1: %
2: % last and final version of infinite continued fractions---started by MAS,
3: % 2/5/05. Edited 15/8/05 DMJC
4: %
5: %
6: % Sort out \delta_1s, and statements of convolution stuff.
7: %
8: \documentclass[11pt,reqno]{amsart}
9: \setlength{\textheight}{23cm}
10: \setlength{\textwidth}{15.5cm}
11: \calclayout
12: \usepackage{amssymb}
13: %\usepackage[all]{xy}
14: %\usepackage{showkeys}
15: \newif\ifaddpics\addpicstrue% Comment out \addpicstrue to remove pictures.
16: \ifaddpics\usepackage{graphicx}\fi
17: %\usepackage{calrsfs}
18: %\usepackage{aeguill}
19: %\usepackage{diagrams}
20: %
21: %%%%%%%% Bibliography macros at end of file.
22: %
23: % Proclamation environments
24: %
25: %
26: \makeatletter
27: \def\swappedhead@plain#1#2#3{%
28: \thmnumber{\@upn{\mdseries #2}}\thmname{\@ifnotempty{#2}{. }#1}%
29: \thmnote{ \textmd{\upshape(#3)}}}
30: \makeatother
31: \theoremstyle{plain}
32: \newtheorem{result}{Theorem}
33: \renewcommand{\theresult}{\Alph{result}}
34: \swapnumbers
35: \newtheorem{thm}[subsection]{Theorem}
36: \newtheorem{lem}[subsection]{Lemma}
37: \newtheorem{prop}[subsection]{Proposition}
38: \newtheorem{cor}[subsection]{Corollary}
39: %
40: \newcommand{\emphdef}{\textit} % Emphasising definitions
41: \newcommand{\tcite}[1]{\textup{\cite{#1}}} % Theorem references
42: %
43: % Equation and item numbering
44: %
45: \numberwithin{equation}{section}
46: \renewcommand{\theequation}{\thesection.\arabic{equation}}
47: \renewcommand{\labelenumi}{\textup{(\roman{enumi})}}
48: \renewcommand{\theenumi}{\textup{(\roman{enumi})}}
49: %
50: \newenvironment{bulletlist}{\begin{list}{\labelitemi}%
51: {\setlength{\leftmargin}{\parindent}\def
52: \makelabel ##1{\hss \llap {\upshape ##1}}}}{\end{list}}
53: %
54: \newcommand{\acknowledge}{\subsection*{Acknowledgements}}
55: %
56: \newcommand{\thismonth}{\ifcase\month\or
57: January\or February\or March\or April\or May\or June\or
58: July\or August\or September\or October\or November\or December\fi
59: \space\number\year}
60: %
61: \newcommand{\sideremark}[1]{\marginpar{\small #1}}
62: %
63: % Lowering of subscripts and raising of superscripts
64: %
65: \makeatletter
66: \newcommand{\low}{\@ifnextchar^{}{^{\vphantom x}}}
67: \newcommand{\high}{\@ifnextchar_{}{_{\vphantom I}}}
68: \makeatother
69: %
70: % Euler script font and wedge symbol
71: %
72: \DeclareSymbolFont{script}{U}{eus}{m}{n}
73: \DeclareSymbolFontAlphabet{\mathscr}{script}
74: \DeclareMathSymbol{\EuWedge}{0}{script}{"5E}
75: %
76: % Slanted roman font for operators and symbols in math mode
77: %
78: \DeclareMathAlphabet{\mathrmsl}{OT1}{cmr}{m}{sl}
79: %
80: % Macros to create operators and symbols
81: % Operators automatically add space before the next symbol if needed
82: % Vertical centering of single character operators is suppressed using \null
83: %
84: \newcommand{\symb}[2]{\newcommand{#1}{{\mathit{#2}}}}
85: \newcommand{\rssymb}[2]{\newcommand{#1}{{\mathrmsl{#2}}}}
86: \newcommand{\calsymb}[2]{\newcommand{#1}{{\mathcal{#2}}}}
87: \newcommand{\bbsymb}[2]{\newcommand{#1}{{\mathbb{#2}}}}
88: %
89: \newcommand{\lieoper}[2]{\newcommand{#1}{\mathop
90: {\mathfrak{#2}\null}\nolimits}}
91: \newcommand{\oper}[3][n]{\newcommand{#2}{\mathop
92: {\mathrm{#3}\null}\ifx n#1\nolimits\else\limits\fi}}
93: \newcommand{\rsoper}[3][n]{\newcommand{#2}{\mathop
94: {\mathrmsl{#3}\null}\ifx n#1\nolimits\else\limits\fi}}
95: %
96: % Character shorthands
97: %
98: \bbsymb\C{C} \bbsymb\F{F} \bbsymb\IH{H}\bbsymb\N{N} \bbsymb\IP{P}
99: \bbsymb\Q{Q} \bbsymb\R{R} \bbsymb\U{U} \bbsymb\V{V} \bbsymb\W{W}
100: \bbsymb\Z{Z}
101: \bbsymb\E{E}
102: %
103: \newcommand{\ip}[1]{\langle #1 \rangle}
104: \renewcommand{\phi}{\varphi}
105: \newcommand{\ci}{C^\infty}
106: \newcommand{\ve}{\varepsilon}
107: \newcommand{\p}{\partial}
108: \newcommand{\hf}{{\mbox{\small$\frac{1}{2}$}}}
109: \newcommand{\thf}{{\mbox{\small$\frac{3}{2}$}}}
110: \newcommand{\qrt}{{\mbox{\small$\frac{1}{4}$}}}
111: \newcommand{\m}{\mbox{}}
112: \newcommand{\tr}{\mbox{tr}}
113: \newcommand{\ovnab}{\overline{\nabla}}
114: \newcommand{\oZ}{\overline{Z}}
115: \newcommand{\oD}{\overline{D}}
116: \newcommand{\tx}{\tilde{x}}
117: \newcommand{\ty}{\tilde{y}}
118: \newcommand{\ove}{\overline{e}}
119: \newcommand{\ovth}{\overline{\theta}}
120: \newcommand{\ovg}{\overline{g}}
121: \newcommand{\cD}{\mathcal D}
122: \newcommand{\cH}{\mathcal H}
123: \newcommand{\oH}{\overline{\mathcal H}}
124: \newcommand{\rd}{\mbox{d}}
125: \newcommand{\dv}{\mbox{div}}
126: \newcommand{\ric}{\mbox{ric}}
127: \newcommand{\supp}{\mbox{supp}}
128: \newcommand{\sign}{\mbox{sgn}}
129: %
130: \renewcommand{\geq}{\geqslant} \renewcommand{\leq}{\leqslant}
131: %
132: \theoremstyle{definition}
133: \newtheorem{dfn}[subsection]{Definition}
134: \newtheorem{rmk}[subsection]{Remark}
135: %
136: % Operator constructions
137: %
138: \newcommand{\del}{\partial} % directional derivative
139: \newcommand{\db}{\overline\partial} % del-bar derivative
140: \newcommand{\zb}{\overline z}
141: \newcommand{\ub}{\overline u}
142: %
143: %
144: %
145: \begin{document}
146: \title[Continued fractions and Einstein manifolds]{Continued fractions
147: and Einstein manifolds\\ of infinite topological type}
148: \author{David M. J. Calderbank}
149: \address{School of Mathematics\\
150: University of Edinburgh\\ King's Buildings, Mayfield Road\\ Edinburgh
151: EH9 3JZ\\ Scotland.}
152: \email{dc511@york.ac.uk}
153: \curraddr{Department of Mathematics\\ University of York\\ Heslington\\
154: York YO10 5DD\\ England.}
155: \author{Michael A. Singer}
156: \email{m.singer@ed.ac.uk}
157:
158: \begin{abstract} We present a construction of complete self-dual Einstein
159: metrics of negative scalar curvature on an uncountable family of manifolds of
160: infinite topological type.
161: \end{abstract}
162:
163: \maketitle
164: \section{Introduction}
165:
166: \subsection{Summary}
167:
168: Let $\alpha$ be an irrational number, $0<\alpha<1$, and consider the {\em
169: modified continued fraction expansion} of $\alpha$,
170: \begin{equation}\label{e1.9.6.5}
171: \alpha = \cfrac{1}{e_1-\cfrac{1}{e_2- \cfrac{1}{e_3-\cdots}}},
172: \end{equation}
173: with $e_j\geq 2$ for all $j$. We associate to $\alpha$ a noncompact
174: $4$-manifold $M_\alpha$, which is connected and simply connected, but has
175: $b_2(M_\alpha)= \infty$, $H_2(M_\alpha)$ being generated by an infinite
176: sequence of embedded $2$-spheres $S_j$, with
177: \begin{equation}\label{e2.9.6.5}
178: S_j \cdot S_{j+1} = -1,\; S_j\cdot S_j = e_j,\;
179: S_j\cdot S_k = 0\mbox{ for }|j-k|>1.
180: \end{equation}
181: Then, subject to the uniform bound
182: $3\leq e_j\leq N$, we construct a complete self-dual Einstein metric
183: $g_\alpha$, of negative scalar curvature, on $M_\alpha$. The key point about
184: the construction is that $M_\alpha$ and $g_\alpha$ are toric: there is a smooth
185: action of $T^2$ on $M_\alpha$ which preserves $g_\alpha$.
186:
187: The work in this paper complements that in our previous papers \cite{CaSi:emcs}
188: and \cite{CaSi:temco}. Indeed, in \cite{CaSi:emcs}, we associated to each {\em
189: rational} number $\alpha$, $0<\alpha<1$, a noncompact $4$-manifold $M_\alpha$
190: and, subject to the condition $e_j\geq 3$, we constructed a complete SDE metric
191: $g_\alpha$ on $M_\alpha$, where, as above, $M_\alpha$ is connected and simply
192: connected, but now $H_2(M_\alpha)$ is generated by a \emph{finite} sequence of
193: spheres $S_j$ which satisfy \eqref{e2.9.6.5}.
194:
195: We note a parallel in hyperk\"ahler geometry: the $A_n$-gravitational
196: instantons constructed by the Gibbons--Hawking Ansatz are analogous to the case
197: that $\alpha$ is rational, while the hyperk\"ahler manifolds of infinite
198: topological type in \cite{AKL:rfitt} correspond to $\alpha$ being irrational.
199: But there are very important differences: in the hyperk\"ahler case, the
200: self-intersections $e_j$ all have to be $2$, which is complementary to the
201: condition $e_j\geq 3$ that we impose here. On the other hand, despite the
202: uniform bound $e_j \leq N$, we obtain uncountably many non-diffeomorphic SDE
203: manifolds in this way---for this, it is enough to allow the $e_j$ to
204: take only the values $3$ and $4$, for example.
205:
206: The plan of this paper is as follows. In \S\ref{s2.14.6.5} we gather some
207: elementary facts about the continued fraction expansion \eqref{e1.9.6.5}. The
208: only result that may be new here is Theorem~\ref{t1.3.5.5}; we are grateful to
209: Chris Smyth for assisting us with its proof. In \S\ref{s10.14.6.5}, we give
210: the construction of $M_\alpha$. This is a straightforward extension of the work
211: in \cite{CaSi:emcs}, which, as we have indicated, corresponds to the case that
212: $\alpha$ is rational. (That work, in turn, rests on the combinatorial
213: description of toric $4$-manifolds due to Orlik and Raymond \cite{OrRa:at4}.)
214: In \S\ref{s11.14.6.5} we write down a SDE metric $g_\alpha$ on $M_\alpha$.
215: This is defined initially on a dense open subset $U\subset M_\alpha$ (the set
216: on which the $T^2$-action is free) but at the end of the section we show that
217: $g_\alpha$ extends smoothly to the whole manifold. The argument is given in
218: detail partly to make this paper more self-contained, and also to clarify one
219: point omitted from \cite{CaSi:emcs}\footnote{For readers of that paper, we did
220: not show in \cite[\S 5]{CaSi:emcs} that $\lim_{\rho\to 0}
221: \rho^{-1}|\det\Phi(\rho,\eta)| > 0$, and this is needed for the smooth
222: extension of the metric.}. In \S\ref{s1.5.7.5} we give some technical estimates
223: on the functions which enter the definition of $g_\alpha$. These are needed for
224: the smooth extension of $g_\alpha$ from $U$ to $M_\alpha$, and also pave the
225: way for the proof that $g_\alpha$ is complete in \S\ref{s1.9.8.5}.
