R(X)}
R(X) = \min \ \left\lbrace \frac{|q|}{|q-\bc_\nu(\Box)|} \ \right\rbrace \cup \ \left\lbrace \frac{|q_y|}{|q_y - \bc(\Delta_y)|} \right\rbrace_{\bc(\Delta_y) \not\in H} .
\end{equation}
\end{theorem}
\begin{example}
To see that this formula is truly a generalization of Li's result, consider the situation of a toric downgrade. Suppose \(X\) is a Fano toric variety given by a polytope \(P\). Let \(\Phi = \Phi_0 \otimes \{0\} + \Phi \otimes \{\infty\}\) be the divisorial polytope obtained through the downgrade procedure in Example~\ref{ex:polytopedowngrade}. Under the associated surjection \(p: M' \to M\) we have \(p(\bc(\Delta_0)) = p(\bc(\Delta_\infty)) = \bc_\nu(\Box)\). It is easy to see \(\emptyset \neq S \subset \{0,\infty\}\). By convexity of \(P\) then for \(y \in S\) we have \(|q| \ge |p(q_y)|\), and moreover:
\[
\frac{|q|}{|q-\bc_\nu(\Box)|} \ge \frac{|p(q_y)|}{|p(q_y)-\bc_\nu(\Box)|} = \frac{|p(q_y)|}{|p(q_y-\bc(\Delta_y))|} = \frac{|q_y|}{|q_y-\bc(\Delta_y)|}
\]
Clearly then the minimum in \normalfont{(}\ref{eq:R(X)}\normalfont{)} is obtained at one of \(y \in \{0,\infty\}\)
\end{example}
\begin{example}
Consider the \((\CC^*)^2\)-threefold 2.30 from Example \ref{ex:sym}. There are \(3\) normal toric degenerations, given by the polytopes \(\Delta_{0},\Delta_1,\Delta_{\infty}\). It can easily be checked in this case that \(S = \emptyset\), see Figure~\ref{fig2:a} for example, or Appendix~\ref{appendix1}.
\begin{figure}[h]
\subcaptionbox{\label{fig2:a}}{\degendiagram}\hfill%
\subcaptionbox{\label{fig2:b}}{\momentdiagram}%
\caption{Some of the calculation of $R(X)$ for threefold 2.30. (a) Degeneration polytope $\Delta_1$ with barycenter $\bc(\Delta_1)$ and $q_1,n_1$ shown, (b) Moment polytope $\Box$ with Duistermaat-Heckmann barycenter $\bc_\nu(\Box)$ and $q$ shown.}
\end{figure}
Therefore \(R(X)\) is given by the first term in the minimum. We calculate \(\bc_\nu(\Box)= (0,-6/23)\) and \(q = (0,1)\). Then:
\[
R(X) = \frac{1}{1+ 6/23} = \frac{23}{29}.
\]
\end{example}
\begin{corollary}[{\cite[Corollary 1.2]{cable2019greatest}}] \label{cor:R(X)}
In Table~\ref{table:name} below we give \(R(X)\) for \(X\) a Fano threefold admitting a \(2\)-torus action appearing in the list of Mori and Mukai \cite{mori1981classification}. We include only those where \(R(X) <1\). Note all admit a K\"ahler-Ricci soliton by Theorem~\ref{thm:sol}.
\end{corollary}
\begin{table}[H] \centering
\begin{tabular}{l l}
\ & \ \\
\hline
X & R(X) \\
\hline
2.30 & \(23/29\) \\
2.31 & \(23/27\) \\
3.18 & \(48/55\) \\
3.21 & \(76/97\) \\
3.22 & \(40/49\) \\
3.23 & \(168/221\) \\
3.24 & \(21/25\) \\
4.5* & \(64/69\) \\
4.8 & \(76/89\) \\
\hline
\ & \ \\
\end{tabular}
\caption{Calculations for complexity \(1\) threefolds appearing in the list of Mori and Mukai for which \(R(X) <1\)} \label{table:name}
\end{table}
Let \(X\) be a \(T\)-variety of complexity one associated to a divisorial polytope \(\Psi: \Box \to \wdiv_\QQ(\PP^1)\), see Section~\ref{prelim:Tvar}.
It follows from Theorem~\ref{thm:DS} that:
\begin{equation} \label{eq:R(X)inf}
R(X) = \inf_{(\X,\L)}( \sup(t | \DF_t(\X,\L,\cdot) \ge 0) ),
\end{equation}
where \((\X,\L)\) varies over all special test configurations for \((X,L)\). We have an explicit description of special test configurations and their Donaldson Futaki invariants, see Section~\ref{sec:eqKstab}. We will calculate \(R(X)\) by considering first the product configurations and then the non-product configurations. To calculate the values \(\sup(t | \DF_t(\X,\L) \ge 0)\) for a given configuration we need to first consider some elementary convex geometry.