226:
227: For the reader familiar with \cite{CaSi:emcs}, we make some remarks about the
228: extension of the construction of $g_\alpha$ from finite to infinite continued
229: fractions. From the point of view of \cite[\S4]{CaSi:emcs}, one would like to
230: allow $k$ to go to $\infty$ in
231: \begin{equation*}
232: \sqrt{\rho}F(\rho,\eta) = \sum_{j=0}^{k+1}w_j \sqrt{\rho^2 +(\eta- y_j)^2},
233: \end{equation*}
234: where
235: \begin{equation*}
236: w_j = m_{j+1}-m_j, \quad y_j = \frac{n_{j+1}-n_j}{m_{j+1}-m_j}
237: \end{equation*}
238: and the $n_j/m_j$ are the continued fraction approximants to
239: $\alpha$. But the sequence $w_j$ increases rapidly with $j$, while the
240: $y_j$ also converge to $\alpha$, so the status of such a limit is
241: unclear. However, we noted in \cite[\S5]{CaSi:emcs} that the
242: above sum represents an eigenfunction of the hyperbolic laplacian with
243: boundary data equal to $\eta\mapsto m_j \eta -n_j$ for $y_j \leq \eta
244: \leq y_{j-1}$ and this makes perfectly good sense also for infinite
245: continued fractions. This observation is really the key to the
246: construction in this paper.
247:
248: \subsection{Notation} \label{s1.14.6.5} Denote by $\cH^2$ the
249: hyperbolic plane, by $\oH{}^2$ its conformal compactification. We shall
250: always identify $\cH^2$ with the upper half-plane $\{(x,y)\in \R^2:
251: y>0\}$; then the hyperbolic metric is given by $(\rd x^2 +\rd
252: y^2)/y^2$ and $\oH{}^2 =\{(x,y): y \geq 0\}\cup\{\infty\}$. Note that
253: half-space coordinates near $\infty$ can be defined by
254: $$
255: (\tx, \ty) = \left(-\frac{x}{x^2+y^2}, \frac{y}{x^2+y^2}\right).
256: $$
257: Further, denote by $T^2$ the standard $2$-torus, identified with
258: $\R^2/(2\pi \Z)^2$. We shall write $z = (z_1,z_2)$ for standard
259: linear coordinates on $\R^2$. The circle-subgroup generated by
260: $m\del_{z_1} + n\del_{z_2}$ (for $m,n\in\Z$) will be denoted by $S^1_{(m,n)}$.
261: We shall denote by $\ve$ the standard skew form
262: \begin{equation}\label{e1.9.8.5}
263: \ve(z',z'') = \det(z',z'') = z_1'z''_2 - z'_2z''_1.
264: \end{equation}
265:
266: \subsection{Acknowledgement} We thank Chris Smyth for useful
267: conversations on continued fractions and for the proof of
268: Theorem~\ref{t1.3.5.5}. We also thank Jim Wright for useful conversations.
269:
270: \section{Continued fractions and toric $4$-manifolds}
271: \label{s2.14.6.5}
272: \subsection{Continued fractions}
273: \label{s3.14.6.5}
274: Let us begin with our irrational number $\alpha$ and its continued fraction
275: expansion \eqref{e1.9.6.5}. Set
276: \begin{equation}\label{e2.2.5.5}
277: (m_0,n_0) = (0,-1),\quad (m_0,n_0) = (1,0),
278: \end{equation}
279: and
280: \begin{equation}\label{e3.2.5.5}
281: (m_j,n_j)\mbox{ coprime with }
282: \frac{n_j}{m_j} =
283: \cfrac{1}{e_1-\cfrac{1}{e_2- \cdots \cfrac{1}{e_{j-1}}}}.
284: \end{equation}
285: The most important properties of this sequence of pairs $(m_j,n_j)$
286: are summarized as follows.
287: \begin{lem} \label{l2.14.6.5}
288: For each $j\geq 1$,
289: \begin{equation}\label{e6.2.5.5}
290: (m_{j+1},n_{j+1}) = e_j(m_j,n_j) - (m_{j-1},n_{j-1}),
291: \end{equation}
292: and for $j\geq 0$,
293: \begin{equation}\label{e7.2.5.5}
294: m_{j}n_{j+1} - m_{j+1}n_{j} = 1.
295: \end{equation}
296: \end{lem}
297: \begin{proof}
298: If
299: \begin{equation}\label{e5.2.5.5}
300: M_j = \begin{pmatrix} 0& 1 \cr -1 & e_j\end{pmatrix}
301: \end{equation}
302: then one has, for each $j$:
303: \begin{align}
304: \begin{pmatrix} n_{j-1} \cr m_{j-1} \end{pmatrix} &=
305: M_1M_2\cdots M_{j-1}
306: \begin{pmatrix} -1\cr 0 \end{pmatrix}; \label{e1.14.6.5} \\
307: \begin{pmatrix} n_j \cr m_j \end{pmatrix} &=
308: M_1M_2\cdots M_{j-1}
309: \begin{pmatrix} 0\cr 1 \end{pmatrix};\label{e4.2.5.5} \\
310: \begin{pmatrix} n_{j+1} \cr m_{j+1} \end{pmatrix} &=
311: M_1M_2\cdots M_{j-1}
312: \begin{pmatrix} 1\cr e_j \end{pmatrix}. \label{e2.5.6.5}
313: \end{align}
314: These are all essentially equivalent to each other and are easily
315: proved by induction. Combining these, we have
316: \begin{equation}\label{e1.5.6.5}
317: \begin{pmatrix} n_{j+1} & n_{j} \cr
318: m_{j+1} & m_{j} \end{pmatrix} = M_1M_2\cdots M_{j-1}
319: \begin{pmatrix} 1 & 0 \cr e_j & 1 \end{pmatrix},\quad
320: \begin{pmatrix} n_{j+1} & n_{j-1} \cr
321: m_{j+1} & m_{j-1} \end{pmatrix} = M_1M_2\cdots M_{j-1}
322: \begin{pmatrix} 1 & -1 \cr e_j & 0 \end{pmatrix},
323: \end{equation}
324: and, taking determinants, we obtain \eqref{e7.2.5.5} and the identity
325: \begin{equation}\label{e8.2.5.5}
326: m_{j-1}n_{j+1} - m_{j+1}n_{j-1} = e_j.
327: \end{equation}
328: From \eqref{e7.2.5.5} (with $j$ replaced by $j-1$) it follows that
329: $(m_{j-1},n_{j-1})$ and $(m_j,n_j)$ form a $\Z$-basis of
330: $\Z\oplus\Z$, so that $(m_{j+1},n_{j+1})$ is an integer linear combination
331: of these vectors. The coefficients are
332: determined as in \eqref{e6.2.5.5} by using the identities
333: \eqref{e7.2.5.5} and \eqref{e8.2.5.5}.
334: \end{proof}
335:
336: \subsection{Definition} Set
337: \begin{equation}\label{e9.2.5.5}
338: a_j = \frac{n_{j+1}-n_j}{m_{j+1}-m_j},\quad b_j = \frac{1}{m_{j+1}-m_j}.
339: \end{equation}
340: These are well-defined by the following lemma.
341: \begin{lem} \label{l1.14.6.5}
342: \textup{(i)} For all $j\geq 1$, $m_{j+1}> m_j$, $n_{j+1}>n_j$, and $(n_j/m_j)$
343: is strictly increasing with limit $\alpha$. Hence the sequences $(a_j)$ and
344: $(b_j)$ are positive. Furthermore, $\lim_{j\to\infty} a_j=\alpha$.
345:
346: \noindent\textup{(ii)} Suppose $e_j\geq 3$ for all $j$. Then $m_{j+1} > \phi^2
347: m_j$ and $n_{j+1} > \phi^2 n_j$, where $\phi=(1+\sqrt 5)/2>1$ is the golden
348: ratio. Furthermore, $(a_j)$ and $(b_j)$ are strictly decreasing, and
349: $\lim_{j\to\infty} b_j=0$.
350: \end{lem}
351: \begin{proof} (i) To prove that the sequences $(m_j)$ and $(n_j)$ are
352: strictly increasing, use \eqref{e6.2.5.5} and induction on $j$. In particular
353: $m_j>0$ for $j>0$. The monotonicity of the sequence $n_j/m_j$ now follows from
354: the identity
355: \begin{equation}\label{e2.14.6.5}
356: \frac{n_{j+1}}{m_{j+1}} - \frac{n_{j}}{m_{j}}
357: = \frac{1}{m_jm_{j+1}},
358: \end{equation}
359: where we have used \eqref{e7.2.5.5}. The fact that $n_j/m_j\to\alpha$ as
360: $j\to\infty$ is standard. Now note that
361: \begin{equation}\label{e3.3.5.5}
362: a_j -\frac{n_j}{m_j} = \frac{1}{m_j(m_{j+1}-m_j)}>0
363: \end{equation}
364: so that $\lim_{j\to\infty} a_j = \lim_{j\to\infty} n_j/m_j = \alpha$ as
365: required.
366:
367: (ii) We again use~\eqref{e6.2.5.5} and induction on $j$: e.g., if $m_j>\phi^2
368: m_{j-1}$, then
369: \begin{equation*}
370: m_{j+1} = e_j m_j - m_{j-1} > (3-\phi^{-2})m_j =\phi^2 m_j,
371: \end{equation*}
372: since $\phi^2=(3+\sqrt 5)/2$. Now from \eqref{e6.2.5.5} and the nonnegativity
373: of $(m_j)$,
374: \begin{equation*}
375: m_{j+1} -m_j = (e_j-1)m_j - m_{j-1} \geq (e_j-1)(m_j -m_{j-1}).
376: \end{equation*}
377: Since $e_j-1>1$, the statements about $(b_j)$ are immediate. On the other hand,
378: \begin{equation}\label{e1.3.5.5}
379: a_{j-1} - a_{j} = \frac{e_j-2}{(m_{j+1}-m_j)(m_j-m_{j-1})}>0
380: \end{equation}
381: which gives the monotonicity of the $a_j$.
382: \end{proof}
383:
384: \subsection{Assumption}
385: \label{s5.14.6.5} From now on, suppose that for all $j$, $3\leq e_j
386: \leq N$ for some $N>0$.
387:
388: \subsection{Definition} The {\em envelope} $\eta_\alpha$ of $\alpha$
389: is defined as
390: \begin{equation}\label{e4.3.5.5}
391: \eta_{\alpha}(x) = \begin{cases} m_jx -n_j &
392: \mbox{for }x\in [a_j,a_{j-1}],\\
393: 0& \mbox{for }x\leq \alpha.
394: \end{cases}
395: \end{equation}
396: (Here we set $a_{-1}= \infty$, so that $\eta_{\alpha}(x) =1$ for $x\geq a_0 =
397: 1$.)
398: \begin{prop}\label{eta-prop} $\eta_\alpha$ is continuous, and for $x>\alpha$ it
399: is strictly increasing \textup(hence positive\textup), concave \textup(i.e.,
400: $\eta''_\alpha \leq 0$ in the sense of distributions\textup) and is the linear
401: interpolant of the points $(a_j,b_j)$ \textup(so that $\eta_\alpha(a_j)=b_j$
402: for all $j$\textup).
403: \end{prop}
404: \begin{proof} Trivial, given the previous results.\end{proof}
405:
406: \subsection{Example}
407: \label{s6.14.6.5} Suppose $e_j=3$ for all $j$. Then
408: \begin{equation*}
409: m_j = n_{j+1} = \frac{\phi^{2j} - \phi^{-{2j}}}{\sqrt{5}},\quad
410: a_j=\frac{\phi^{2j-1}+\phi^{-2j+1}}{\phi^{2j+1}+\phi^{-2j-1}}, \quad
411: b_j=\frac{\sqrt 5}{\phi^{2j+1}+\phi^{-2j-1}},
412: \end{equation*}
413: where $\phi = (1 +\sqrt{5})/2$ as above. Then $\alpha
414: =\phi^{-2}=0.381966\ldots$ and
415: \begin{equation*}
416: \frac{a_j - \phi^{-2}}{b_j^2} = \frac{1}{\sqrt{5}}
417: \biggl(\frac{\phi^{4j+2}+2+\phi^{-4j-2}}{\phi^{4j+2}+1}\biggr).
418: \end{equation*}
419: In particular, by Proposition~\ref{eta-prop},
420: \begin{equation} \label{e2.9.8.5}
421: \eta_{\alpha}(x) \lesssim \sqrt{\sqrt{5}(x-\phi^{-2})}
422: \end{equation}
423: for $x-\phi^{-2}$ small. The two functions in \eqref{e2.9.8.5} are shown in
424: Figure~\ref{fig1}.
425: %
426: \ifaddpics
427: \begin{figure}[ht]
428: \begin{center}
429: %
430: \includegraphics[width=.7\textwidth]{cfmfig1.eps}
431: %
432: \caption{The envelope and the square-root function with $\alpha=\phi^{-2}$}
433: \label{fig1}
434: \end{center}
435: \end{figure}
436: \fi
437: %
438: \smallbreak
439:
440: We shall now show that the behaviour of $\eta_\alpha(x)$ is bounded by a
441: multiple of $\sqrt{x-\alpha}$ in general---we are indebted to Chris Smyth for
442: the proof of this fact.
443:
444: \begin{thm} \label{t1.3.5.5}
445: There is a constant $\Omega>0$ such that
446: \begin{equation}\label{eta-bounds}
447: \eta_\alpha(x) \leq \Omega\sqrt{x-\alpha} \mbox{ for }x\geq\alpha.
448: \end{equation}
449: \end{thm}
450: \begin{proof} This is a sequence of elementary deductions, based on
451: Lemmas~\ref{l2.14.6.5} and \ref{l1.14.6.5}.
452: Start by substituting
453: $m_{j+1} =e_jm_j -m_{j-1}$ into
454: \eqref{e1.3.5.5}, to get
455: \begin{equation}\label{11.19.12}
456: a_{j-1} - a_{j} = \left[\frac{e_{j}-2}{(1 - m_{j-1}/m_{j})
457: (e_{j} -1 - m_{j-1}/m_{j})}\right]\frac{1}{m_{j}^2}.
458: \end{equation}
459: Because $0 < m_{j-1}/m_{j} <\phi^{-2}<1/2$ and $e_j\geq 3$ it is elementary to
460: bound the quantity in square brackets between $1/2$ and $2$. Hence
461: \begin{equation}\label{12.19.12}
462: \frac{1}{2m_{j}^2}< a_{j-1} - a_{j} < \frac{2}{m_{j}^2}.
463: \end{equation}
464: Now
465: \begin{equation}\label{e4.3.6.5}
466: a_n- \alpha = \sum_{j=n+1}^\infty (a_{j-1} - a_{j})
467: \end{equation}
468: and since $m_j > 2m_{j-1}$ for all $j$, we obtain
469: \begin{equation}\label{13.19.12}
470: \frac{1}{2m_{n+1}^2} < a_n -\alpha
471: < \frac{2}{m_{n+1}^2}\sum_{j=0}^\infty \phi^{-4j} <\frac{3}{m_{n+1}^2}.
472: \end{equation}
473: (The lower bound is the trivial one coming from the first term in the sum.)
474:
475: Since $b_{j-1} - b_{j} = m_j(a_{j-1} - a_{j})$ we find from \eqref{12.19.12}
476: \begin{equation}\label{14.19.12}
477: \frac{1}{2m_{j+1}}< b_{j-1} - b_{j} < \frac{2}{m_{j+1}};
478: \end{equation}
479: summing as before, we obtain
480: \begin{equation}\label{15.19.12}
481: \frac{1}{2m_{n+1}} < b_n < \frac{2}{m_{n+1}}\sum_{j=0}^\infty \phi^{-2j}
482: < \frac{4}{m_{n+1}}.
483: \end{equation}
484: Combining \eqref{13.19.12} and \eqref{15.19.12} in the obvious way we
485: obtain
486: \begin{equation}\label{16.19.12}
487: \frac{1}{2\sqrt 3}\sqrt{a_n-\alpha} < b_n < 4\sqrt 2\sqrt{a_n - \alpha}.
488: \end{equation}
489: Since $\eta_\alpha$ is piecewise linear, $\eta_\alpha(a_n) =b_n$, and the
490: square-root function is concave, the bound~\eqref{eta-bounds} follows at once.
491: \end{proof}
492:
493: The following technical result will be needed in \S\ref{s1.9.8.5}. In order to
494: state it, set
495: \begin{equation}\label{e1.10.8.5}
496: \mu(x) = m_j\mbox{ if $a_j < x < a_{j-1}$},\quad
497: \nu(x) = n_j\mbox{ if $a_j < x < a_{j-1}$},
498: \end{equation}
499: and
500: \begin{equation}\label{e2.10.8.5}
501: D(x_1,x_2) = (\mu(x_1)\nu(x_2) -\mu(x_2)\nu(x_1))(x_1-x_2).
502: \end{equation}
503:
504: \begin{lem}\label{l1.11.8.5}
505: If $3 \leq e_j \leq N$ for all $j$, then
506: \begin{equation}\label{e3.10.8.5}
507: D(\alpha+ x, \alpha+ x\eta) \geq x(1-\eta)\mbox{ for }x\in (0,1-\alpha)),
508: \eta \in(0,(4N)^{-2}).
509: \end{equation}
510: \end{lem}
511: \begin{proof}
512: We have
513: \begin{equation} \label{e11.5.7.5}
514: D(\alpha+x,\alpha+ x\eta) = [\mu(\alpha + x)\nu(\alpha + x\eta) -
515: \mu(\alpha + x\eta)\nu(\alpha + x)]x(1-\eta);
516: \end{equation}
517: since the quantity in square brackets is $\geq 1$ if $\alpha+x$ and
518: $\alpha +x\eta$ are in disjoint intervals $[a_j,a_{j-1}]$,
519: it is enough to show that if $x$ and $\eta$ are as in
520: \eqref{e3.10.8.5} and
521: \begin{equation}\label{e1.11.8.5}
522: a_j<\alpha+ x< a_{j-1},
523: \end{equation}
524: then $\alpha + x\eta < a_j$. Now if \eqref{e1.11.8.5} is satisfied,
525: then we have $x < 4/m_{j}^2$ from \eqref{13.19.12}, and so
526: \begin{equation}\label{e2.11.8.5}
527: x\eta < \frac{1}{4N^2m_{j}^2} < \frac{1}{4m_{j+1}^2} < a_{j}-\alpha;
528: \end{equation}
529: this follows from \eqref{e3.10.8.5},
530: \eqref{e6.2.5.5} (which implies $m_{j+1} < N m_j$) and
531: \eqref{13.19.12}. In view of the previous remarks, the proof is now complete.
532: \end{proof}
533:
534:
535: \section{Construction of $M_\alpha$}
536: \label{s10.14.6.5}
537: In this section, we associate to any irrational number $\alpha$,
538: $0<\alpha<1$, a toric $4$-manifold $M_\alpha$ of infinite topological
539: type.
540:
541: \subsection{Notation and set-up} \label{s2.5.6.5}The two points
542: $\alpha$ and $\infty$ decompose $\partial \cH^2$ as
543: \begin{equation*}
544: \del\cH^2 = \del\cH^2_+\cup\del\cH^2_- \cup\{\alpha\}\cup\{\infty\},
545: \end{equation*}
546: where
547: \begin{equation*}
548: \del\cH^2_+ = \{(x,0)\in\oH{}^2: x>\alpha\},\quad
549: \del\cH^2_- = \{(x,0)\in\oH{}^2: x<\alpha\}.
550: \end{equation*}
551: Choose a smooth simple arc $\oZ$ in $\oH{}^2$ which joins $\alpha$ to
552: $\infty$ and is such that
553: \begin{equation*}
554: \oZ = Z\cup\{\alpha\}\cup\{\infty\},\quad Z\subset \cH^2.
555: \end{equation*}
556: Then $Z$ decomposes $\cH^2$ as a disjoint union
557: \begin{equation*}
558: \cH^2 = D_+\cup Z\cup D_-,
559: \end{equation*}
560: where $D_\pm$ contains $\del \cH^2_{\pm}$ in its closure. Finally, put
561: \begin{equation*}
562: \oD_{\pm} = D_{\pm}\cup \del \cH^2_{\pm}.
563: \end{equation*}
564: The reader is urged to note that $\oD_+$ does not contain $Z$ or
565: either of the points $\alpha$ and $\infty$.
566:
567: \subsection{Notation} The boundary component $\del \cH^2_+$ is
568: decomposed into the intervals $[a_j,a_{j-1}]$. We shall refer to
569: these as {\em edges} and to the $a_j$ themselves as {\em corners}.
570:
571: \subsection{Construction of $M_\alpha$} Recall the combinatorial
572: description of smooth toric $4$-manifolds of Orlik and Raymond
573: \cite{OrRa:at4}. Let $M$ be a compact, simply connected $4$-manifold, with a
574: smooth action of the 2-torus $T^2$, free on the open subset $U\subset M$. Let
575: $\pi:M \to M/T^2 = P$ be the quotient map. Then $P$ is a topological polygon
576: with a finite number of edges $E_j$, say. If $x$ is a point in the interior of
577: $P$, then $\pi^{-1}(x) = T^2$, while if $x\in \mathrm{int}\, E_j$, then
578: $\pi^{-1}(x)$ is a circle. This circle is a special orbit of the $T^2$-action,
579: with isotropy group $S^1_{\smash{(m_j,n_j)}}$. If $x$ is a corner $E_j\cap
580: E_{j+1}$ of $P$, then $\pi^{-1}(x)$ is a point, fixed by the $T^2$-action.
581:
582: More important for our purposes is the converse construction: to a polygon $P$,
583: with edges $E_j$, labelled by coprime pairs $(m_j,n_j)$, one can construct a
584: toric $4$-orbifold $M$ such that $M/T^2$ is $P$, with special orbits described
585: in the previous paragraph; for $M$ to be a smooth manifold, we require that the
586: pairs $(m_j,n_j)$ and $(m_{j+1}, n_{j+1})$ form a $\Z$-basis for $\Z\oplus \Z$
587: for each $j$. One way to understand this construction is as follows. Starting
588: from the closed polygon $P$, form the product $P\times T^2$. For each point $x
589: \in \partial P$, the labelling gives us a subtorus $T_x$ of $T^2$, which is a
590: circle if $x$ is in the interior of an edge and is $T^2$ itself if $x$ is a
591: corner. The manifold $M$ is formed by contracting $T_x$ to a point. A smooth
592: local model for these `contractions' is given by the toric description of
593: $\R^4$. Using polar coordinates, the $T^2$-action is
594: \begin{equation*}
595: (z_1,z_2)\cdot (r_1e^{i\theta_1}, r_2e^{i\theta_2}) =
596: (r_1e^{i(\theta_1+z_1)}, r_2e^{i(\theta_2+z_2)}).
597: \end{equation*}
598: The quotient space is the closed quadrant $Q = \{(r_1,r_2): r_1\geq 0,
599: r_2\geq 0\}$. We have maps
600: \begin{equation}\label{e1.20.6.5}
601: Q \times T^2 \stackrel{\beta}{\longrightarrow}
602: \R^4\stackrel{\pi_c}{\longrightarrow} Q
603: \end{equation}
604: where $\pi_c$ is the quotient map $R^4\to\R^4/T^2=Q$ and
605: \begin{equation*}
606: \beta(r_1,r_2,\theta_1,\theta_2) = (r_1e^{i\theta_1},
607: r_2e^{i\theta_2})
608: \end{equation*}
609: contracts $S^1_{0,1}$ along the $r_1$-axis, $S^1_{1,0}$ along the $r_2$-axis,
610: and the whole $T^2$ at the corner of $Q$. Returning to the construction of
611: $M$, if $x\in \partial P$ is in the interior of an edge, with isotropy group
612: $S^1_{(m,n)}$, then we can change basis in the lattice so that $(m,n)$ is
613: mapped to $(0,1)$. Then a neighbourhood $N$ of $x$ in $P$ can be identified
614: with a neighbourhood $N'$ of $r_1=1$, say, on the $r_1$-axis in $Q$. The part
615: of $M$ over $N$ can then be defined to be $\pi_c^{-1}(N')$. Similarly, if
616: $x$ is a corner of $P$, then we can change basis in the lattice so that the
617: labels of the two edges through $x$ are mapped to $(1,0)$ and $(0,1)$. Again,
618: we identify a small neighbourhood $N$ of $x$ with a neighbourhood $N'$ of $0$
619: in $Q$ and construct the part of $M$ over $N$ as $\pi^{-1}_c(N')$. This
620: concludes our outline of the reconstruction of a toric $4$-manifold from a
621: labelled polygon.