\section{A short digression into convex geometry}
Let \(V\) be a real vector space and \(P \subset V\) be a full dimensional convex polytope containing the origin. Fix some point \(b \in \text{int}(P)\). Let \(q \in \partial P\) be the intersection of \(\partial P\) with the ray \(\tau = \RR^+ (-b)\). Suppose \(n \in V^\vee\) is an outer normal to a face containing \(q\). For \(a \in \partial P\) write \(\mathcal{N}(a) = \{w\in V^\vee \ | \langle a,w \rangle = \max_{x \in P} \langle x,c \rangle \}\). For \(w \in \mathcal{N}(a) \) let \(\Pi(a,w)\) be the affine hyperplane tangent to \(P\) at \(a\) with normal \(w\). For \(w \in \inte(\tau^\vee) \) there is a well-defined point of intersection of \(\Pi(a,w)\) and \(\tau\) which we denote \(r_w\). See Figure~\ref{schematic} for a schematic.
\begin{figure}[h] \centering
\diagram1
\caption{An Example in \(V \cong \RR^2\)}
\label{schematic}
\end{figure}
\begin{lemma} \label{R(X):Lemma3.1}
Fix \(w \in \inte(\tau^\vee) \backslash (\RR^+ n)\). For \(s \in [0,1]\) set \(w(s) := sn + (1-s)w\). As \(n \in \tau^\vee \) we may consider \(r(s) := r_{w(s)}\). For \(0 \le s' < s \le 1\) we then have:
\[
\frac{|r(s)|}{|r(s)-b|} < \frac{|r(s')|}{|r(s')-b|}.
\]
\end{lemma}
\begin{proof}
Without loss of generality we may assume \(s' = 0\). For \(s \in [0,1]\) the points \(r(s), q,b\) are collinear, so \(|r(s)|= |r(s)-q|+|q|\) and \(|r(s)-b| =|r(s)-q| +|q| + |b| \). Therefore:
\[
\frac{|r(s)|}{|r(s) - b|} = \frac{|r(s)-q|+|q|}{|r(s)-q| +|q| + |b|}.
\]
Hence it is enough for \(|r(s)-q| < |r(0)-q|\) for \(s >0\). Since \(q \neq 0\) is fixed this is equivalent to:
\[
\frac{|r(s) - q|}{|q|} < \frac{|r(0) - q|}{|q|}.
\]
For each \(s \in [0,1]\) choose \(a(s) \in \partial P\) such that \(w(s) \in \mathcal{N}(a(s))\). Write \(a = a(0)\) for convenience. We then have:
\[
\frac{|r(s)-q|}{|q|} = \frac{\langle a(s)-q, w \rangle }{\langle q,w \rangle}.
\]
Note \(n \in \mathcal{N}(q)\). Now \(\langle a(s)-q,n \rangle \le 0\) and \( \langle a(s) - q,w \rangle \le \langle a - q,w \rangle\). Clearly we have \(\langle q , n \rangle > 0\). Then:
\begin{align*}
\frac{|r(s)-q|}{|q|} &= \frac{\langle a(s) - q, w(s) \rangle }{\langle q, w(s) \rangle} \\ \\ &= \frac{s\langle a(s)-q, n \rangle + (1-s)\langle a(s)-q,w \rangle}{s \langle q , n \rangle + (1-s) \langle q, w \rangle} \\ \\ &\le \frac{(1-s)\langle a-q, w \rangle }{s \langle q , n \rangle + (1-s) \langle q, w \rangle} \\ \\ &< \frac{\langle a-q, w \rangle }{\langle q,w \rangle} \\ \\ &= \frac{|r(0) - q|}{|q|}.
\end{align*}
\end{proof}
Let \(V,P,b,q,\tau,n\) be as in the introduction to this section. Fix some open halfspace \(H \subset V^\vee\) given by \(u \ge 0\) for some \(u \in V \backslash \{0\}\). This defines a projection map \(p: V \to V/\langle u \rangle.\) Consider the function \(F_b: V^\vee \times [0,1] \to \RR\) given by:
\[
F_b(w,t) := t \langle b,w \rangle+ (1-t) \max_{x \in P} \langle x, w \rangle
\]
\begin{corollary} \label{cor:convexgeomextra}
For any \(W \subseteq V^\vee\) containing \(n\) we have:
\begin{equation} \label{R(X):1}
\sup (t \in [0,1] \ | \ \forall_{w \in W} \ F_b(t,w) \ge 0) = \frac{|q|}{|q-b|}.
\end{equation}
If for some choice of \(n\) we have \(n \not\in H\) then:
\begin{equation} \label{R(X):2}
\sup (t \in [0,1] \ | \ \forall_{w \in H} \ F_b(t,w) \ge 0) = \frac{|\tilde{q}|}{|\tilde{q} - p(b)|},
\end{equation}
where \(\tilde{q}\) is the intersection of the ray \(p(\tau)\) with the boundary of \(p(P)\).
\end{corollary}
\begin{proof}
Note that:
\[
\sup (t \in [0,1] \ | \ \forall_{w \in W} \ F_b(t,w) \ge 0) = \inf_{w \in W} \sup (t \in [0,1] \ | \ F_b(t,w) \ge 0).