622:
623: With this understood, we can describe $M_\alpha$ using the notation introduced
624: in \S\ref{s2.5.6.5}. That is, we form the product $\oD_+\times T^2$, and
625: contract $S^1_{(m_j,n_j)}$ over each point of $[a_j,a_{j-1}]$; the result is an
626: open manifold that we shall call $M_\alpha$. Notice that a sequence of points
627: $p_j\in M_\alpha$ with $\pi(p_j)$ converging to a point on $\oZ$ will have no
628: convergent subsequence in $M_\alpha$. On the other hand, if $\pi(p_j)$
629: converges to a point on $\del \cH^2_+$, then $p_j$ {\em does} have a convergent
630: subsequence.
631:
632: \subsection{Remark} Note that for $j\geq 1$, $S_j = \pi^{-1}[a_j,a_{j-1}]$
633: is a smoothly embedded $2$-sphere. Suitably oriented, we have
634: \begin{equation*}
635: S_j\cdot S_{j+1} = - 1,\quad S_j\cdot S_j = e_j.
636: \end{equation*}
637:
638:
639: \section{A self-dual Einstein metric on $M_\alpha$}
640: \label{s11.14.6.5}
641:
642: For background to the material presented in this section, the reader is
643: referred to \cite{CaPe:emt} and \cite{CaSi:emcs}. We present the essential
644: formulae here: the proofs can be found in the cited papers.
645:
646: Let $F(x,y)$ be a solution of the equation
647: \begin{equation}\label{e3.9.8.5}
648: \Delta F = \tfrac{3}{4}F%\;\;\lim_{y\to 0}[y^{1/2}F(x,y)] \to f_0(x).
649: \end{equation}
650: where $\Delta = y^2(\del_x^2 + \del_y^2)$ is the Laplacian of
651: $\cH^2$. Let
652: \begin{equation*}
653: f(x,y) = \sqrt{y}F(x,y)
654: \end{equation*}
655: and introduce the auxiliary quantities
656: \begin{equation*}
657: v_1 = (f_y,xf_y -y f_x),\quad v_2 = (f_x, xf_x +yf_y -f)
658: \end{equation*}
659: and
660: \begin{equation*}
661: w = f_x^2 + f_y^2 - y^{-1}ff_y, \mbox{ so that }\ve(v_1,v_2) = yw.
662: \end{equation*}
663: (Recall from \eqref{e1.9.8.5} that $\ve$ stands for the standard
664: symplectic form on $\R^2$.)
665: Set
666: \begin{equation*}
667: g = \frac{|w|}{f^2}\left(\rd x^2 +\rd y^2 + \frac{\ve(v_1,\rd z)^2
668: +\ve(v_2,\rd z)^2}{w^2}\right).
669: \end{equation*}
670: This is a self-dual Einstein metric on $U\times T^2$, where $U\subset
671: \cH^2$ is the set on which $w\not=0$, $f\not=0$. Moreover, the sign of
672: the scalar curvature is opposite to the sign of $w$.
673:
674: The next task is to choose the eigenfunction $F$ so as to obtain a
675: metric that extends smoothly over the special $T^2$-orbits. This is
676: very conveniently done using the following integral formulae which
677: show how to recover $F$ from its (renormalized) boundary value $u(x)=f(x,0)$.
678:
679: \subsection{Integral formulae}
680: Let $u$ be a distribution on
681: $\R$, viewed as the finite part of $\del \cH^2$. If $u$ does not
682: grow too fast at $\infty$, we define
683: \begin{equation}\label{e11.3.5.5}
684: f(x,y) = \int \frac{y^2}{2((x-x_1)^2 +y^2)^{3/2}}u(x_1)\,\rd x_1 ,
685: \end{equation}
686: which we shall also write as
687: \begin{equation} \label{e12.3.5.5}
688: f(x,y) = [k_y*u](x,y), \mbox{ where } k_y(x) = \frac{y^2}{2(x^2+y^2)^{3/2}}.
689: \end{equation}
690: Then $F(x,y) = y^{-\hf}f(x,y)$ satisfies \eqref{e3.9.8.5} and has
691: boundary data given by $u$ in the sense that
692: \begin{equation*}
693: f(x,y)\to u(x)\mbox{ as }y\to 0
694: \end{equation*}
695: in the sense of distributions.
696:
697: If $f$ is given by \eqref{e11.3.5.5}, there is also a formula for $w$
698: in terms of the boundary value $u$.
699: For this, introduce
700: \begin{equation*}
701: (\mu(x),\nu(x)) = (u'(x), xu'(x) -u(x)),
702: \end{equation*}
703: so that $(\mu,\nu)$ is the boundary value of $v_2$ and
704: $\mu(x)x-\nu(x) = u(x)$.
705: If
706: \begin{equation} \label{e2.3.6.5}
707: D(x_1,x_2) = (\mu(x_1)\nu(x_2)-\mu(x_2)\nu(x_1))(x_1-x_2),
708: \end{equation}
709: then
710: \begin{equation}\label{e11.6.5.5}
711: w(x,y) =\hf y^{-2}\int\!\!\int k_y(x-x_1)k_y(x-x_2)D(x_1,x_2)\,\rd x_1 \rd x_2.
712: \end{equation}
713: This is essentially \cite[(5.16)]{CaSi:emcs}.
714:
715: \subsection{Remark} At the beginning of this section we made the
716: assumption that $u$ should not grow too fast at $\infty$. Our
717: explicit choice of coordinates obscures the invariance of the
718: preceding formulae. It is more natural to interpret $u$ as a
719: distributional section of the $-\hf$-power of the density bundle on
720: $\del \cH^2$. Then the above integral formulae become fully invariant
721: under $PSL_2(\R)$, acting by isometries on $\cH^2$ and projectively on
722: its boundary. The distributions that we shall
723: actually use will satisfy $u(\pm x) = \pm 1$ for all sufficiently
724: large $|x|$; it is not difficult to check that $u(x)|\rd
725: x|^{-\hf}$ is then smooth in a neighbourhood of $\infty$.
726:
727: \subsection{Definition of $g_\alpha$}
728: In order to obtain a metric on $M_\alpha$, apply the previous formulae with
729: $u$ equal to the odd extension of
730: $\eta$ to the left of $\alpha$:
731: \begin{equation*}
732: u(x) = \eta(x) - \eta(2\alpha-x).
733: \end{equation*}
734: It is clear from \eqref{e11.3.5.5} that $f(\alpha,y)=0$ for all $y\geq
735: 0$; accordingly we set
736: $Z = \{(\alpha,y):y>0\}$ (cf.\ \S\ref{s2.5.6.5}).
737:
738: \begin{thm} The metric $g_\alpha$ is defined on $D_+\times T^2$ and
739: extends smoothly to $M_\alpha$.
740: \end{thm}
741: \begin{proof} See \cite[Theorem 5.2.1]{CaSi:emcs}. That result
742: applies to show that $w>0$ in $\cH^2$ and that $f>0$ in $D_+$, so that
743: $g_\alpha$ is defined on $D_+\times T^2$. We explain why the metric extends to
744: $M_\alpha$ since this was not done in \cite{CaSi:emcs} and we did not
745: explicitly check that $w$ has the needed boundary behaviour. This is now fixed
746: in Proposition~\ref{p1.5.7.5} below.
747:
748: For this, we need to understand the asymptotic behaviour of $f$ and $w$ as
749: $y\to 0$. So pick $x_0>\alpha$. We distinguish two cases according to whether
750: $x_0$ is or is not one of the $a_j$. The easier case corresponds to $a_j < x_0
751: < a_{j-1}$ for some $j$. Then by Proposition~\ref{p1.6.6.5}, with $m_1=m_2 =
752: m_j$, $n_1=n_2=n_j$, we have
753: \begin{equation}\label{e2.5.7.5}
754: f(x,y) = m_jx-n_j +y^2f_1(x) +\cdots,
755: \end{equation}
756: and by part (i) of Proposition~\ref{p1.5.7.5},
757: \begin{equation}\label{e1.5.7.5}
758: w(x,y) = w_0(x) + w_1(x)y^2 +\cdots, \mbox{ where }w_0(x)>0,
759: \end{equation}
760: for $|x-x_0|$ and $y\geq 0$ sufficiently small.
761: Write $(m_j,n_j) = (m,n)$, to simplify the notation. Then computing with
762: these formulae,
763: \begin{align*}
764: v_1 &= y(2f_1,2xf_1 - m) + O(y^3)\\
765: v_2 &= (m,n) + O(y^2)\\
766: \ve(v_1,v_2) &= y(2uf_1 -(u')^2).
767: \end{align*}
768: Thus, for small $y$,
769: \begin{equation*}
770: g_\alpha \simeq
771: \frac{w_0}{u^2}\left(
772: \rd x^2 +\rd y^2 + \frac{
773: y^2(2f_1\rd z_2 + (u'-2xf_1)\rd z_1)^2
774: + (m\rd z_2 -n\rd z_1)^2}{(2uf_1 - (u')^2)}\right)
775: \end{equation*}
776: In particular
777: \begin{equation*}
778: g_\alpha(m\del_{z_1}+n\del_{z_2},
779: m\del_{z_1}+n\del_{z_2}) \simeq y^2.
780: \end{equation*}
781: If we introduce new coordinates
782: \begin{align*}
783: \rd \theta &= m\rd z_2 - n\rd z_1 \\
784: \rd \psi &= m_1\rd z_2 - n_1\rd z_1,
785: \end{align*}
786: where $|mn_1-m_1n|=1$, we have
787: \begin{equation*}
788: g_\alpha = \frac{w_0}{u^2}
789: \left( \rd x^2 + a(x)\rd\theta^2 + \rd y^2
790: + y^2\rd \psi^2+\cdots)\right),\quad (a(x)>0)
791: \end{equation*}
792: which shows that $g_\alpha$ does extend as a smooth metric to
793: $\pi^{-1}(U)$, where $U$ is a small neighbourhood of the
794: boundary-point $x_0$.
795:
796: To complete the proof, we must consider the behaviour of $g_\alpha$ at
797: one of the corners $a_j$. By a suitable change of variables, we
798: may assume that $x_0 = a_0 = 1$, so that $u(x) =x$ for $1-\delta<x
799: \leq 1$, $u(x) =1$ for $1\leq x < 1+\delta$.
800:
801: By Proposition~\ref{p1.6.6.5}, we have
802: \begin{equation*}
803: f(x,y) =1 + \hf(x-1) - \hf((x-1)^2+y^2)^{1/2} + f_1(x)y^2 +\cdots
804: \end{equation*}
805: and by part (ii) of Proposition~\ref{p1.5.7.5}
806: \begin{equation}\label{e4.5.7.5}
807: w(x,y) = \hf((x-1)^2+y^2)^{-1/2} + w_1(x) + \cdots
808: \end{equation}
809: where the omitted terms contain only even powers of $y$. Using the
810: change of variable
811: \begin{equation*}
812: x-1 + iy = (r_2 + ir_1)^2
813: \end{equation*}
814: we compute, to leading order,
815: \begin{align*}
816: v_1 &= -\frac{r_1r_2}{r_1^2+r_2^2}(1,1) + \cdots\\
817: v_2 &= \frac{1}{r_1^2+r_2^2}(r_1^2,-r_2^2) + \cdots\\
818: w &= \frac{1}{2}\frac{1}{r_1^2+r_2^2}+\cdots\\
819: \rd x^2 + \rd y^2 &=4(r_1^2+r_2^2)(\rd r_1^2 +\rd r_2^2).
820: \end{align*}
821: Substituting these into the formula for $g_\alpha$ we find
822: \begin{equation*}
823: g_\alpha = (\rd r_1^2 +r_1^2 \rd z_1^2) +
824: (\rd r_2^2 +r_2^2 \rd z_2^2) +\cdots
825: \end{equation*}
826: exactly as required for smooth extension to $M_\alpha$ in a
827: neighbourhood of the corner.
828: \end{proof}
829:
830:
831: \section{Boundary behaviour of $f$ and $w$}\label{s1.5.7.5}
832:
833: This section is devoted to a study of the integral formulae for $f$
834: and $w$, \eqref{e11.3.5.5} and \eqref{e11.6.5.5}. As in the previous
835: section, we assume that $u$
836: is a distribution on $\R$ such that
837: \begin{equation} \label{e2.6.6.5}
838: \mbox{$u$ is equal to a constant multiple of $\sign(x)$ outside a compact set.}
839: \end{equation}
840: Since
841: \begin{equation} \label{e3.6.6.5}
842: K_y ''(x) = k_y(x),\quad K_y(x) =\hf \sqrt{x^2+y^2},
843: \end{equation}
844: we can rewrite \eqref{e11.3.5.5} as
845: \begin{equation} \label{e4.6.6.5}
846: f(x,y) = \int K_y(x-x_1)u''(x_1)\,\rd x_1,
847: \end{equation}
848: the condition \eqref{e2.6.6.5} being used to justify the integration by parts.
849:
850: We now start our study of the behaviour of \eqref{e11.3.5.5} as $y\to
851: 0$. For this, fix $\delta >0$, and define
852: \begin{equation}\label{e5.6.6.5}
853: B_0 = (-\delta,\delta) \times (0,\delta),\quad
854: B = (-\delta,\delta) \times [0,\delta).
855: \end{equation}
856: Fix also a smooth cut-off function $\beta$, equal to $1$ for $|x|\leq
857: 2\delta$ and equal to $0$ for all $|x|\geq 3\delta$.
858: \begin{lem} \label{l1.6.6.5}
859: \textup{(i)} In \eqref{e11.3.5.5}, suppose that $u=0$ in $(-2\delta,2\delta)$.
860: Then $f(x,y)$ is real-analytic for $(x,y)\in B$ and has an expansion in this
861: domain of the form
862: \begin{equation}\label{e6.6.6.5}
863: f(x,y) = y^2f_2(x) + y^4f_4(x) + \cdots
864: \end{equation}
865: \textup{(ii)} In \eqref{e11.3.5.5}, suppose that $u$ has compact
866: support, and that $u$ is $C^2$ in $(-3\delta,3\delta)$. Then $f(x,y)$
867: is continuous for $(x,y)$ in $B$ and real-analytic for $(x,y)\in B_0$.
868: Moreover, if $(x,y)\in B$,
869: \begin{equation}\label{e7.6.6.5}
870: |f(x,y) - u(x)| \leq \hf |y| \int |(\beta u)''(x_1)|\,\rd x_1\mbox{ as }y\to 0.
871: \end{equation}
872: \end{lem}
873: \begin{proof}
874: For (i), note that if $(x,y)\in B$ and
875: $x_1$ is in the support of $u$, we have $|x-x_1|> \delta$ and so $k_y(x-x_1)$
876: can be expanded as a convergent series
877: \begin{equation}\label{e21.3.5.5}
878: \frac{y^2}{2((x-x_1)^2+y^2)^{3/2}}=
879: \frac{y^2}{2|x-x_1|^3}\left(1-\frac{3}{2}\frac{y^2}{(x-x_1)^2}+\cdots \right).
880: \end{equation}
881: The result follows at once from this, using term-by-term integration.
882:
883:
884: For part (ii), note first that
885: \begin{equation}\label{e85.6.6.5}
886: f(x,y) = \int k_y(x-x_1)\beta(x_1)u(x_1)\,\rd x_1 +
887: \int k_y(x-x_1)(1-\beta(x_1))u(x_1)\,\rd x_1 = I_1 + I_2.
888: \end{equation}
889: By part (i), $I_2$ is real-analytic and $O(y^2)$ if $(x,y)\in B$. On the other
890: hand, by \eqref{e4.6.6.5}
891: \begin{equation}\label{e9.6.6.5}
892: I_1(x,y) = \int \hf\sqrt{(x-x_1)^2+y^2}(\beta u)''(x_1)\,\rd x_1.
893: \end{equation}
894: Hence $I_1$ is real-analytic in $B_0$ and continous in $B$, with
895: \begin{equation}\label{e10.6.6.5}
896: I_1(x,0) = \int \hf |x-x_1|(\beta u)''(x_1)\,\rd x_1 = u(x).
897: \end{equation}
898: The estimate \eqref{e7.6.6.5} is obtained by noting
899: \begin{equation}\label{e11.6.6.5}
900: |I_1(x,y) - I_1(x,0)| \leq \hf
901: \int \bigl(\sqrt{(x-x_1)^2+y^2} - |x-x_1|\bigr)|(\beta u)''(x_1)| \,\rd x_1
902: \end{equation}
903: and using the triangle inequality,
904: \begin{equation*}
905: 0\leq \sqrt{(x-x_1)^2+y^2} - |x-x_1| \leq y.
906: \end{equation*}
907: \end{proof}
908:
909: We turn to now to the boundary behaviour of \eqref{e11.3.5.5} when $u$
910: is piecewise linear near $0$.
911:
912: \begin{lem} Suppose in the above that
913: \begin{equation}\label{e12.6.6.5}
914: u(x) = a\,\sign(x) + b|x| + c x\mbox{ for }|x| < 3\delta.
915: \end{equation}
916: Then for $(x,y)\in B_0$,
917: \begin{equation}\label{e22.3.5.5}
918: f(x,y) = a\frac{x}{(x^2+y^2)^{1/2}} + b(x^2+y^2)^{1/2} + cx + O(y^2)
919: \end{equation}
920: where $O(y^2)$ stands for a real-analytic function of $(x,y)$ which
921: goes to $0$ like $y^2$ as $y\to 0$.
922: \label{l1.5.7.5}\end{lem}
923:
924: \begin{proof} With $\beta$ as before,
925: \begin{equation*}
926: (\beta u)''(x) = 2a\delta'(x) + 2 b \delta(x) + \beta'' u
927: \end{equation*}
928: so that, using \eqref{e4.6.6.5},
929: \begin{equation*}
930: f(x,y) = 2a K_y(x) + 2b K'_y(x) + \int K_y(x-x_1)\beta''(x_1)u(x_1)\,\rd x_1
931: = f_1(x,y) + f_2(x,y) +f_3(x,y).
932: \end{equation*}
933: Since $\beta''(x_1)u(x_1)$ vanishes for $|x|\leq 2\delta$, it follows
934: as in Lemma~\ref{l1.6.6.5} that $f_3(x,y)$ is real-analytic in $B$. By
935: part (ii) of the same lemma
936: $f_3(x,0) =cx$ (for we know that
937: $f(x,0) = u(x)$, at least if $x\not=0$). Since
938: \begin{equation*}
939: f_1(x,y) = a\sqrt{x^2+y^2}, \quad f_2(x,y) = \frac{b x}{\sqrt{x^2+y^2}},
940: \end{equation*}
941: the proof is complete.
942: \end{proof}
943:
944: Using these lemmas, the proof of the following result is immediate.
945: \begin{prop} Suppose that
946: \begin{equation*}
947: u(x) = \begin{cases} m_1 x - n &\mbox{for }0<x< 3\delta;\\
948: m_2 x - n &\mbox{for }-3\delta<x< 0.
949: \end{cases}
950: \end{equation*}
951: Then if \eqref{e5.6.6.5} is satisfied, we have
952: \begin{equation*}
953: f(x,y) = \hf(m_1-m_2)\sqrt{x^2+y^2} + \hf(m_1+m_2)x - n + O(y^2).
954: \end{equation*}
955: \label{p1.6.6.5}
956: \end{prop}
957: \begin{proof}
958: Write
959: \begin{equation*}
960: u(x) = \hf(m_1-m_2)|x| + \hf(m_1+m_2)x -n
961: \end{equation*}
962: and apply Lemma~\ref{l1.5.7.5}.
963: \end{proof}
964:
965: We now give our result concerning the boundary behaviour of
966: $w(x,y)$. Let the notation be as in \S\ref{s2.5.6.5}.
967:
968: \begin{prop} \textup{(i)} Let $c$ be a point at which $f(x)$ is smooth. Then
969: $w(x,y)$ is smooth and positive in $B$.
970:
971: \noindent\textup{(ii)} Let $c$ be one of the $a_j$. Then in $B_0$ we have
972: \begin{equation*}
973: w(x,y) = \frac{1}{\sqrt{(x-c)^2+y^2}} +O(1),
974: \end{equation*}
975: where $O(1)$ stands for a smooth positive function in $B$.
976: \label{p1.5.7.5}\end{prop}
977:
978: \begin{proof} In case (i), write \eqref{e11.6.5.5} in the form
979: \begin{equation}\label{e1.27.7.5}
980: w(x,y) = \frac{1}{4}\int\left[ k_y(x-x_1) \int
981: ((x-x_2)^2+y^2)^{-3/2}D(x_1,x_2)\,\rd x_2\right]\, \rd x_1.
982: \end{equation}
983: Because $D(x_1,x_2)=0$ if $x_1$ and $x_2$ are sufficiently close to
984: $c$, the function
985: \begin{equation}\label{e2.27.7.5}
986: x\mapsto \int |x-x_2|^{-3}D(x,x_2)\,\rd x_2
987: \end{equation}
988: is smooth for $x$ near $c$, and because $k_y(x) \to \delta(x)$ as
989: $y\to 0$, we obtain
990: \begin{equation*}
991: w(x,0) = \frac{1}{4}\int |x-x_2|^{-3}D(x,x_2)\,\rd x_2
992: \mbox{ for $x$ sufficiently close to $c$.}
993: \end{equation*}
994: In particular, $w(x,0)>0$ by the non-negativity of $D(x_1,x_2)$. With
995: a little more work it can be shown that $w(x,y)$ is smooth in $B$,
996: following the proof of Lemma~\ref{l1.6.6.5}.
997:
998: In case (ii), we use a cut-off function $\beta$, equal to $1$ near $c$
999: and with small support, to split the integral as
1000: \begin{equation*}
1001: w(x,y) = w_1(x,y) + w_2(x,y),
1002: \end{equation*}
1003: where
1004: \begin{equation*}
1005: w_1(x,y) = \frac{1}{4}\int\left[ k_y(x-x_1) \int
1006: ((x-x_2)^2+y^2)^{-3/2}(1-\beta(x_2))D(x_1,x_2)\,\rd x_2\right]\,
1007: \rd x_1,
1008: \end{equation*}
1009: and
1010: \begin{equation*}
1011: w_2(x,y) = \frac{1}{4}\int\left[ k_y(x-x_1) \int
1012: ((x-x_2)^2+y^2)^{-3/2}\beta(x_2)D(x_1,x_2)\,\rd x_2\right]\,
1013: \rd x_1.
1014: \end{equation*}
1015: Now $w_1(x,0)$ is smooth and positive near $c$ by the same argument as
1016: before, and if we write
1017: \begin{equation*}
1018: w_2(x,y) = w_3(x,y) + w_4(x,y)
1019: \end{equation*}
1020: where
1021: \begin{equation*}
1022: w_3(x,y) = \frac{1}{4}\int\left[ k_y(x-x_1) (1-\beta(x_1))\int
1023: ((x-x_2)^2+y^2)^{-3/2}\beta(x_2)D(x_1,x_2)\,\rd x_2\right]\,
1024: \rd x_1,
1025: \end{equation*}
1026: and
1027: \begin{equation*}
1028: w_4(x,y) = \frac{1}{4}\int\left[ k_y(x-x_1) \beta(x_1)\int
1029: ((x-x_2)^2+y^2)^{-3/2}\beta(x_2)D(x_1,x_2)\,\rd x_2\right]\,
1030: \rd x_1,
1031: \end{equation*}
1032: then $w_3(x,0)$ is also smooth and positive near $c$ by symmetry. Thus
1033: it remains only to consider $w_4(x,y)$.
1034:
1035:
1036:
1037: % By a translation, we may set $c=0$; now pick $\delta>0$
1038: %so small so that $f(x)$ is smooth in $(-3\delta,3\delta)\setminus
1039: % 0$ and let the box $B$ and the bump-function $\beta$ be as before.
1040: %
1041: %In case (i), note that $D(x_1,x_2)$ vanishes
1042: %
1043: %We decompose
1044: %\begin{equation}\label{e6.5.5.5}
1045: %w(x,y) = \hf y^{-2}\int\int k_y(x-x_1)k_y(x-x_2)D(x_1,x_2)\,\rd x_1\rd x_2.
1046: %\end{equation}
1047: %as a sum
1048: %\begin{equation}
1049: %w = w_0 +w_1+w_2 +w_3
1050: %\end{equation}
1051: %where
1052: %\begin{equation}\label{e6.5.5.5}
1053: %w_j(x,y) =\hf y^{-2}\int\int k_y(x-x_1)k_y(x-x_2)D_j(x_1,x_2)\,\rd x_1\rd x_2.
1054: %\end{equation}
1055: %and
1056: %\begin{eqnarray*}
1057: %D_0(x_1,x_2) &=& (1-\beta(x_1))(1-\beta(x_2))D(x_1,x_2), \\
1058: %D_1(x_1,x_2) &=& \beta(x_1)(1-\beta(x_2))D(x_1,x_2),\\
1059: %D_2(x_1,x_2) &=& (1-\beta(x_1))\beta(x_2)D(x_1,x_2),\\
1060: %D_3(x_1,x_2) &=& \beta(x_1)\beta(x_2))D(x_1,x_2).