\]
Moreover \(\sup (t \in [0,1] \ | \ F_b(t,w) \ge 0) = 1 > F_b(t,n)\) for \(\langle b,w \rangle \ge 0\), so without loss of generality we may assume \(W \subseteq \inte(\tau^\vee)\). For \(w \in W\) then:
\begin{align*}
\sup (t \in [0,1] \ | \ F_b(t,w) \ge 0) &= \frac{\max_{x \in P} \langle x, w \rangle}{ \max_{x \in P} \langle x, w \rangle - \langle b, w \rangle } \\ &= \frac{ \langle a,w \rangle}{\langle a ,w \rangle - \langle b,w \rangle} \\ &= \frac{ \langle p_w,w \rangle}{\langle p_w ,w \rangle - \langle b,w \rangle} = \frac{|p_w|}{|p_w-b|}.
\end{align*}
Hence:
\[
\sup (t \in [0,1] \ | \ \forall_{w \in W} \ F_b(t,w) \ge 0) = \inf_{w \in W} \frac{|p_w|}{|p_w-b|}.
\]
Now for \(w \in W\) consider the continuity path \(w(s) = sn + (1-s)w\). By Lemma \ref{R(X):Lemma3.1} if \(n \in W\) then the above infimum is attained when \(s=1\) and we obtain \normalfont{(}\ref{R(X):1}\normalfont{)}. Otherwise the infimum is attained at some \(w \in \partial W\). For \normalfont{(}\ref{R(X):2}\normalfont{)} restricting \(F_b\) to \(\partial H \times [0,1]\) gives:
\begin{align*}
F_b(w,t) = t \langle p(b) ,w \rangle+ (1-t) \max_{x \in p(P)} \langle x, w \rangle.
\end{align*}
Applying \normalfont{(}\ref{R(X):1}\normalfont{)} to the polytope \(p(P)\) in the vector space \(\partial H\) we obtain \normalfont{(}\ref{R(X):2}\normalfont{)}.
\end{proof}
\section{Proof of Theorem 4}
\subsection{Product Configurations}
Recall the formula \normalfont{(}\ref{eq:productDF}\normalfont{)} for the twisted Donaldson-Futaki invariant of a product configuration \(\X \cong X \times \A^1\).
Let \(q \in N_\RR\) be the point of intersection of the ray generated by \(-\bc(\Psi)\) with \(\partial \Box\). Applying \normalfont{(}\ref{R(X):1}\normalfont{)}, from Corollary~\ref{cor:convexgeomextra}, to \normalfont{(}\ref{eq:R(X)inf}\normalfont{)} we obtain:
\[
\sup(t | \DF_t(\X,\L,\cdot) \ge 0) = \frac{|q|}{|q-\bc_\nu(\Box)|}.
\]
\subsection{Non-Product Configurations}
Recall the description of special non-product test configurations of \(X\) from Section~\ref{sec:eqKstab}, and in particular the formula for the twisted Donaldson-Futaki in Lemma~\ref{lem:nonprodDF}. Set \(H := N_\QQ \times \RR^+\).
\begin{proposition}
For any non-product configuration \((\X,\L)\), with special fiber one of the \(\Delta_y\), let \(\sigma_y\) be the cone of outer normals to \(\Delta_y\) at the unique point of intersection of \(\partial \Delta_y\) with the ray generated by \(-\bc(\Delta_y)\). Denote this point of intersection by \(q_y\). Then:
\[
\sup(t | \DF_t(\X,\L,\cdot) \ge 0) =\begin{cases}
\frac{|q_y|}{|q_y - \bc(\Delta_y)|} & \sigma_y \cap H \neq \emptyset ;\\ \\
\frac{|q|}{|q-\bc_\nu(\Box)|} & \sigma_y \cap H = \emptyset.
\end{cases}
\]
\end{proposition}
\begin{proof}
Extend \(\DF_t(\X,\L,\cdot)\) linearly to the whole of \(N_\RR \times \RR\). In the case \(\sigma_y \cap H \neq \emptyset\) we may apply (1) from Corollary 2 with \(P = \Delta_y\) and \(b = \bc(\Delta_y)\). Otherwise we may apply \normalfont{(}\ref{R(X):2}\normalfont{)} from Corollary~\ref{cor:convexgeomextra}, noting that \(p(\Delta_y) = \Box\) and \(p(\bc(\Delta_y)) = \bc_\nu( \Box)\).
\end{proof}
\begin{proof}[Proof of Theorem~\ref{thm:R(X)}]
With Remark 2 in mind, observe that a special test configuration must either be product or non-product. Any non-product configurations \(\Delta_y\) with \(\sigma_y \cap H \neq \emptyset\) have their contribution to the infimum already accounted for and we may exclude them. The result follows.
\end{proof}
\begin{proof}[of Corollary~\ref{cor:R(X)}]
We observe \(\bc(\Delta_y) \in H \) for every special test configuration polytope \(\Delta_y\) of each threefold in this list, see \ref{appendix1}. We may then calculate \(R(X)\) using just the base polytope \(\Box\) and its Duistermaat-Heckman barycenter. The divisorial polytopes and Duistermaat-Heckman measures were originally given in \cite{suss2013fano}, and may be also be found in Appendix~\ref{appendix1}.
\end{proof}