1061: %\end{eqnarray*}
1062: %We note that each $w_j\geq 0$, with equality only possible for $w_3$,
1063: %if $D_3=0$, and this is the case if and only if $(\mu,\nu)$ is
1064: %constant in $\supp(\beta)$.
1065: %Consider
1066: %\begin{equation}\label{e21.6.6.5}
1067: %v(x,x_2,y):=y^{-2}\int k_y(x-x_1)(1-\beta(x_1))D(x_1,x_2)\,\rd x_1.
1068: %\end{equation}
1069: %For each fixed $x_2$, $v(x,x_2,y)$ is smooth for $(x,y)\in B$ and
1070: %$v(x,x_2,0)$ is strictly positive. This follows from
1071: %Lemma~\ref{l1.6.6.5}, and the fact that $D(x_1,x_2)>0$ for $|x_1-x_2|$
1072: %sufficiently large. Since
1073: %\begin{eqnarray}
1074: %w_0(x,y) &=& \hf\int k_y(x-x_2)(1-\beta(x_2))v(x,x_2,y)\,\rd x_2\\
1075: %w_2(x,y) &=& \hf\int k_y(x-x_2)\beta(x_2)v(x,x_2,y)\,\rd x_2
1076: %\end{eqnarray}
1077: %it follows that
1078: %\begin{equation}
1079: %w_0 = O(y^2)
1080: %\end{equation}
1081: %in all cases.
1082: %
1083: %Now we can complete the discussion of case (i). For in that case,
1084: %$\beta(x_2)v(x,x_2,y)$ is linear in $x_2$ for $|x_2| < 3\delta$, and
1085: %so by Lemma~\ref{l1.6.6.5} $w_2$ is smooth and positive, with boundary
1086: %value
1087: %$$
1088: %w_2(x,0) = v(x,x,0) >0.
1089: %$$
1090: %Interchanging the roles of $x_1$ and $x_2$, we conclude also that
1091: %$w_1(x,y)$ is smooth with positive boundary value. Finally we note
1092: %that $w_3=0$ in $B$ because $\beta(x_1)\beta(x_2)D(x_1,x_2)=0$. Adding
1093: %up the contributions, we conclude as required that $w$ is smooth and
1094: %positive in $B$.
1095: %
1096: %Moving to case (ii), we note that $x_2 \mapsto v(x,x_2,y)$ is
1097: %piecewise linear near $0$, and so we can write
1098: %\begin{equation}
1099: %\beta(x_2)v(x,x_2,y)=
1100: %a(x,y) \sign(x_2) + b(x,y) |x_2| + O(1)\mbox{ for }|x-x_2| <3\delta
1101: %\end{equation}
1102: %where $O(1)$ is a bounded smooth function of $(x,y)$. Applying
1103: %Lemma~\ref{l1.5.7.5},
1104: %\begin{equation}
1105: %w_2(x,y)=
1106: %a(x,y) \frac{x}{\sqrt{x^2+y^2}} + b(x,y) \sqrt{x^2+y^2} + O(1)\mbox{
1107: % for }|x-c|<\delta, 0<y <\delta.
1108: %\end{equation}
1109: %By symmetry, $w_1$ will also have an analogous expansion.
1110: %
1111:
1112: In order to simplify the notation, use a translation to set $c=0$ and
1113: assume that $u$ is as in
1114: Proposition~\ref{p1.6.6.5} near $x=0$. Then
1115: \begin{equation*}
1116: D(x_1,x_2) =
1117: -\hf x_2\sign(x_1) -\hf x_1\sign(x_2) + \hf|x_1| + \hf|x_2|
1118: \mbox{ for }(x_1,x_2) \in (-\delta,\delta)\times (-\delta,\delta),
1119: \end{equation*}
1120: assuming, as we may, that
1121: \begin{equation*}
1122: n(m_1-m_2) =1.
1123: \end{equation*}
1124:
1125: Now Lemma~\ref{l1.5.7.5} can be applied, giving
1126: \begin{equation*}
1127: \int k_y(x-x_1)\beta(x_1)D(x_1,x_2)\,\rd x_1
1128: =\hf
1129: - \frac{xx_2}{\sqrt{x^2+y^2}} + \sqrt{x^2+y^2}
1130: - x_1 \sign(x_2) + |x_2| + O(y^2)
1131: \end{equation*}
1132: and then
1133: \begin{align*}
1134: \int\!\!\int k_y(x-x_1)\beta(x_1)\beta(x_2)D(x_1,x_2)\,\rd x_1\rd x_2
1135: &=
1136: - \frac{x^2}{\sqrt{x^2+y^2}} + \sqrt{x^2+y^2} + O(y^2) \\
1137: &= \frac{y^2}{\sqrt{x^2+y^2}} + O(y^2).
1138: \end{align*}
1139: Dividing this by $y^2$ and combining with our conclusions about the
1140: other $w_j$ now gives the stated result.
1141: \end{proof}
1142:
1143: \section{Completeness of $g_\alpha$}
1144: We now have a SDE metric on $M_\alpha$. Our final task is to show that
1145: this metric is complete.
1146: \label{s1.9.8.5}
1147: \begin{thm}
1148: Let
1149: $\Gamma:[0,1)\to M_\alpha$ be a smooth curve with finite length,
1150: \begin{equation}\label{e1.1.12.4}
1151: L(\gamma) = \lim_{t\to 1} d(\Gamma(0),\Gamma(t)) < \infty.
1152: \end{equation}
1153: Then there exists a point $p\in M_\alpha$ such that
1154: \begin{equation}\label{e1.6.5.5}
1155: \lim_{t\to 1} \Gamma(t) = p.
1156: \end{equation}
1157: \end{thm}
1158:
1159: \begin{proof}
1160: Let $\gamma = \pi\circ \Gamma$ be the projection of the curve to
1161: $\oD_+$. Then from the form of the metric $g_\alpha$, it is clear that this
1162: too has finite length with respect to the base metric
1163: \begin{equation}\label{3.22.12}
1164: h = \frac{w} {f^2}(\rd x^2 + \rd y^2).
1165: \end{equation}
1166:
1167: From the estimates proved below, $h$ is uniformly bounded below by a multiple
1168: of the euclidean metric on $\R^2$, so that
1169: \begin{equation*}
1170: \gamma(1) : = \lim_{t\to 1}\gamma(t)
1171: \end{equation*}
1172: exists and lies in $\oD_+\cup Z\cup\{\alpha\}$. There are now three
1173: possibilities to consider. First, suppose $\gamma(1)\in \oD_+$. Then
1174: $\pi^{-1}(\gamma(1))$ is contained in the interior of $M_\alpha$, and
1175: so \eqref{e1.6.5.5} must be satisfied. Next, suppose, if possible,
1176: that $\gamma(1)\in Z$. Since $f$ vanishes on $Z$, $h$ has a
1177: double-pole along $Z$, and it follows that $\gamma$ must have infinite
1178: length, a contradiction. The remaining possibility is that
1179: $\gamma(1) = (\alpha,0)$.
1180:
1181: From the estimates proved below, we have
1182: \begin{equation}\label{e3.6.5.5}
1183: x>\alpha, y>0\mbox{ implies that }\frac{w}{f^2}>C(x+y-\alpha)^{-2}.
1184: \end{equation}
1185: If we set $\xi =x -\alpha - y$, $\eta = x -\alpha +y$, then we get
1186: \begin{equation*}
1187: h \geq C\frac{\rd \xi^2 + \rd \eta^2}{2\eta^2}\mbox{ for }|\xi| \leq \eta
1188: \end{equation*}
1189: (that is, LHS minus RHS is positive-definite). Now our curve $\gamma$
1190: is contained in $\{|\xi| < \eta\}$ and $\lim_{t\to 1}\gamma(t) =
1191: (0,0)$. Hence $l(\gamma)= \infty$ and this contradiction completes the
1192: proof, modulo the estimates established in the next section.
1193: \end{proof}
1194:
1195:
1196:
1197: \subsection{Estimates}
1198:
1199: In order to simplify the notation in this section, we shift variables
1200: so that $\alpha$ is translated to the origin, and we aim to understand the
1201: behaviour of $f$ near $(0,0)$.
1202:
1203: First of all, we have
1204: \begin{prop}\label{p1.5.3.4}
1205: Let $u(x)$ and $f(x,y)$ be as throughout. Then there exist $\ve>0$ and $C>0$
1206: so that
1207: \begin{equation}\label{e4.6.5.5}
1208: f(x,y)\leq C\sqrt{x+y} \mbox{ if } (x,y) \in (0,\ve)\times (0,\ve).
1209: \end{equation}
1210: \end{prop}
1211: \begin{proof}
1212: The result will be established by showing that
1213: \begin{equation}\label{e1.22.8.5}
1214: f(x,\theta x) \leq C\sqrt{x}\mbox{ if }0<\theta \leq 1
1215: \end{equation}
1216: and
1217: \begin{equation}\label{e2.22.8.5}
1218: f(\theta_1 y, y) \leq C\sqrt{y}\mbox{ if }0<\theta_1 \leq 1.
1219: \end{equation}
1220: Since $u$ is odd,
1221: \begin{equation}\label{e1.3.6.5}\begin{split}
1222: f(x,y) &=\int_{0}^\infty \{k_y(x-x_1) - k_y(x+x_1)\}u(x_1)\,\rd x_1\\
1223: &\leq \Omega\int_{0}^\infty \{k_y(x-x_1) - k_y(x+x_1)\}\sqrt{x_1}\,\rd x_1\\
1224: &\leq \Omega\int_{0}^\infty k_y(x-x_1)\sqrt{x_1}\,\rd x_1,
1225: \end{split}
1226: \end{equation}
1227: where we have used the upper bound of Theorem~\ref{t1.3.5.5} and the
1228: positivity of $k_y(x)$.
1229:
1230: To prove \eqref{e1.22.8.5}, introduce a natural rescaling of variables
1231: in \eqref{e1.3.6.5},
1232: \begin{equation}\label{e7.6.5.5}
1233: \xi = x_1/x, \theta = y/x
1234: \end{equation}
1235: so that \eqref{e1.3.6.5} becomes
1236: \begin{equation} \label{1.27.5}
1237: f(x,x\theta) \leq \Omega I = \Omega\sqrt{x}\int_0^\infty
1238: k_\theta(\xi-1)\sqrt{\xi}\,\rd \xi.
1239: \end{equation}
1240: Note that this yields \eqref{e1.22.8.5} for each fixed positive
1241: $\theta$, but the uniformity as $\theta\to 0$ needs a little further work.
1242: For this we split $I$, as in \S\ref{s1.5.7.5}, using a bump-function
1243: $\beta$ identically
1244: equal to $1$ in a neighbourhood of $\xi=1$. We have
1245: \begin{equation*}
1246: I = I_1 + I_2
1247: \end{equation*}
1248: where
1249: \begin{equation*}
1250: I_1 =\int_0^\infty k_\theta(\xi-1)(1-\beta(\xi))\sqrt{\xi}\,\rd \xi,\quad
1251: I_2 =\int_0^\infty k_\theta(\xi-1)\beta(\xi)\sqrt{\xi}\,\rd \xi.
1252: \end{equation*}
1253: Then since $\lim_{\theta\to 0}k_\theta(\xi-1) =\delta(\xi-1)$, we have
1254: \begin{equation*}
1255: I_2\to Cx^{1/2}\beta(1)\sqrt{1}= C\sqrt{x} \mbox{ as }\theta \to 0.
1256: \end{equation*}
1257: By continuity, $I_2$ is uniformly bounded by $C\sqrt{x}$ for all
1258: $0\leq \theta \leq 1$. The estimate \eqref{e1.22.8.5} now follows by
1259: noting that $I_1\geq 0$ and $I_1\to 0$ as $\theta\to
1260: 0$. In fact, it is easily
1261: shown that $I_1 \leq C' \theta^2\sqrt{x}$ for
1262: $0\leq\theta \leq 1$.
1263:
1264: The complementary estimate \eqref{e2.22.8.5} is somewhat simpler:
1265: return to \eqref{e1.3.6.5} and make the change of
1266: variables $x = y\theta_1$, $x_1 = y\xi_1$, so
1267: \begin{equation}\label{2.27.5}
1268: f(y\theta_1,y ) \leq \Omega\sqrt{y}
1269: \int_0^\infty k_1(\xi_1-\theta_1)\sqrt{\xi_1}\,\rd \xi_1.
1270: \end{equation}
1271: Clearly the integral is uniformly bounded for $\theta_1\in [0,1]$
1272: and this completes the proof.
1273: \end{proof}
1274:
1275: Next we need a lower bound on $w(x,y)$. This is given by
1276: \begin{prop}\label{p2.5.3.4}
1277: Suppose that in the continued fraction expansion of $\alpha$, $3\leq e_j \leq
1278: N$, for some $N$. Then there exist $\ve>0$ and $C>0$ so that
1279: \begin{equation}\label{e3.5.3.4}
1280: w(x,y)\geq C(x+y)^{-1}\mbox{ if } (x,y) \in (0,\ve)\times (0,\ve).
1281: \end{equation}
1282: \end{prop}
1283: \begin{proof} The argument is closely analogous to the proof of the previous
1284: proposition. In particular, \eqref{e3.5.3.4} will be established by proving the
1285: separate inequalities
1286: \begin{equation}
1287: w(x,\theta x) \geq Cx^{-1}\mbox{ if }0<\theta \leq 1
1288: \end{equation}
1289: and
1290: \begin{equation}\label{e5.9.8.5}
1291: w(\theta_1 y, y) \geq Cy^{-1}\mbox{ if }0<\theta_1 \leq 1.
1292: \end{equation}
1293:
1294: We use~\eqref{e11.6.5.5}. Because $D\geq 0$ and
1295: $k_y(x)>0$, it is enough to prove the lower bound for
1296: \begin{equation}\label{e3.3.6.5}
1297: I(x,y) := \hf y^{-2}\int_{x_1=0}^\infty \int_{x_2=0}^{x_1}k_y(x-x_1)k_y(x-x_2)
1298: D(x_1,x_2)\,\rd x_1\rd x_2
1299: \end{equation}
1300: where the integral has been restricted to the intersection of the
1301: positive quadrant with the region $x_2\leq x_1$. In this integral,
1302: make the change of variables
1303: \begin{equation*}
1304: x_1 = x\xi, x_2 = x\xi\eta, y = \theta x.
1305: \end{equation*}
1306: Then
1307: \begin{equation*}
1308: I(x,\theta x) =
1309: \frac{1}{2x^2\theta^2}\int_{\xi=0}^\infty\int_{\eta=0}^1
1310: k_\theta(\xi-1)k_\theta(\xi\eta-1)D(x\xi,x\xi\eta)\xi\rd\xi\rd \eta
1311: \end{equation*}
1312: \begin{equation}\label{e5.3.6.5}
1313: \hspace{1in} \geq
1314: \frac{1}{2x\theta^2}\int_{\xi=0}^{a/x}\int_{\eta=0}^b
1315: k_\theta(\xi-1)k_\theta(\xi\eta-1)\xi^2(1-\eta)\rd\xi\rd \eta,
1316: \end{equation}
1317: where $a = 1 -\alpha$, $b = 1/16N^2$ as in Lemma~\ref{l1.11.8.5}. For each
1318: $\theta$ this gives the required $O(1/x)$ lower bound. To see this is uniform
1319: as $\theta \to 0$, rewrite \eqref{e5.3.6.5} as follows and take the limit:
1320: \begin{equation}\label{e33.3.6.5}
1321: \frac{1}{4x} \int_0^{a/x} k_\theta(\xi-1)\left[\xi^2
1322: \int_{\eta=0}^b(1-\eta)((1-\xi\eta)^2 +\theta^2)^{-3/2}\rd\eta\right] \rd \xi
1323: \longrightarrow
1324: \frac{1}{4x}
1325: \int_0^b(1-\eta)^{-2}\rd\eta\mbox{ as }\theta \to 0.
1326: \end{equation}
1327: By continuity, the required uniform lower bound follows.
1328:
1329: In order to obtain the other bound \eqref{e5.9.8.5}, return to \eqref{e3.3.6.5}
1330: and make the substitutions
1331: \begin{equation*}
1332: x = \theta_1 y, x_1 = y\xi, x_2 = y\xi\eta
1333: \end{equation*}
1334: to give
1335: \begin{equation}\label{e5.11.8.5}
1336: w(y\theta_1, y) \geq
1337: \frac{1}{2y}\int_{\xi=0}^{a/y}\int_0^b
1338: k_1(\xi-\theta_1)k_1(\xi\eta-\theta_1)\xi^2(1-\eta)\,\rd\xi\rd\eta.
1339: \end{equation}
1340: This gives \eqref{e5.9.8.5} at once.
1341: \end{proof}
1342:
1343: \subsection{Remark} As in \cite[\S5]{CaSi:emcs} the distribution
1344: $u$ can be perturbed to the left of $\alpha$ to yield an infinite-dimensional
1345: family of SDE metrics on $M_\alpha$. These perturbations need to preserve the
1346: monotonicity and convexity properties enjoyed by $u$, as well as the boundary
1347: condition $u=-1$ for $x\ll 0$. Provided these perturbations are sufficiently
1348: small and supported away from $\alpha$, the resulting metrics will be complete;
1349: the details are left to the interested reader.
1350:
1351:
1352:
1353: % References
1354: %
1355: %
1356: % Bibliography format
1357: %
1358: \newcommand{\bauth}[1]{\mbox{#1}} \newcommand{\bart}[1]{\textit{#1}}
1359: \newcommand{\bjourn}[4]{#1\ifx{}{#2}\else{ \textbf{#2}}\fi{ (#4)}}
1360: \newcommand{\bbook}[1]{\textsl{#1}}
1361: \newcommand{\bseries}[2]{#1\ifx{}{#2}\else{ \textbf{#2}}\fi}
1362: \newcommand{\bpp}[1]{#1} \newcommand{\bdate}[1]{ (#1)} \def\band/{and}
1363: %
1364: \newif\ifbibtex%\bibtextrue% Comment out \bibtextrue to remove BiBTeX.
1365: %
1366: \ifbibtex
1367: %
1368:
1369:
1370:
1371:
1372: \bibliographystyle{genbib}
1373: \bibliography{papers}
1374: %
1375:
1376: %
1377:
1378: \else
1379:
1380: \begin{thebibliography}{10}
1381:
1382: %\bibitem{And:emc}
1383: %\bauth{M.~T. Anderson}, \bart{Einstein metrics with prescribed conformal
1384: % infinity on {$4$}-manifolds}, Preprint, SUNY Stonybrook\bdate{2001},
1385: % math.DG/0105243.
1386:
1387: \bibitem{AKL:rfitt}
1388: \bauth{M.~T. Anderson, P. B. Kronheimer, and C. R. LeBrun}
1389: \bart{Complete {R}icci-flat {K}\"ahler manifolds of infinite topological
1390: type} \bjourn{Comm. Math. Phys.}{125}{}{1989} \bpp{637--642}.
1391:
1392: %\bibitem{BPV:ccs}
1393: %\bauth{W.~Barth, C.~Peters \band/ A.~Van~de Ven}, \bbook{Compact Complex
1394: % Surfaces}, Springer-Verlag, Berlin Heidelberg New York Tokyo\bdate{1984}.
1395:
1396: %\bibitem{Bel:ncr}
1397: %\bauth{F.~A. Belgun}, \bart{Normal {CR} structures on compact $3$-manifolds},
1398: % \bjourn{Math. Z.}{238}{}{2001} \bpp{441--460}.
1399:
1400: %\bibitem{Biq:mab}
1401: %\bauth{O.~Biquard}, \bart{M{\_1e}triques autoduales sur la boule}, Preprint,
1402: % IRMA Strasbourg\bdate{2000}, math.DG/0010188.
1403:
1404: %\bibitem{Biq:mes}
1405: %\bauth{O.~Biquard}, \bart{M\_1etriques d_1{E}instein asymptotiquement
1406: % sym\_1etriques}, \bjourn{Ast\_1erisque}{265}{}{2000}.
1407:
1408: %\bibitem{BGMR:3s7}
1409: %\bauth{C.~P. Boyer, K.~Galicki, B.~M. Mann \band/ E.~G. Rees}, \bart{Compact
1410: % $3$-{S}asakian $7$-manifolds with arbitrary second {B}etti number},
1411: % \bjourn{Invent. Math.}{131}{}{1998} \bpp{321--344}.
1412:
1413: \bibitem{CaPe:emt}
1414: \bauth{D.~M.~J. Calderbank \band/ H.~Pedersen}, \bart{Selfdual {E}instein
1415: metrics with torus symmetry}, \bjourn{J.~Diff. Geom.}{60}{}{2002}
1416: \bpp{485--521}, math.DG/0105263.
1417:
1418: \bibitem{CaSi:emcs}
1419: \bauth{D.~M.~J. Calderbank \band/ M.~A.~Singer}, \bart{Einstein
1420: metrics and complex singularities},
1421: \bjourn{Invent. Math.}{156}{}{2004} \bpp{405--443},
1422: math.DG/0206229.
1423:
1424: \bibitem{CaSi:temco}
1425: \bauth{D.~M.~J. Calderbank \band/ M.~A.~Singer}, \bart{Toric selfdual
1426: Einstein metrics on compact orbifolds},
1427: \bjourn{Duke Math. J.}{}{}{to appear},math.DG/0405020.
1428:
1429: %\bibitem{FeGr:ci}
1430: %\bauth{C.~Fefferman \band/ C.~R. Graham}, \bart{Conformal invariants}, in
1431: % \bbook{The Mathematical Heritage of {\_1E}lie Cartan} (Lyon, 1984),
1432: % Ast{\_1e}risque\bdate{1985}, pp.~95--116.
1433:
1434: \bibitem{Gal:mcm}
1435: \bauth{K.~Galicki}, \bart{Multi-centre metrics with negative cosmological
1436: constant}, \bjourn{Class. Quantum Grav.}{8}{}{1991} \bpp{1529--1543}.
1437:
1438: \bibitem{GiHa:gmi}
1439: \bauth{G.~W. Gibbons \band/ S.~W. Hawking}, \bart{Gravitational
1440: multi-instantons}, \bjourn{Phys. Lett.}{B 78}{}{1978} \bpp{430--432}.
1441:
1442: \bibitem{GrLe:emc}
1443: \bauth{C.~R. Graham \band/ J.~M. Lee}, \bart{Einstein metrics with prescribed
1444: conformal infinity on the ball}, \bjourn{Adv. Math.}{87}{}{1991}
1445: \bpp{186--225}.
1446:
1447: %\bibitem{Hit:pg}
1448: %\bauth{N.~J. Hitchin}, \bart{Polygons and gravitons},
1449: %\bjourn{Math. Proc. Camb.
1450: % Phil. Soc.}{85}{}{1979} \bpp{465--476}.
1451:
1452: %\bibitem{Hit:tem}
1453: %\bauth{N.~J. Hitchin}, \bart{Twistor spaces, {E}instein metrics and
1454: % isomonodromic deformations}, \bjourn{J.~Diff. Geom.}{42}{}{1995}
1455: % \bpp{30--112}.
1456:
1457: %\bibitem{Joy:hqqq}
1458: %\bauth{D.~D. Joyce}, \bart{The hypercomplex quotient and the quaternionic
1459: % quotient}, \bjourn{Math. Ann.}{290}{}{1991} \bpp{323--340}.
1460:
1461: %\bibitem{Joy:qcs}
1462: %\bauth{D.~D. Joyce}, \bart{Quotient constructions for compact self-dual
1463: % {$4$}-manifolds}, Merton College, Oxford\bdate{1991}.
1464:
1465: %\bibitem{Joy:esd}
1466: %\bauth{D.~D. Joyce}, \bart{Explicit construction of self-dual
1467: %{$4$}-manifolds},
1468: % \bjourn{Duke Math. J.}{77}{}{1995} \bpp{519--552}.
1469:
1470: %\bibitem{KoSi:gt}
1471: %\bauth{A.~G. Kovalev \band/ M.~A. Singer}, \bart{Gluing theorems for complete
1472: % anti-self-dual spaces}, \bjourn{Geom. Funct. Anal.}{11}{}{2001}
1473: % \bpp{1229--1281}.
1474:
1475: \bibitem{Kro:ale}
1476: \bauth{P.~B. Kronheimer}, \bart{The construction of {ALE} spaces as
1477: hyper-{K\"a}hler quotients}, \bjourn{J.~Diff. Geom.}{29}{}{1989}
1478: \bpp{665--683}.
1479:
1480: \bibitem{OrRa:at4}
1481: \bauth{P.~Orlik \band/ F.~Raymond}, \bart{Actions of the torus on $4$-manifolds
1482: \textup{I}}, \bjourn{Trans. Amer. Math. Soc.}{152}{}{1970} \bpp{531--559}.
1483:
1484:
1485: \end{thebibliography}
1486: \fi
1487: \end{document}
1488: \bibitem{LeBr:hcc}
1489: \bauth{C.~R. LeBrun}, \bart{{$\mathcal H$}-space with a cosmological constant},
1490: \bjourn{Proc. Roy. Soc. London}{A 380}{}{1982} \bpp{171--185}.
1491:
1492: \bibitem{LeBr:pac}
1493: \bauth{C.~R. LeBrun}, \bart{Counterexamples to the generalized positive action
1494: conjecture}, \bjourn{Comm. Math. Phys.}{118}{}{1988} \bpp{591--596}.
1495:
1496: \bibitem{LeBr:cp2}
1497: \bauth{C.~R. LeBrun}, \bart{Explicit self-dual metrics on
1498: {$\CP2\connect\cdots\connect\CP2$}}, \bjourn{J.~Diff. Geom.}{34}{}{1991}
1499: \bpp{223--253}.
1500:
1501: \bibitem{LeBr:cqk}
1502: \bauth{C.~R. LeBrun}, \bart{On complete quaternionic-{K\"a}hler manifolds},
1503: \bjourn{Duke Math. J.}{63}{}{1991} \bpp{723--743}.
1504:
1505: \bibitem{RM:pc}
1506: \bauth{R.~R. Mazzeo}, \bart{Private communication.}
1507:
1508:
1509: \bibitem{Ped:emm}
1510: \bauth{H.~Pedersen}, \bart{{E}instein metrics, spinning top motions and
1511: monopoles}, \bjourn{Math. Ann.}{274}{}{1986} \bpp{35--39}.
1512:
1513: \bibitem{Rol:reh}
1514: \bauth{Y.~Rollin}, \bart{Rigidit\_1e d_1{E}instein du plan hyperbolique
1515: complexe}, \bjourn{C. R. Math. Acad. Sci. Paris}{334}{}{2002} \bpp{671--676}.
1516:
1517: \bibitem{Tod:p6}
1518: \bauth{K.~P. Tod}, \bart{Self-dual {E}instein metrics from the {P}ainlev\_1e
1519: {VI} equation}, \bjourn{Phys. Lett.}{A 190}{}{1994} \bpp{221--224}.
1520:
1521: \end{thebibliography}
1522:
1523: \fi
1524:
1525:
1526: \end{thebibliography}
1527: \end{document}
1528:
1529:
1530:
1531:
1532:
1533:
1534:
1535: We shall also need the following technical result about the function
1536: $D(x_1,x_2)$, defined as
1537: \begin{equation}
1538: D(x_1,x_2) := (m_j n_k - m_k n_j)(x_1-x_2)\mbox{ for }(x_1,x_2) \in
1539: (a_j, a_{j-1}) \times (a_k, a_{k-1}).
1540: \end{equation}
1541: (See \S\ref{} below for the significance of this function.)
1542:
1543: \begin{lem}\label{l4.12.3.4} There exists a constant $\lambda\in(0,1)$ so that
1544: \begin{equation*}
1545: (x_1,x_2) \in S:= \{(x_1,x_2): \alpha < x_1, x_2 < 1, (x_2-\alpha) \leq
1546: \lambda(x_1-\alpha)\}\Rightarrow (x_1,x_2) \geq x_1-x_2.
1547: \end{equation*}
1548: \end{lem}
1549: \begin{proof} If $\alpha<x_1< x_2 < 1$, then for some $k\leq j$ ,
1550: \begin{equation}\label{e10.12.3.4}
1551: a_{j} \leq x_1 < a_{j-1}, a_k < x_2 < a_{k-1},
1552: \end{equation}
1553: and then
1554: \begin{equation}
1555: D(x_1,x_2) = (m_j n_k - m_k n_j)(x_1-x_2).
1556: \end{equation}
1557: The idea is to show that we can choose $\lambda$ so that if
1558: $(x_1,x_2)\in S$, then $j$
1559:
1560: \begin{equation}\label{e11.12.3.4}
1561: a_{k+1} \leq x_2 < a_k,
1562: \end{equation}
1563: then $D(x_1,x_2) \geq x_1-x_2$, for $m_jn_{k+1} - m_{k+1}n_j>0$
1564: for such $j$ and $k$. It is therefore enough to show that if
1565: \eqref{e10.12.3.4} holds and $(x_1,x_2) \in S$, then \eqref{e11.12.3.4}
1566: also holds. From \eqref{13.19.12}, there is a constant so that
1567: \begin{equation}\label{e12.12.3.4}
1568: x_2-\alpha \leq c(x_1-\alpha) \leq
1569: c(a_{j-1}-\alpha ) \leq \frac{c}{N^2 m_j^2}.
1570: \end{equation}
1571: Using the bound $e_j \leq N$, we find $N m_j > m_{j+2}$. Inserting
1572: this into \eqref{e12.12.3.4} and using \eqref{13.19.12} again,
1573: \begin{equation}
1574: x_2-\alpha < \frac{c}{m_{j+2}^2} < a_j - \alpha
1575: \end{equation}
1576: proving \eqref{e11.12.3.4}, for some $k\geq j$ as required.
1577: \end{proof}
1578:
1579:
1580:
1581:
1582:
1583: First, note that $w>0$ in $\cH^2$. By the integral formula, we can establish
1584: by showing that $D(x_1,x_2)\geq 0$ and is not identically zero. This follows
1585: from the integral formula, for if $x_1> x_2>\alpha$, with say
1586: $$
1587: a_k \leq x_2 \leq a_{k-1}\leq a_j \leq x_1 \leq a_{j-1}
1588: $$
1589: then
1590: $$
1591: (\mu(x_2),\nu(x_2)) = (m_k,n_k),\quad (\mu(x_1),\nu(x_1)) = (m_j,n_j)
1592: $$
1593: so that
1594: $$
1595: D(x_1,x_2) = (m_jn_k -m_k n_j)(x_1-x_2)=m_jm_k\left(\frac{n_k}{m_k}
1596: -\frac{n_j}{m_j}\right)(x_1-x_2) \geq 0
1597: $$
1598: with strict inequality if $x_2$ and $x_1$ are in distinct intervals
1599: (i.e. $j<k$). Since $D(x_1,x_2) = D(x_2,x_1)$ this establishes the positivity
1600: in the positive quadrant $x_1,x_2>\alpha$.
1601:
1602: Now from the oddness of $u$,
1603: $$
1604: (\mu(2\alpha -x), \nu(2\alpha -x)) = (\mu(x),\nu(x) - 2\alpha\mu(x))
1605: $$
1606: so that
1607: $$
1608: D(2\alpha-x_1,2\alpha -x_2) = D(x_1,x_2),
1609: $$
1610: and
1611: $$
1612: D(2\alpha-x_1,x_2) = (x_1+x_2-2\alpha)(2\alpha \mu(x_1)\mu(x_2) -
1613: \mu(x_1)\nu(x_2)-\mu(x_2)\nu(x_1))
1614: $$
1615: By this last, if $x_1,x_2>\alpha$, $D(2\alpha-x_1,x_2)>0$, for
1616: $$
1617: \nu(x)/\mu(x) <\alpha\mbox{ if }x>\alpha.
1618: $$
1619: By the symmetry of $D$ it now follows that $D\geq 0$ and is strictly positive
1620: sufficiently far from the diagonal.
1621:
1622: Next, note that the skew symmetry of $u$ carries over to $f$,
1623: $$
1624: f(\alpha+x,y) = - f(\alpha -x,y)
1625: $$
1626: so that $f=0$ on $Z$. But $f>0$ on $D_+$ by the maximum principle. It follows
1627: that all the quantities needed to define $g_\alpha$ are smooth and positive in
1628: $D_+$, which proves that $g_\alpha$ is defined on $D_+\times T^2$.
1629:
1630: Next we turn to the analysis of $g_\alpha$ near $\del \cH^2_+$. Obviously, for
1631: this we need to understand the asymptotic behaviour of $f$ and $w$ as $y\to 0$.
1632: So pick $x_0>\alpha$. We distinguish two cases according to whether $x_0$ is or
1633: is not one of the $a_j$. The easier case corresponds to $a_j < x_0 < a_{j-1}$
1634: for some $j$. Then by Theorem~\ref{}, we have
1635: \begin{equation}
1636: f(x,y) = m_jx-n_j +y^2f_1(x) +\cdots,\quad w(x,y) = w_0(x) + w_1(x)y^2 +\cdots
1637: \end{equation}
1638: for $|x-x_0|$ and $y\geq 0$ sufficiently small, where $w_0(x)>0$, and the
1639: expansions involve only even powers of $y$. Write $(m_j,n_j) = (m,n)$, to
1640: simplify the notation. Then computing with these expansions,
1641: \begin{align*}
1642: v_1 &= y(2f_1,2xf_1 - m) + O(y^3)\\
1643: v_2 &= (m,n) + O(y^2)\\
1644: \ve(v_1,v_2) &= y(2uf_1 -(u')^2).
1645: \end{align*}
1646: Thus, for small $y$,
1647: $$
1648: g_\alpha \simeq
1649: \frac{w_0}{u^2}\left(
1650: \rd x^2 +\rd y^2 + \frac{
1651: y^2(2f_1\rd z_2 + (u'-2xf_1)\rd z_1)^2
1652: + (m\rd z_2 -n\rd z_1)^2}{(2uf_1 - (u')^2)}\right)
1653: $$
1654: In particular
1655: $$
1656: g_\alpha(m\del_{z_1}+n\del_{z_2},
1657: m\del_{z_1}+n\del_{z_2}) \simeq y^2.
1658: $$
1659: If we introduce new coordinates
1660: \begin{align*}
1661: \rd \theta &= m\rd z_2 - n\rd z_1 \\
1662: \rd \psi &= m_1\rd z_2 - n_1\rd z_1,
1663: \end{align*}
1664: where $|mn_1-m_1n|=1$, we have
1665: $$
1666: g_\alpha = \frac{w_0}{u^2}
1667: \left( \rd x^2 + a(x)\rd\theta^2 + \rd y^2
1668: + y^2\rd \psi^2+\cdots)\right),\quad (a(x)>0)
1669: $$
1670: which shows that $g_\alpha$ does extend as a smooth metric to $\pi^{-1}(U)$,
1671: where $U$ is a small neighbourhood of the boundary-point $x_0$.
1672:
1673: To complete the proof, we must consider the behaviour of $g_\alpha$ at one of
1674: the ``corners'' $a_j$. By a suitable change of variables, we may assume that
1675: $x_0 = a_0 = 1$, so that $u(x) =x$ for $1-\delta<x \leq 1$, $u(x) =1$ for
1676: $1\leq x < 1+\delta$.
1677:
1678: By Theorem~\ref{}, we have expansions
1679: \begin{align*}
1680: f(x,y) &= 1 + \hf(x-1) - \hf((x-1)^2+y^2)^{1/2} + f_1(x)y^2 +\cdots\\
1681: w(x,y) &= \hf((x-1)^2+y^2)^{-1/2} + w_1(x) + \cdots \\
1682: \end{align*}
1683: where the omitted terms contain only even powers of $y$. Using the change of
1684: variable
1685: $$
1686: x-1 + iy = (r_2 + ir_1)^2
1687: $$
1688: we compute, to leading order,
1689: \begin{align*}
1690: v_1 &= -\frac{r_1r_2}{r_1^2+r_2^2}(1,1) + \cdots,\\
1691: v_2 &= \frac{1}{r_1^2+r_2^2}(r_1^2,-r_2^2) + \cdots\\
1692: w &=\frac{1}{2}\frac{1}{r_1^2+r_2^2}+\cdots\\
1693: \rd x^2 + \rd y^2 &=4(r_1^2+r_2^2)(\rd r_1^2 +\rd r_2^2).
1694: \end{align*}
1695: Substituting these into the formula for $g_\alpha$ we find
1696: \begin{equation}
1697: g_\alpha = (\rd r_1^2 +r_1^2 \rd z_1^2) +
1698: (\rd r_2^2 +r_1^2 \rd z_2^2) +\cdots
1699: \end{equation}
1700: exactly as required for smooth extension to $M_\alpha$ in a neighbourhood of
1701: the corner.
1702: \end{proof}
1703:
1704: