math0305399/hda.tex
1: \documentclass{pspum-l}
2: 
3: %% OWN MACROS BEGINS HERE
4: 
5: %% Own package inclusions
6: \usepackage[dvips]{graphicx}
7: \usepackage[latin1]{inputenc}
8: \usepackage[T1]{fontenc}
9: \usepackage{amssymb}
10: 
11: %% Own macro definitions - Simon
12: \newcommand{\abs}[1]{\ensuremath{\left\vert #1 \right\vert}}
13: \newcommand{\hdim}{\text{\textup{dim}}_{\text{\textup{H}}}}
14: 
15: %% A hack to make spacing before subsubsections
16: \makeatletter
17: \renewcommand{\subsubsection}{\@startsection
18:   {subsubsection}%
19:   {3}%
20:   {0pt}%
21:   {-.5\baselineskip}%
22:   {-\fontdimen2\font
23:     plus -\fontdimen3\font}%
24:   {\normalfont\itshape}%
25: }
26: \makeatother
27: 
28: %% Own macro definitions - Maurice
29: 
30: \DeclareFontFamily{OMS}{rsfs}{\skewchar\font'177}
31: \DeclareFontShape{OMS}{rsfs}{m}{n}{%
32:   <5> rsfs5
33:   <6> <7> rsfs7
34:   <8> <9> <10> rsfs10
35:   <10.95> <12> <14.4> <17.28> <20.74> <24.88> rsfs10
36: }{}
37: \DeclareSymbolFont{rsfscript}{OMS}{rsfs}{m}{n}
38: \DeclareSymbolFontAlphabet{\mathrsfs}{rsfscript}
39: \newcommand{\scr}[1]{\mathrsfs{#1}}  
40: \DeclareMathOperator{\di}{diam}
41: \DeclareMathOperator{\SL}{SL}
42: \DeclareMathOperator{\codim}{codim}
43: 
44: %% OWN MACRO SECTION ENDS HERE
45: 
46: % If your version of amsproc.cls is version 1.2 (before December 1999),
47: % uncomment the following definition.
48: %\renewcommand{\subjclassname}{%
49: %  \textup{2000} Mathematics Subject Classification}
50: 
51: % Update the information and uncomment if AMS is not the copyright holder.
52: \copyrightinfo{2003}{American Mathematical Society}
53: 
54: \newtheorem{theorem}{Theorem}[section]
55: \newtheorem{lemma}[theorem]{Lemma}
56: 
57: %% Own environments
58: \newtheorem{cor}[theorem]{Corollary}
59: 
60: %% End of own environments
61: 
62: \theoremstyle{definition}
63: \newtheorem{definition}[theorem]{Definition}
64: \newtheorem{example}[theorem]{Example}
65: \newtheorem{xca}[theorem]{Exercise}
66: 
67: \theoremstyle{remark}
68: \newtheorem{remark}[theorem]{Remark}
69: 
70: \numberwithin{equation}{section}
71: 
72: \begin{document}
73: 
74: \title{Hausdorff Dimension and Diophantine Approximation}
75: 
76: % Remove or comment out any unused author tags.
77: % author one information
78: \author{M. Maurice Dodson}
79: \address{University of York\\
80:   Heslington \\
81:   York \\
82:   YO10 5DD \\
83:   UK
84: }
85: %\curraddr{}
86: \email{mmd1@york.ac.uk}
87: %\thanks{}
88: 
89: % author two information
90: \author{Simon Kristensen}
91: \address{University of York\\
92:   Heslington \\
93:   York \\
94:   YO10 5DD \\
95:   UK
96: }
97: %\curraddr{}
98: \email{sk17@york.ac.uk}
99: \thanks{Research funded by EPSRC grant no. GR/N02832/01 with
100:   additional support from INTAS grant no. 001--429.}
101: 
102: % Use this \subjclass if you are using amsproc version 2.0 (December 1999).
103: \subjclass[2000]{Primary 11J83; Secondary 28A78, 37F50, 28A80}
104: % Use this one if you are using an older version of amsproc.
105: %\subjclass{}
106: %\date{}
107: 
108: \keywords{Diophantine approximation, exceptional sets, Hausdorff
109:   dimension, fractals, small divisors, dynamical systems, complex
110:   dynamics} 
111: 
112: %\begin{abstract}
113: %\end{abstract}
114: 
115: \maketitle
116: 
117: 
118: \section{Introduction }
119: 
120: Dimension provides an indication of the size and complexity of a set
121: and various kinds, such as box counting, packing and Hausdorff
122: dimensions, play an important role in the study of
123: fractals~\cite{FalcFG}.  For example, the Hausdorff dimension of the
124: Cantor middle third set is $\log 2/\log 3$ (proved by Hausdorff in his
125: seminal paper~\cite{Hausdorff19}), that of the Koch snowflake curve is
126: $\log 4/\log 3$ and it has recently been shown that the boundary of
127: the Mandelbrot set, a very complicated set of Lebesgue measure 0 or
128: \emph{null} set in the complex plane with topological dimension 1, has
129: Hausdorff dimension 2~\cite{shishikura1998}.  On the other hand,
130: Diophantine approximation is a quantitative analysis of rational
131: approximation and so, at least at first sight, is less geometrical.
132: The purpose of this article is to show that Hausdorff dimension plays
133: an important part in this theory too.
134: 
135: In order to keep the article accessible, the emphasis is on
136: approximation of real numbers by rationals and the less well known
137: topic of approximation of complex numbers by ratios of Gaussian
138: integers. The more general theory, which recently has seen some
139: spectacular advances, will be referred to and some applications
140: sketched. The article is organised as follows.  We begin with a brief
141: treatment of Hausdorff measure and Hausdorff dimension. We then
142: explain some of the principal results in Diophantine approximation and
143: the Hausdorff dimension of related sets, originating in the pioneering
144: work of Vojt{\v{e}}ch Jarník~\cite{NSJobit}.  We conclude with some
145: applications of these results to the metrical structure of exceptional
146: sets associated with some famous problems. It is not intended that all
147: the recent developments be covered but they can be found in the
148: references cited.
149: 
150: \section{Hausdorff measure and dimension}
151: 
152: Felix Hausdorff introduced the notion of Hausdorff dimension in a
153: remarkable and influential paper~\cite{Hausdorff19} that extended
154: Carath\'eodory's approach to Lebesgue measure~\cite{Carat14} in a
155: simple but far-reaching way.  (Kahane's Foreword to the book
156: \emph{Fractals} ~\cite{Cherbit91} includes a short and moving
157: biography of Hausdorff.)  Dimension had been taken to be a
158: non-negative integer but by a simple observation, which Hausdorff
159: described modestly as `a small contribution', he modified
160: Cara\-th\'eodory's definition of measure to obtain a measure
161: associated with a dimension that could be any non-negative real
162: number.  We shall assume a knowledge of Lebesgue measure and as usual,
163: we shall often say \emph{almost no} to indicate a null set -- thus
164: almost no numbers are rational -- and we shall say \emph{almost all}
165: to indicate a set whose complement is null, so that almost all numbers
166: are irrational.
167: 
168: For familiar sets such as the interval, circle and the plane, the
169: Hausdorff dimension\ (defined below) coincides with the usual notion
170: of dimension and is respectively 1,\,1 and 2.  However, a significant
171: difference is that \emph{any\/} set in Euclidean space has a Hausdorff
172: dimension (a non-measurable set in $\mathbb{R}^n$ has full Hausdorff
173: dimension $n$).  In particular, null sets, such as Cantor's middle
174: third set or the set of badly approximable numbers
175: (see~\S\ref{sec:toa}), have a Hausdorff dimension and this gives a way
176: of discriminating between them.  It is also natural to study the
177: Hausdorff dimension of \emph{exceptional sets} which are sets
178: associated with the invalidity of some result, making it desirable
179: that they be null.  A brief and more or less self-contained account of
180: Hausdorff measure and dimension is now given (more detailed
181: expositions can be found
182: in~\cite{MDAM,FalcGFSshort,FalcFG,Fed,MattilaGS,Rogers}).
183: 
184: 
185: \subsection{Hausdorff measure} 
186: 
187: Carathéodory's approach to the measure of a set $E$ in $\mathbb{R}^n$
188: was based on `approximating' $E$ by countable covers consisting of
189: small `simple' sets $U$ in $ \mathbb{R}^n$.  Hausdorff's idea was to
190: introduce for a given cover, $\scr{C}$ say, of $E$ the sum (sometimes
191: termed the \emph{$s$-length} of the cover $\scr{C}$)
192: \begin{equation*}
193:   \ell^s({\scr{C}}):=
194:   \sum_{U\in{\scr{C}}} (\di U)^{s},
195: \end{equation*}
196: where $\di U=\sup\{\vert \mathbf{x}-\mathbf{y}\vert_2\colon
197: \mathbf{x},\mathbf{y} \in U\}$ is the diameter of $U$ ($\vert
198: \mathbf{x}-\mathbf{y}\vert_2$ is the usual Euclidean distance between
199: $\mathbf{x}$ and $\mathbf{y}$) and where $s$ \emph{is a non-negative
200:   real number that is not necessarily an integer}.  Hausdorff
201: considered a monotonic function $l$ which allows more discrimination
202: but for simplicity we shall stick to the more familiar widely used
203: special case $l(\di U)=(\di U)^s$, associated with what is now usually
204: called the Hausdorff dimension but also sometimes called the
205: Hausdorff-Besicovitch dimension~\cite{Rogers}.  The possibly infinite
206: number $\ell^s({\scr{C}})$ gives an indication of the number of
207: subsets $U$ in $\scr{C}$ needed to cover $E$.  In order to effect the
208: approximation, the diameter of the sets $U$ in the cover is restricted
209: to be at most $\delta>0$.
210: 
211: Let 
212: \begin{equation*}
213:   \mathcal{H}^s_\delta(E):=\inf_{\scr{C}_\delta} 
214:   \sum_{U\in{\scr{C}}_\delta} (\di
215:   U)^{s} =\inf \ell^s({\scr{C_\delta}}),
216: \end{equation*}
217: where the infimum is taken over all covers $\scr{C}_\delta$ of $E$ by
218: sets $U$ with $\di U \leqslant \delta$; such covers are called
219: $\delta$-covers.  For a point $\mathbf{x}$,
220: $\mathcal{H}^s_\delta(\{\mathbf{x}\})=1$ when $s=0$ and vanishes when
221: $s>0$.  As $\delta$ decreases, $\mathcal{H}^s_\delta$ can only
222: increase as there are fewer $U$'s available, \emph{i.e.}, if
223: $0<\delta<\delta'$, then
224: \begin{equation*}
225:   \mathcal{H}^s_{\delta'}(E)\leqslant \mathcal{H}^s_{\delta}(E). 
226: \end{equation*}
227: The set function $\mathcal{H}^s_\delta$ is an outer measure on
228: $\mathbb{R}^n$ but the limit $\mathcal{H}^s$ (which can be infinite)
229: as $\delta \to 0$, given by
230: \begin{equation}
231: \label{eq:HHd}
232:   \mathcal{H}^s(E)=\lim_{\delta\to 0}\mathcal{H}^s_\delta (E) \ = 
233:   \sup_{\delta>0} \mathcal{H}^s_\delta (E)\in [0,\infty],
234: \end{equation}
235: is better behaved.  From its construction by covers,
236: $\mathcal{H}^s(E)\leqslant \mathcal{H}^s(F)$ for any $E\subset F$ and
237: indeed is subadditive and a regular outer measure.  The restriction to
238: the $\sigma$-field of $\mathcal{H}^s$ measurable sets (which includes
239: open and closed sets, limsup and liminf sets and $G_\delta$ and
240: $F_\sigma$ sets) is usually called the \emph{Hausdorff $s$-dimensional
241: measure}.  Hausdorff $1$-dimensional measure coincides with
242: 1-dimensional Lebesgue measure and in higher dimensions, Hausdorff
243: $n$-dimensional measure is comparable to $n$-dimensional Lebesgue
244: measure, \emph{i.e.},
245: \begin{equation*}
246:   \mathcal{H}^n(E)\asymp \vert E\vert ,
247: \end{equation*}
248: where $\vert E\vert $ is the Lebesgue measure of $E$ and where for $a,b>0$,
249: $a\asymp b$ means there exist constants $c,c'>0$ such that $a\leqslant
250: c b\leqslant c' a$ or $a=O(b)$, $b=O(a)$ in Landau's $O$-notation.
251: Thus a set of positive $n$-dimensional Lebesgue measure has positive
252: Hausdorff $n$-measure.
253: 
254: Because it is defined in terms of the diameter of the covering sets,
255: Hausdorff $s$-measure is unchanged by restriction to closed, convex or
256: open sets.  It is also unchanged by isometries and so in particular by
257: translations and rotations.  It is, however, affected by scaling in
258: the natural way (as are fractals): for any $r\geqslant 0$,
259: \begin{equation*}
260:   \mathcal{H}^s(r E)=r^s \mathcal{H}^s(E).
261: \end{equation*}
262: 
263: \subsection{Hausdorff dimension} 
264: \label{sec:HD}
265: Zero-dimensional Hausdorff measure $\mathcal{H}^0(E)$ is simply
266: counting measure; thus the Hausdorff $s$-measure of a set of $k$
267: points is $k$ when $s=0$ and 0 for $s>0$.  This pattern is typical.
268: When the set $E$ is infinite, $\mathcal{H}^s(E)$ is either 0 or
269: $\infty$, except for possibly one value of $s$. To see this, the
270: definition of $\mathcal{H}^{s}_\delta(E)$ implies that there is a
271: $\delta$-cover $\scr{C}_\delta$ of $E$ such that
272: \begin{equation*}
273:   \sum_{C\in\scr{C}_\delta}(\di C)^{s} \leqslant
274:   \mathcal{H}^{s}_\delta(E)+1 \leqslant \mathcal{H}^{s}(E)+1 \leqslant
275:   \infty.
276: \end{equation*}
277: Suppose that $\mathcal{H}^{s_{_0}}(E)$ is finite and
278: $s=s_0+\varepsilon$, $\varepsilon>0$.  Then for each member $C$ of the
279: cover $\scr{C}_\delta$, $(\di C)^{s_0+\varepsilon}\leqslant
280: \delta^\varepsilon (\di C)^{s_0}$, so that the sum
281: \begin{equation*}
282:   \sum_{C\in\scr{C}_\delta}( \di C)^{s_0+\varepsilon}\leqslant
283:   \delta^\varepsilon \sum_{C\in\scr{C}_\delta}(\di C)^{s_0}.
284: \end{equation*}
285: Hence 
286: \begin{equation*}
287:   \mathcal{H}_\delta^{s_0+\varepsilon}(E)\leqslant
288:   \sum_{C\in\scr{C}_\delta} ( \di C)^{s_0+\varepsilon} \leqslant
289:   \delta^\varepsilon \sum_{C\in\scr{C}_\delta} (\di
290:   C)^{s_0}\leqslant\delta^\varepsilon( \mathcal{H}^{s_{_0}}(E)+1)
291: \end{equation*}
292: and so
293: \begin{equation*}
294:   0\leqslant \mathcal{H}^s(E)= \mathcal{H}^{s_{_0}+\varepsilon}(E)=
295:   \lim_{\delta\to0}\mathcal{H}_\delta^{s_{_0}+\varepsilon}(E)\leqslant
296:   \lim_{\delta\to0} \delta^\varepsilon( \mathcal{H}^{s_0}(E)+1)=0.
297: \end{equation*}
298: On the other hand suppose $\mathcal{H}^{s_{_0}}(E)>0$.  If for any
299: $\varepsilon>0$, $\mathcal{H}^{s_{_0}-\varepsilon}(E)$ were finite,
300: then by the above $\mathcal{H}^{s_{_0}}(E)=0$, a contradiction, whence
301: $\mathcal{H}^{s-\varepsilon}(E)=\infty$.
302: \begin{figure}[htb]
303:   \begin{center}
304:     \begin{picture}(300,220) 
305:       %horizontal axis
306:       \put (40,10){\line(1,0){220}}
307:       \thicklines
308:       \put (40,200){\line(1,0){140}}
309:       \put (180,10){\vector(1,0){100}}
310:       \thinlines 
311:       \put (180,0){$s_0=\hdim{E}$}
312:       \put (40,0){$0$}
313:       \put (180,45){\circle*{3}}
314:       \put (270,0){$s$}
315:       %vertical axis
316:       \put (40,10){\line(0,1){115}}
317:       \multiput (40,120)(0,10){8}{\line(0,1){5}}
318:       \put (15,200){$\infty$}
319:       \put (40,200){\circle*{4}}
320:       \put (40,45){\circle*{3}}
321:       \multiput (180,10)(0,20){6}{\line(0,1){10}}
322:       \multiput (180,130)(0,10){7}{\line(0,1){5}}
323:       \put (0,120){$\mathcal H^s(E)$}
324:       \put (-5,45){$\mathcal H^{s_0}(E)$}
325:     \end{picture}
326:     \caption{ 
327:       The graph of the Hausdorff measure of $E$.  The Hausdorff
328:       dimension $\hdim{E}=s_0$, the point of discontinuity.  }
329:     \label{fig:HMgraph}
330:   \end{center}
331: \end{figure}
332: 
333: Thus for each infinite set $E$ in $n$-dimension\-al Euclidean space,
334: there exists a unique non-negative exponent $s_0$ such that
335: \begin{displaymath}
336:   \mathcal{H}^s(E)= 
337:   \begin{cases}
338:     \infty, 
339:     & \  0 \leqslant  s < s_0, \\
340:     0,  &    \  s_0 < s < \infty,   
341:   \end{cases}
342: \end{displaymath}
343: as shown in Figure~\ref{fig:HMgraph} (reproduced with the permission
344: of the Cambridge University Press from~\cite{MDAM}). The critical
345: exponent
346: \begin{equation}
347:   \label{eq:HDdef}
348:   s_0  = \inf\{s\in[0,\infty)\colon \mathcal{H}^s(E) = 0\}
349: \end{equation} 
350: where the Hausdorff $s$-measure crashes is called the \emph{Hausdorff
351: dimension} of the set $E$ and is denoted by $\hdim E$. Thus the
352: Hausdorff dimension of a finite set is 0, as it is for a countable
353: set.  It is clear that
354: \begin{equation}
355:   \label{eq:hdests}
356:   \text{if } \ \mathcal{H}^s(E)=0 \ \text{ then } \ \hdim E\leqslant
357:   s; \ \text{ and if } \mathcal{H}^s(E)>0 \ \text{ then } \ \hdim
358:   E\geqslant s.
359: \end{equation}
360: 
361: The Hausdorff dimension tells us nothing about the Hausdorff
362: $s$-measure at the critical exponent $s_0=\hdim E$, only that this is
363: the appropriate exponent to investigate the measure.  The sudden
364: change in Hausdorff $s$-measure at $s_0=\hdim E$ can be compared to the
365: focal length of a microscope.  If the lens is too close, the image
366: fills the eyepiece and cannot be resolved; if the lens is too far
367: away, the image is invisible.  At the focal length, the image is in
368: focus and can be resolved.
369: 
370: The main properties of Hausdorff dimension for sets in $\mathbb{R}^n$ 
371: are
372: \begin{enumerate}
373: \item[(i)] If $E \subseteq F$ then $\hdim E\leqslant \hdim F$.
374: \item[(ii)]  $\hdim E\leqslant n$.
375: \item[(iii)]  If  $\vert E\vert >0$, then $\hdim E = n$.\label{item:1}
376: \item[(iv)]  The dimension of a point is $0$.
377: \item[(v)] If $\hdim E < n$, then $\vert E\vert =0$ 
378:   (however $\hdim E =n$ does not imply $\vert E\vert >0$). 
379: \item[(vi)] $\hdim (E_1\times E_2) \geqslant \hdim E_1+\hdim E_2$
380: \item[(vii)] $\hdim \cup_{j=1}^\infty E_j=\sup\{\hdim E_j\colon 
381:   j\in\mathbb{N}\}$.
382: \end{enumerate}
383: It can be shown that the Hausdorff dimension of any countable set is 0
384: and that of any open set in $\mathbb{R}^n$ is
385: $n$~\cite[p.~29]{FalcFG}.  The nature of the construction of Hausdorff
386: measure ensures that the Hausdorff dimension of a set is unchanged by
387: an invertible transformation which is bi-Lipschitz.  This implies that
388: for any set $S \subseteq \mathbb{R} \setminus\{0\}$, $\hdim
389: S^{-1}=\hdim S$, where $S^{-1}=\{s^{-1}\colon s\in S\}$.  Thus on the
390: whole, Hausdorff dimension behaves as a dimension should, although the
391: natural formula 
392: \begin{displaymath}
393:   \hdim( E_1\times E_2)= \hdim E_1+\hdim E_2
394: \end{displaymath}
395: does \emph{not} always hold~\cite[\S5.3]{FalcGFSshort} (it does hold
396: for certain sets, \emph{e.g.}, cylinders, such as $E \times I$, where
397: I is an interval: $\hdim (E\times I)=\hdim E + \hdim I = \hdim E+1$ by
398: (iii), see~\cite{MDAM}).
399: 
400: The general character of $\delta$-covers in the definition of
401: Hausdorff outer measure can be difficult to work with and for many
402: applications in higher dimensions, it is convenient to restrict the
403: elements in the $\delta$-covers of a set to simpler sets such as balls
404: or cubes.  For example, covers consisting of hypercubes
405: \begin{displaymath}
406:   H=\{\mathbf{x}\in\mathbb{R}^n\colon
407:   \vert \mathbf{x}-\mathbf{a}\vert_\infty<\delta\},
408: \end{displaymath} 
409: where $\vert \mathbf{x}\vert_\infty:=\max\{\vert x_j\vert \colon
410: 1\leqslant j\leqslant n\}$ is the {\em height} of
411: $\mathbf{x}\in\mathbb{R}^n$, centred at $\mathbf{a}\in\mathbb{R}^n$
412: and with sides of length $2\delta$ are used extensively.  While outer
413: measures corresponding to these more convenient restricted covers are
414: not the same as Hausdorff measure, they are comparable and so have the
415: same critical exponent~\cite[Chapter~5]{MattilaGS}.  Thus there is no
416: loss as far as dimension is concerned if the sets $U$ are chosen to be
417: balls or hypercubes. Of course, the two measures are identical for
418: sets with Hausdorff $s$-measure which is either $0$ or $\infty$. Such
419: sets are said to obey a `$0$-$\infty$' law, this being the appropriate
420: analogue of the more familiar `$0\,$-$1$' law in
421: probability~\cite[p~339]{KandT}.  Sets which do not satisfy a
422: $0$-$\infty$ law, \emph{i.e.}, sets which satisfy
423: \begin{equation}
424:   \label{eq:3}
425:   0<{\mathcal{H}}^{\hdim E}(E)<\infty,
426: \end{equation}
427: are called \emph{$s$-sets}; these occur surprisingly
428: often~\cite[p.~29]{FalcFG} and enjoy some nice properties (see
429: Chapters 2--4 of \cite{FalcGFSshort}).  One example is the Cantor set
430: which has Hausdorff $s$-measure 1 when $s=\log 2/\log
431: 3$~\cite[p.~14]{FalcGFSshort}.  However it seems that $s$-sets are of
432: less interest in Diophantine approximation where the sets that arise
433: naturally, such as the set of badly approximable numbers or the set of
434: numbers approximable to a given order (see next section), obey a
435: $0$-$\infty$ law. The first
436: steps in this direction were taken by Jarník, who proved that the
437: Hausdorff $s$-measure of set of numbers rationally approximable to
438: order $v$ (see~\S\ref{sec:toa}) was $0$ or
439: $\infty$~\cite{Ja30a,Ja30b}.  This result turns on an idea related to
440: density of Hausdorff measure.
441: \begin{lemma}
442:   \label{lem:qis}
443:   Let $E$ be a null set in $\mathbb{R}$. Suppose that   for any interval 
444:   $(a,b)$
445:   and  $s\in [0,1]$,
446:   \begin{equation}
447:     \label{eq:qind}
448:     \mathcal{H}^s(E\cap(a,b)) \leqslant K(b-a) \mathcal{H}^s(E).  
449:   \end{equation}
450:   Then $\mathcal{H}^s(E)=0$ or $\infty$.
451: \end{lemma}
452: \begin{proof}
453:   Suppose the contrary, \emph{i.e.}, suppose $0 < \mathcal{H}^s(E)
454:   <\infty $.  Since $E$ is null, given $\varepsilon>0$, there exists a
455:   cover of $E$ by open intervals $(a_j,b_j)$ such that
456:   \begin{displaymath}
457:     \sum_j (b_j-a_j) < \varepsilon.
458:   \end{displaymath}
459:   By~\eqref{eq:qind}, there exists a constant $K>0$ such that
460:   \begin{equation*}
461:     0 < \mathcal{H}^s(E) = \mathcal{H}^s(\cup_j (a_j,b_j)\cap
462:     E) \leqslant K\mathcal{H}^s(E)\sum_j(b_j-a_j) <K \varepsilon
463:     \mathcal{H}^s(E)< \mathcal{H}^s(E)
464:   \end{equation*}
465:   for $\varepsilon<1/K$, a contradiction. 
466: \end{proof}
467: 
468: The proof for a general outer measure is essentially the same.  The
469: sets we encounter in Diophantine approximation are generally not
470: $s$-sets and some satisfy this `quasi-independence' property.
471: 
472: For other definitions of dimension, such as box-counting and packing
473: dimension, and their relationship with Hausdorff dimension,
474: see~\cite[Chapter~3]{FalcFG}.
475: 
476: \subsection{The determination of Hausdorff dimension}
477: Unless some general result is available, the Hausdorff dimension
478: $\hdim E$ of a null set $E$ is usually determined in two steps, with
479: the correct upward inequality $\hdim E\leqslant s_0$ and downward
480: inequality $\hdim E\geqslant s_0$ being established separately.
481: 
482: \subsubsection{The upper bound}
483: \label{sec:ub}
484: In view of~\eqref{eq:hdests}, an upper bound can be obtained by
485: finding a value of $s$ for which $\mathcal{H}^s(E)$ vanishes.  To find
486: such a value, it suffices to exhibit a cover $\{H\}$ of $E$
487: ($E\subseteq \cup H$) by hypercubes $H$ of arbitrarily small
488: sidelength and $s$-length.  This can often be done by adapting the
489: estimate involved in showing that Lebesgue measure is 0. When $E$ is a
490: limsup set, \emph{i.e.},
491: \begin{equation*}
492:   E=\limsup_{N\to\infty}E_N=\bigcap_{N=1}^\infty 
493:   \bigcup_{k=N}^\infty
494:   E_k
495:   =\{\mathbf{x}\in\mathbb{R}^n\colon \mathbf{x}\in E_k  
496:   \mathrm{ \ for\ infinitely\ many \ } k\in \mathbb{N}\}
497: \end{equation*}
498: for a sequence of sets $E_n$, a simple Hausdorff measure counterpart
499: of the convergence case of the Borel-Cantelli lemma often gives the
500: correct upper bound for $\hdim E$.  This is useful in Diophantine
501: approximation.
502: \begin{lemma}
503:   \label{lem:HCLemma}  
504:   Let
505:   \begin{equation*}
506:     E= \{\mathbf{x}\in\mathbb{R}^n\colon \mathbf{x}\in E_k
507:     \mathrm{ \ for\ infinitely\ many \ } k\in \mathbb{N}\}. 
508:   \end{equation*}
509:   If for some $s>0$,  
510:   \begin{equation}
511:     \label{eq:slength}
512:     \sum_{k=1}^\infty \di(E_k)^s < \infty,   
513:   \end{equation}
514:   then $\mathcal{H}^s(E)=0$ and $\hdim E\leqslant s$.
515: \end{lemma}
516: \begin{proof}
517:   From the definition, for each
518:   $N=1,2,\dots$,
519:   \begin{equation*}
520:     E\subseteq \bigcup_{k=N}^\infty E_k,
521:   \end{equation*}
522:   so that the family $\scr{C}^{(N)}=\{E_k\colon k\geqslant N\}$ is a
523:   cover for $E$.  By~\eqref{eq:slength},
524:   \begin{displaymath}
525:     \lim_{N\to \infty} \sum_{k=N}^\infty \di(E_k)^s 
526:     =0.
527:   \end{displaymath}
528:   Hence $\lim_{k\to\infty} \di(E_k) = 0$ and therefore given
529:   $\delta>0$, $\scr{C}^{(N)}$ is a $\delta$-cover of $E$ for $N$
530:   sufficiently large. But 
531:   \begin{equation*}
532:     \mathcal{H}_\delta^s(E)
533:     =\inf_{\scr{C}_\delta} \sum_{U\in{\scr{C_\delta}}} (\di U)^{s}
534:     \leqslant   \ell^s(\scr{C}^{(N)})
535:     =\sum_{k=N}^\infty \di(E_k)^s  \to 0
536:   \end{equation*}
537:   as $N\to\infty$.  Thus $\mathcal{H}_\delta^s(E)=0$ and
538:   by~\eqref{eq:HHd}, $\mathcal{H}^s(E)=0$, whence $\hdim E\leqslant
539:   s$.
540: \end{proof}
541: 
542: \subsubsection{The lower bound}
543: \label{sec:lb}
544: The lower bound is often harder (though by no means always,
545: see~\cite{dd00}).  It requires showing that given \emph{any} $s<s_0$
546: and \emph{any} cover $\{C\}$ of $E$ with the diameters of the covering
547: elements arbitrarily small, the $s$-length $\sum_C (\di C)^s \geqslant
548: \delta$ for some positive $\delta$.  This can be very difficult and
549: has led to the development of a variety of methods.  In Diophantine
550: approximation, the \emph{regular systems} introduced by Baker and
551: W.~M.~Schmidt~\cite{BS} and the more general \emph{ubiquitous
552: systems}~\cite{DRV90a} depend on a good supply of approximating
553: elements (\emph{e.g.}, the rationals).  These and related techniques
554: have proved effective in obtaining lower bounds for the Hausdorff
555: dimension of sets of number theoretic interest (see the survey
556: article~\cite{bbdbaker} for more details). A more fundamental approach
557: is the so-called mass distribution principle.
558: \begin{lemma}
559:   Let $\mu$ be a measure supported on a bounded Borel set $E$ in
560:   $\mathbb{R}^n$. Suppose that for some $s\geqslant 0$, there are
561:   strictly positive constants $c$ and $\delta$ such that $\mu(B)
562:   \leqslant c\, (\di B)^s$ for any ball $B$ in $\mathbb{R}^n$ with
563:   $\di B\leqslant \delta$.  Then ${\mathcal{H}}^s(E)\geqslant
564:   \mu(E)/c$. 
565: \end{lemma}
566: \begin{proof}
567:   The proof is short.  Let $\{B_k\}$ be a $\delta$-cover of $E$ by
568:   balls $B_k$.  Then
569:   \begin{equation*}
570:     \mu(E)\leqslant \mu\left(\bigcup_k B_k\right) \leqslant \sum_k
571:     \mu(B_k) \leqslant c\sum_k (\di B)^s.
572:   \end{equation*}
573:   Taking infima over all such covers, we see that
574:   ${\mathcal{H}}_\delta^s(E)\geqslant \mu(E)/c$, whence on letting
575:   $\delta \to 0$,
576:   \begin{displaymath}
577:     {\mathcal{H}}^s(E)\geqslant \mu(E)/c>0.
578:   \end{displaymath}
579: \end{proof}
580: This simple lemma is surprisingly useful and gives the easy part of
581: Frostman's lemma~\cite{Frostman} which is now stated in full.  The
582: Vinogradov notation $a\ll b$ for $a,b>0$ means that $a=O(b)$.
583: \begin{lemma} 
584:   \label{lem:Frostman}
585:   Let $E$  be a Borel subset of $\mathbb{R}^n$.
586:   Then
587:   \begin{displaymath}
588:     \mathcal{H}^s(E)>0 
589:   \end{displaymath} 
590:   if and only if there exists a measure $\mu$ on $\mathbb{R}^n$
591:   supported on $E$ with $\mu(E)$ finite such that $\mu(B) \ll (\di B
592:   )^s$ for all sufficiently small balls $B$.
593: \end{lemma}
594: Thus if $E$ supports a probability measure $\mu$ ($\mu(E)=1$) with $
595: \mu(B)\ll (\di B )^s $ for all sufficiently small balls $B$, then
596: $\hdim E\geqslant s$. The converse is more difficult but can be proved
597: using net measures (see~\cite{CarlesonES,FalcFG,MattilaGS}).
598: 
599: \section{Diophantine approximation} 
600: At its simplest level, Diophantine approximation is concerned with
601: approximating real numbers by rationals. Hardy and Wright's classic
602: \emph{Introduction to the theory of numbers}~\cite{HW} contains an
603: excellent account while the more advanced~\cite{Casselshort,SchmidtDA}
604: are devoted wholly to Diophantine approximation.  The theory extends
605: to approximating vectors in $\mathbb{R}^n$ (simultaneous Diophantine
606: approximation) and to matrices (systems of linear forms).  For
607: simplicity, we will stick mainly to one particular direction in the
608: one dimensional real and complex cases and treat the extensions to
609: higher dimensions and other settings fairly briefly.  Since the
610: rationals $\mathbb{Q}$ are a dense subset of the real numbers
611: $\mathbb{R}$, given any real number $\alpha$ and any positive
612: $\varepsilon$, there exists a rational $p/q$ such that
613: \begin{equation}
614:   \label{eq:1}
615:   \left\vert \alpha - \dfrac{p}{q}\right\vert  < \varepsilon.
616: \end{equation}
617: The numerator $p$ is often of no interest and the size of the
618: expression
619: \begin{equation}
620:   \label{eq:4}
621:   \Vert q \alpha\Vert =\min\{\vert q \alpha -p\vert\colon 
622:   p\in\mathbb{Z}\},
623: \end{equation}
624: the distance of $q\alpha$ from the integers $\mathbb{Z}$, is
625: considered.  Although convenient, it will not be used much here in
626: order to keep the notational burden to a minimum.
627: 
628: In simultaneous Diophantine approximation, one considers the system of
629: $n$ inequalities
630: \begin{equation*}
631:   \left\vert \alpha_k-\dfrac{p_k}{q}\right\vert <\varepsilon, \ \ k=,1\dots,n.
632: \end{equation*}
633: This system can be expressed more concisely as a single vector
634: inequality with $\boldsymbol{\alpha}=(\alpha_1,\dots,\alpha_n)\in
635: \mathbb{R}^n$, $\mathbf{p}\in\mathbb{Z}^n$, $q\in\mathbb{N}$, by
636: considering the expression
637: \begin{equation*}
638:   \left\vert \boldsymbol{\alpha}-\dfrac{\mathbf{p}}{q}\right\vert_\infty ,
639: \end{equation*}
640: where for $\mathbf{x}\in\mathbb{R}^n$,
641: $\vert \mathbf{x}\vert_\infty=\max\{\vert x_1\vert ,\dots,\vert x_n\vert \}$ is the \emph{height}
642: of $\mathbf{x}$, or, on multiplying by $q$, the expression
643: \begin{equation*}
644:   \vert q\boldsymbol{\alpha}-\mathbf{p}\vert_\infty =\Vert
645:   q\boldsymbol{\alpha}\Vert =\max\{\Vert q\alpha_j\Vert\colon
646:   j=1,\dots,n\}.
647: \end{equation*}
648: The last inequality has a \emph{dual} or linear form version: given
649: $\boldsymbol{\alpha}\in\mathbb{R}^n$ and $\varepsilon>0$, one
650: considers the inequality
651: \begin{equation*}
652:   \vert\mathbf{q}\cdot\boldsymbol{\alpha}-p\,\vert<\varepsilon,
653: \end{equation*}
654: where $\mathbf{q}\in\mathbb{Z}^n$ and $p\in\mathbb{Z}$.
655: The last two inequalities can be combined into a single 
656: general one.  The
657: system of $n$ real linear forms
658: \begin{equation*}
659:   \xi_1 a_{1j} + \dots + \xi_m a_{mj}, \ j=1,\dots,n,
660: \end{equation*}
661: in $m$ real variables $\xi_1,\dots,\xi_n$, can be written more
662: concisely as $\boldsymbol{\xi} A$, where $A=(a_{ij})$ and the system
663: of $n$ inequalities in $m$ variables
664: \begin{align*}
665:   &\vert  q_1 a_{11} + \dots + q_m a_{m1} -p_1\vert < \varepsilon\\
666:   &\qquad \qquad\qquad \qquad \vdots \\
667:   & \vert  q_1 a_{1n} + \dots + q_m a_{mn} -p_n\vert < \varepsilon,
668: \end{align*}
669: can be written $\vert \mathbf{q} A-\mathbf{p}\vert_\infty <
670: \varepsilon$.  Further details are in~\cite{Casselshort,HW,SchmidtDA}.
671: The theory extends naturally to the fields of $p$-adic
672: numbers~\cite{CasselsLF,Lutz} and formal power
673: series~\cite{kristensen03,kristensen02}.  Less obviously, it also
674: extends to discrete groups acting on hyperbolic space.  This is
675: relevant to Diophantine approximation over the Gaussian integers or
676: rationals considered below in~\S\ref{sec:DAGI}, so an outline is now
677: given.
678: 
679: The hyperbolic space setting sprang from the observations that the the
680: real axis is the set of limit points of the rationals and that the
681: rationals can be characterised as the parabolic vertices of the
682: modular group $\SL(2,\mathbb{Z})$, \emph{i.e.}, as the orbit of the
683: point at infinity under the linear fractional or Möbius
684: transformations
685: \begin{equation}
686: \label{eq:Mobius}
687:   z\mapsto \dfrac{az+b}{cz+d}, \ a,b,c,d \in \mathbb{Z}, \ ad-bc=1,
688: \end{equation}
689: of the extended upper half plane
690: $\mathbb{H}^{\,2}=\{z=x+iy\in\mathbb{C}\colon
691: y\geqslant0\}\cup\{\infty\}$.  For each element
692: \begin{equation*}
693:   g= \left(\begin{matrix} a & b \\
694:       c & d\\
695:     \end{matrix}\right)
696: \end{equation*}
697: in $\SL(2,\mathbb{Z})$, the point $\infty$ in the extended real line
698: evidently goes to $a/c$ under the group action and is also a fixed
699: point of the map $z\mapsto z+1$.  The maps $g$ form the \emph{modular}
700: group, $\SL(2,\mathbb{Z})$, a discrete subgroup of
701: $\SL(2,\mathbb{R})$, which is essentially (modulo the centre) the
702: group of orientation preserving Möbius transformations of the upper
703: half plane $\mathbb{H}^{\,2}$ to itself.  When the upper half plane is
704: endowed with the hyperbolic metric derived from $d\rho = \abs{dz}/y$,
705: it is a model for two dimensional hyperbolic space
706: $(\mathbb{H}^{\,2},\rho)$. The Möbius group $M(\mathbb{H}^{\,2})$ is
707: the group of isometries of $(\mathbb{H}^{\,2},\rho)$. Because the
708: group $\SL(2,\mathbb{Z})$ is discrete, points in the orbit can
709: accumulate only on the boundary $\mathbb{R}\, \cup\, \{\infty\}$ of
710: $\mathbb{H}^{\,2}$ and because the group elements are isometries with
711: respect to the hyperbolic metric, the limit set of any orbit is
712: the extended real line $\mathbb{R}\, \cup\{\infty\}$. A discrete subgroup of
713: $M(\mathbb{H}^{\,2})$ is called a \emph{Fuchsian} group.  Further
714: details of this rich and beautiful theory are
715: in~\cite{AhlforsMT,AndersonHG,BeardonGDG,NichollsETDG,Patterson76a}
716: and there is a short survey in Chapter~7 of~\cite{MDAM}.
717: 
718: These observations allow the classical theory, including the metrical
719: theory, of Diophantine approximation to be translated into Fuchsian
720: groups acting on the hyperbolic plane and to the much more general
721: setting of \emph{Kleinian} groups acting on $(n+1)$-dimensional
722: hyperbolic space $(\mathbb{H}^{\,n+1},\rho)$, $n\geqslant 2$ (Kleinian
723: groups are the discrete subgroups of the M\"obius group
724: $M(\mathbb{H}^{\,n+1})$ of isometries of $\mathbb{H}^{\,n+1}$; further
725: details are in the references given). The \emph{Picard} group
726: $\SL(2,\mathbb{Z}[i])$ consists of $2\times 2$ matrices over
727: $\mathbb{Z}[i]$ with determinant 1 and has an action on
728: $\mathbb{C}$ given by 
729: \begin{equation}
730: \label{eq:Picard}
731:   z \mapsto  \frac{az + b}{cz+d},  
732: \ a,b,c,d \in \mathbb{Z}[i], \ ad-bc=1.
733: \end{equation}
734: The limit set of the Picard group is the extended complex plane (or
735: Riemann sphere) $\mathbb{C}\cup \{\infty\}$ and the orbit of the point
736: at $\infty$ under the group is the set of ratios of Gaussian integers
737: ~\cite{MP93}.
738: Thus the Picard group plays a role precisely parallel to that played
739: by the modular group, expressed by~\eqref{eq:Mobius}, in approximating
740: real numbers by ratios of integers.
741: 
742: In the literature cited, hyperbolic space is usually taken in the
743: equivalent Poincar\'e form of the open unit ball
744: $\mathbb{B}^{n+1}=\{\mathbf{x}\in \mathbb{R}^{n+1} :
745: \abs{\mathbf{x}}_2<1\}$, where the ball is now endowed with the  
746: equivalent hyperbolic metric $\rho$ given by
747: $d\rho=\abs{d\mathbf{x}}_2/(1-\abs{\mathbf{x}}_2^2)$. To ease
748: comparison, we shall adopt this viewpoint, even though the upper half
749: plane model, as used by Sullivan in~\cite{Sullivan82}, is more
750: natural for Diophantine approximation. We choose to consider the
751: $(n+1)$-dimensional hyperbolic space, as the results in Diophantine
752: approximation are results about the boundary of $\mathbb{H}^{\,n+1}$,
753: which is $n$-dimensional.
754: 
755: The analogue $\mathfrak{p}$ of the point at infinity for Kleinian
756: groups is not quite straightforward. First of all the nature of the
757: elements $g$ of the Kleinian group $G$ implies that each $g$ has at
758: most two fixed points on the boundary of the ball. The special point
759: $\mathfrak{p}$ is called \emph{a parabolic} fixed point if it is the
760: unique fixed point on the boundary of some element in $G$; otherwise
761: they are called \emph{hyperbolic} fixed points. The orbit of a special
762: point $\mathfrak{p}$ under the action of a Kleinian group $G$
763: corresponds to the rationals $\mathbb{Q}$. The limit set $\Lambda(G)$
764: of the orbit under $G$ of a point in $\mathbb{H}^{\,n+1}$ lies in the
765: boundary $\mathbb{S}^n$. Given $\alpha\in
766: \Lambda(G)\subseteq\mathbb{S}^n$, one considers the 
767: quantity 
768: \begin{equation*}
769:   \vert \alpha- g(\mathfrak{p})\vert_2,
770: \end{equation*}
771: where $\vert \cdot \vert_2$ is the usual Euclidean metric in
772: $\mathbb{R}^{n+1}$.  Analogues of the principal theorems in Diophantine
773: approximation have been obtained with relatively minor technical
774: restrictions and will be discussed below; a brief survey is in
775: Chapter~7 of~\cite{MDAM}.  There is a striking dynamical
776: interpretation of the approximation in terms of flows on the
777: associated quotient space $\mathbb{H}^{n+1}/G$; more details are 
778: in~\cite{MDAM,FM95,MP93,Sullivan82,VelaniKTHD}.  
779: We now return to the one-dimensional theory.
780: 
781: \subsection{Dirichlet's theorem}
782: 
783: It is not difficult to make~\eqref{eq:1} more precise: given any real
784: number $\alpha$ and any positive integer $q$, there exists an integer $p$
785: such that $\vert q\alpha-p\vert  < 1$, and indeed such that
786: \begin{displaymath}
787:   \left\vert \alpha - \dfrac{p}{q}\right\vert  \leqslant \dfrac{1}{2q}.
788: \end{displaymath}
789: There are  denominators $q$ for which more can be said by
790: using Dirichlet's celebrated  `box argument' 
791: (see~\cite{Casselshort,HW}).
792: \begin{theorem}
793:   \label{thm:DT}
794:   For each real number $\alpha$ and any positive integer $N \geq 1$,
795:   there exists a rational $p/q$ with denominator satisfying
796:   $1\leqslant q \leqslant N$, such that
797:   \begin{equation*}
798:     \left\vert\alpha - \dfrac{p}{q}\right\vert < \dfrac{1}{qN}
799:     \leqslant \dfrac{1}{q^2}.
800:   \end{equation*}
801: \end{theorem}
802: 
803: \begin{proof}
804:   Let $[\alpha]$ be the integer part of $\alpha$ and $\{\alpha\}$ its
805:   fractional part, so that $\alpha=[\alpha]+\{\alpha\}$.  Divide the
806:   interval $[0,1)$ into $N$ subintervals $[k/N,(k+1)/N)$, where
807:   $k=0,1,\dots,N-1$, of length $1/N$.  The $N+1$ numbers $ \{r
808:   \alpha\}$, $r=0,1,\dots,N$, fall into the interval [0,1) and so two,
809:   $\{r \alpha\}$, $\{r'\alpha\}$ say, must fall into the same
810:   subinterval, $[k/N,(k+1)/N)$ say. Suppose that $r > r'$. Then
811:   \begin{displaymath}
812:     \vert \{r \alpha\} -\{r'\alpha\}\vert = \vert
813:     r\alpha-[r\alpha]-r'\alpha+[r'\alpha]\vert = \vert q\alpha-p\vert
814:     < \dfrac{1}{N},
815:   \end{displaymath}
816:   where $q=r-r', p = [r\alpha]-[r'\alpha] \in \mathbb{Z}$ and $1
817:   \leqslant q \leqslant N$.  Dividing by $q$ gives the quantitative 
818:   inequality
819:   \begin{equation}
820:     \label{eq:DTq}
821:     \left\vert\alpha - \dfrac{p}{q}\right\vert < \dfrac{1}{qN}
822:   \end{equation}
823:   and since $1\leqslant q\leqslant N$, the final inequality is
824:   immediate.
825: \end{proof}
826: A nice sharpening is in~\cite[p.~1]{HarmanMNT}.  When $p,q$ are
827: restricted to having highest common factor 1, the inequality
828: \begin{equation}
829:   \label{eq:DTio}
830:   \left\vert\alpha - \dfrac{p}{q}\right\vert < \dfrac{1}{q^2}
831: \end{equation}
832: has only finitely many solutions if and only if $\alpha$ is a
833: rational. Thus almost all real numbers satisfy~\eqref{eq:DTio} for
834: infinitely many rationals $p/q$.  Without this
835: restriction,~\eqref{eq:DTio} holds infinitely often for all
836: $\alpha\in\mathbb{R}$.
837: 
838: Dirichlet's theorem is one of the fundamental results in the theory of
839: Diophantine approximation. It can be viewed as a result about covers
840: and plays a central part in the Jarník--Besicovitch theorem, discussed
841: below. The theorem generalises to the simultaneous Diophantine
842: approximation of $n$ real numbers
843: $\alpha_1,\dots,\alpha_n$~\cite[Theorem~200]{HW} and asserts that 
844: given $N\in\mathbb{N}$, there exists $q\in\mathbb{N}$, with 
845: $q\leqslant N$ and $\mathbf{p}=(p_1,\dots,p_n)\in\mathbb{Z}^n$
846: such that
847: \begin{equation*}
848:   \left\vert \boldsymbol{\alpha}-\dfrac{\mathbf{p}}{q}\right\vert_\infty
849:   <\dfrac{1}{qN^{1/n}}.
850: \end{equation*}
851: In particular, for simultaneous Diophantine approximation in the
852: plane, given any $\boldsymbol{\alpha} = (\alpha_1, \alpha_2) \in
853: \mathbb{R}^2$ and $N \in \mathbb{N}$, there exists $q \in \mathbb{N}$ 
854: with $q \leqslant N$ and $p_1, p_2 \in \mathbb{Z}$ such that
855: \begin{equation}
856:   \label{eq:5}
857:   \max \left\{ \abs{\alpha_1 - \dfrac{p_1}{q}}, \abs{\alpha_2 -
858:   \dfrac{p_2}{q}} \right\} < \dfrac{1}{q N^{1/2}} \leqslant q^{-3/2}. 
859: \end{equation}
860: 
861: There is a so-called \emph{dual} version: given
862: $\boldsymbol{\alpha}\in\mathbb{R}^n$ and $N\in\mathbb{N}$, there
863: exists $\mathbf{q}\in\mathbb{Z}^n$ and a $p\in\mathbb{Z}$ such that
864: \begin{equation*}
865:   \vert \mathbf{q}\cdot\boldsymbol{\alpha}-p\vert <N^{-n}. 
866: \end{equation*}
867: These can be combined into a result for systems of $n$ real linear
868: forms~\cite[Chapter~1, Theorem~VI]{Casselshort}:
869: \begin{equation*}
870:   \vert \mathbf{q} A-\mathbf{p}\vert_\infty < N^{-m/n}.  
871: \end{equation*}
872: 
873: In the setting of a Kleinian group acting on hyperbolic space, the
874: analogue of the denominator in Theorem~\ref{thm:DT} corresponding to
875: $g(\mathfrak{p})$ is defined to be
876: \begin{equation}
877:   \label{eq:lg}
878:   \lambda_g:= \vert \det(Dg\vert_0)\vert ^{-1}=\dfrac{1}{2} \cosh 
879:   \rho(0,g(0))\asymp
880:   e^{\rho(0,g(0))}
881: \end{equation}
882: in the ball model, and so $\lambda_g\to\infty$ as $\vert g(0)\vert \to
883: 1$, \emph{i.e.}, as the orbit of the origin moves towards the
884: boundary. Here $Dg\vert_0$ denotes the Jacobian of $g$ evaluated at
885: the origin. For finitely generated Fuchsian groups of the first kind
886: taken to be acting on the closed unit disc $\Delta$, the elements $g$
887: are of the form
888: \begin{equation*}
889:   g= \left(\begin{matrix} a & b \\
890:       \overline{b} & \overline{a}\\
891:     \end{matrix}\right), \ a,b \in \Delta, \vert a\vert^2- 
892:   \vert
893:   b\vert^2=1 
894: \end{equation*}
895: and $\lambda_g=2( \vert a\vert^2+ \vert b\vert^2)$. Hedlund's lemma,
896: which is a classical result in the theory of discontinuous groups, is a
897: partial analogue of Dirichlet's theorem.  Let $\zeta\in \mathbb{S}^1$,
898: the unit circle.  For any $\xi\in \mathbb{S}^1$ which is not a
899: parabolic point, there exist infinitely many $g\in G$ such that
900: \begin{equation*}
901:   \vert \xi-g(\zeta)\vert <\dfrac{C}{\lambda_g}\, ,
902: \end{equation*}
903: for some $C>0$.  A complete analogue of Dirichlet's theorem, including
904: the quantitative inequality~\eqref{eq:DTq}, was obtained by
905: Patterson~\cite{Patterson76a} for Fuchsian groups and later he and
906: others extended it to Kleinian
907: groups~\cite{Patterson76b,Stratmann94,SV95,VelaniKTHD}.  However, the
908: statements in the Kleinian group setting differ for parabolic and
909: hyperbolic fixed points and so for simplicity the result will be
910: stated when $G$ has a unique parabolic point $\mathfrak{p}$.  Let
911: $N\geqslant 2$. Then for any $\xi\in\Lambda(G)$, there exists a $g\in
912: G$ with $\lambda_g<N$ such that
913: \begin{equation*}
914:   \vert \xi- g(\mathfrak{p})\vert_2 <\dfrac{C}{\sqrt{N\lambda_g}},
915: \end{equation*}
916: where $C$ is a constant depending only on $G$. 
917: 
918: \subsection{Types of approximation}   
919: \label{sec:toa}
920: The equation~\eqref{eq:DTio} given by Dirichlet's theorem is
921: essentially best possible as Hurwitz~\cite{Hurwitz1891} showed that
922: each $\alpha \in \mathbb{R}$ satisfied the inequality
923: \begin{equation*}
924:   \left\vert \alpha-\dfrac{p}{q}\right\vert  < \dfrac{1}{\sqrt{5}q^2}
925: \end{equation*}
926: for infinitely many positive integers $q$ and that this was best
927: possible in the sense that the constant $1/\sqrt{5}$ cannot be reduced
928: for numbers $\alpha$ equivalent to the golden ratio $(\sqrt{5}\,+1)/2$
929: \cite{Casselshort,Forder63,HW,SchmidtDA}.
930: 
931: \subsubsection{Badly approximable numbers} 
932: \label{sec:BA}
933: A number $\beta\in \mathbb{R}$ is called \emph{badly approximable} or
934: of \emph{constant type} if there exists a $K=K(\beta)$ such that
935: \begin{equation}
936:   \label{eq:ba}
937:   \left\vert \beta-\dfrac{p}{q}\right\vert \geqslant \dfrac{K}{q^2} 
938: \end{equation}
939: for all $p/q\in\mathbb{Q}$. Using \eqref{eq:4}, \eqref{eq:ba} can be
940: written as $q \Vert q\beta \Vert \geqslant K$. In view of Hurwitz's
941: theorem, the constant $K<1/\sqrt5$. Quadratic irrationals, such as
942: $\sqrt2$ and the golden ratio $(\sqrt{5}+1)/2$, are badly
943: approximable. This is proved in~\cite[\S11.4]{HW} using the fact that
944: the partial quotients in the continued fraction expansion for a
945: quadratic irrational are periodic. However, the proof relies only on
946: the boundedness of the partial quotients, which therefore
947: characterises the badly approximable numbers. The set of badly
948: approximable numbers will be denoted by $\mathfrak{B}$. The notion
949: extends naturally to higher dimensions and to the more general
950: settings mentioned above. 
951: 
952: Badly approximable numbers are important in applications, particularly
953: in stability questions for certain dynamical systems
954: \cite{MR93g:11011}.  For example the `noble' numbers, which are
955: equivalent to the golden ratio, have been conjectured to be the most
956: robust in the breaking up of invariant tori~\cite{Mackay83}. One very
957: practical application involved the design of rocket casings. These
958: were made using ruled surfaces and vibrations from the motors were
959: propagated along the generators. To reduce the effects of resonance
960: and delay the onset of catastrophic vibration, the ratio of the
961: circumference to the length of the casing was chosen to be a quadratic
962: irrational (V.~I.~ Arno'ld, personal communication). The desirable
963: properties of badly approximable numbers (and in particular of the
964: golden ratio) appear to be related to their occurrence in nature. It
965: has recently been discovered that the ratio between two step heights
966: on the surface of certain quasi-crystals is given by the golden ratio
967: (see~\cite{chown02:_why} for statements of this result and additional
968: examples).
969: 
970: The notion of badly approximable numbers carries over to higher
971: dimensions, including systems of linear forms~\cite{Schmidt69},
972: $p$-adics~\cite{AA95a} and fields of formal power series
973: \cite{kristensen03} as well as to Kleinian groups acting on hyperbolic
974: space~\cite{BJ97,FM95,Patterson76a,Patterson89}.  In the hyperbolic
975: space setting, a point $\beta$ in $\Lambda(G)$ is said to be badly
976: approximable with respect to $\mathfrak{p}$ if there exists a positive
977: constant $K=K(\beta)$ such that
978: \begin{equation*}
979:   \abs{\beta - g(\mathfrak{p})}_2\geqslant K/\lambda_g
980: \end{equation*}
981: for all $g\in G$.
982: 
983: \subsubsection{Diophantine type}
984: \label{sec:Dt}
985: The concept of a badly approximable number has extensions to
986: restricted classes of real numbers and points in $\mathbb{R}^n$ that are 
987: useful in connection with stability and other questions
988: (see~\S\ref{sec:applns}) and fortunately enjoys full measure. Let
989: $K>0$, $v>1$.  The real number $\alpha$ is said to be of
990: \emph{Diophantine type} $(K,v)$~\cite{ArnoldMM} (the definition has
991: been altered slightly for consistency) if
992: \begin{equation*}
993:   \left\vert \alpha-\dfrac{p}{q}\right\vert \geqslant \dfrac{K}{q^{v+1}} 
994: \end{equation*}
995: for all rationals $p/q$; the set of numbers of Diophantine type
996: $(K,v)$ is denoted by $\mathcal{D}(K,v)$. The union
997: \begin{equation*}
998:   \mathcal{D}_v=\bigcup_{K>0}\mathcal{D}(K,v)
999: \end{equation*}
1000: consists of numbers of Diophantine type $v$ (\emph{i.e.}, of type
1001: $(K,v)$ for some $K>0$) and the union
1002: \begin{equation*}
1003:   \mathcal{D}= \bigcup_{v>1} \mathcal{D}_v= \bigcup_{K>0,\, v>1}\mathcal{D}(K,v)
1004: \end{equation*}
1005: is the set of numbers of Diophantine type $v$ for some $v>1$.  Note
1006: that $\mathcal{D}_1=\mathfrak{B}$, the set of badly approximable
1007: numbers.  We will see in~\S\ref{sec:ktamda} that $\mathfrak{B}$ is
1008: null but that when $v>1$, $\mathcal{D}_v$ has full measure, which is
1009: pleasing since points of Diophantine type have desirable approximation
1010: properties for certain applications.  Again, the notion of Diophantine
1011: type extends naturally to higher dimensions~\cite{DRV94}.  We will be
1012: particularly interested in the case of a single linear form and
1013: accordingly we extend the definition of Diophantine type for a real
1014: number to a point in $\mathbb{R}^n$. A point
1015: $\boldsymbol{\beta}\in\mathbb{R}^n$ is of (dual) Diophantine type
1016: $(K,v)$ if for all $p\in\mathbb{Z}$ and non-zero
1017: $\mathbf{q}\in\mathbb{Z}^n$,
1018: \begin{equation*}
1019:   \vert \mathbf{q}\cdot \boldsymbol{\beta}-p\vert \geqslant
1020:    \dfrac{K}{\vert \mathbf{q}\vert_\infty^{v}}.
1021: \end{equation*}
1022: 
1023: \subsubsection{Well approximable numbers} 
1024: \label{sec:wan} 
1025: 
1026: In applications, we will be interested in points $\alpha$ which are
1027: not of Diophantine type $(K,v)$ for any $K>0$, \emph{i.e.}, in one
1028: dimension with the set
1029: \begin{equation}
1030:   \label{eq:vwa1}
1031:   E_v=\left\{\alpha\in\mathbb{R}\colon  {\text{ for any }} \ K>0, 
1032:     \left\vert  \alpha-\dfrac{p}{q}\right\vert  <
1033:     \dfrac{K}{q^{v+1}}  {\text{ for some }} \ \dfrac{p}{q}\in\mathbb{Q}
1034:   \right\}=\mathbb{R}\setminus \mathcal{D}_v,
1035: \end{equation}
1036: the complement of $\mathcal{D}_v$.  These numbers are closely related
1037: to numbers which are \emph{rationally approximable to order
1038:   $v+1$}~\cite[\S11.4]{HW}, \emph{i.e.}, to the set
1039: \begin{equation}
1040:   \label{eq:Rv}
1041:   R_v=\left\{\alpha\in\mathbb{R}\colon  {\text{ for some }} \ K>0, 
1042:     \left\vert  \alpha-\dfrac{p}{q}\right\vert  <
1043:     \dfrac{K}{q^{v+1}}  {\text{ for infinitely many }} \ 
1044:     \dfrac{p}{q}\in\mathbb{Q} \right\}
1045: \end{equation}
1046: (the exponent $v+1$ in the denominator is a normalisation to keep the
1047: notation the same as elsewhere.) By Dirichlet's theorem, all real
1048: numbers are rationally approximable to order $2$ and quadratic
1049: irrationals are rationally approximable to order exactly $2$.
1050: 
1051: The constant $K$ in the definitions of $R_v$ and $E_v$ is of less
1052: significance than the exponent, which motivates the next definition.
1053: A number $\alpha$ which satisfies
1054: \begin{equation}
1055:   \label{eq:Wvineq}
1056:   \left\vert  \alpha-\dfrac{p}{q}\right\vert  < \dfrac{1}{q^{v+1}}
1057: \end{equation}
1058: for infinitely many $ p/q\in \mathbb{Q}$ will be called {\em
1059:   $v$-approximable}; if $v>1$, $\alpha$ is called \emph{very well
1060:   approximable}.  Thus Liouville numbers, which
1061: satisfy~\eqref{eq:Wvineq} for any $v$, are very well approximable.
1062: We let $W_v$ denote the limsup set of $v$-approximable numbers,
1063: \emph{i.e.},
1064: \begin{equation*}
1065:   \label{eq:Wv}
1066:   W_v=  \left\{\alpha\in\mathbb{R}\colon   
1067:     \left\vert  \alpha-\dfrac{p}{q}\right\vert  <
1068:     \dfrac{1}{q^{v+1}}  {\text{ for infinitely many }} \
1069:     \dfrac{p}{q}\in\mathbb{Q} \right\}. 
1070: \end{equation*}
1071: 
1072: The sets $E_v$, $R_v$, $W_v$ are related and decrease as $v$
1073: increases, as follows from the inclusions
1074: \begin{equation}
1075:   \label{eq:UEWincl}
1076:   R_{v+\varepsilon} \subset E_v\subset W_v\subset   R_{v},
1077: \end{equation}
1078: where $\varepsilon>0$ is arbitrary.  Clearly $W_v\subset R_v$.  Next
1079: consider the complementary inclusion $\mathbb{R}\setminus W_v\subset
1080: \mathcal{D}_v$ and let $ \alpha \not \in W_v$.  Then for all but
1081: finitely many positive integers $q$,
1082: \begin{equation*}
1083:   \left\vert \alpha-\dfrac{p}{q}\right\vert \geqslant
1084:   \dfrac{1}{q^{v+1}}, \quad j=1,\dots, r.
1085: \end{equation*}
1086: Moreover, for the exceptional values of $q$, say
1087: $q^{(1)},\dots,q^{(r)}$, 
1088: \begin{equation*}
1089:   K=\min\left\{\abs{q^{(j)}\alpha-p^{(j)}}\left(q^{(j)}\right)^v\colon 
1090:     j=1,\dots,r\right\} >0,
1091: \end{equation*}
1092: since $ q\alpha\in \mathbb{Z}$ implies $ k q\alpha\in\mathbb{Z}$ for
1093: each $k\in\mathbb{Z}$.  Thus $\vert \alpha-p/q\vert  \geqslant
1094: K/q^{v+1}$ for all $p/q\in \mathbb{Q}$ and so
1095: $\alpha\in\mathcal{D}_v$, \emph{i.e.}, $\alpha\notin E_v$.
1096: 
1097: To establish the final inclusion, let $\alpha\in R_{v+\varepsilon}$,
1098: so that for some $K>0$,
1099: \begin{equation*}
1100:   \left\vert\alpha-\dfrac{p}{q}\right\vert <
1101:   \dfrac{K}{q^{v+1+\varepsilon}}
1102: \end{equation*}
1103: for infinitely many $p/q\in\mathbb{Q}$. Given any $K'>0$, choose as we 
1104: may, a denominator $q_0$ sufficiently large so that
1105: $q_0^{-\varepsilon}\leqslant K'/K$. Then there exists a
1106: $p_0\in\mathbb{Z}$ such that
1107: \begin{equation*}
1108:   \left\vert \alpha-\dfrac{p_0}{q_0}\right\vert< 
1109:   \dfrac{K'}{q_0^{v+1}}
1110: \end{equation*}
1111: and $\alpha\in E_v$.  The Jarník--Besicovitch
1112: theorem~(\S\ref{sec:JBt}) gives us the Hausdorff dimension of $W_v$
1113: and allows us to deduce from~\eqref{eq:UEWincl} that all the sets have
1114: the same Hausdorff dimension.
1115: 
1116: \subsubsection{$\Psi$-approximable  numbers} 
1117: \label{sec:psiapp}
1118: Now we look at the set of points which enjoy a more general
1119: approximation by rationals.  A function $\Psi\colon
1120: \mathbb{N}\to\mathbb{R}^+$ such that
1121: \begin{equation*}
1122:   \label{eq:appfn}
1123:   \lim_{q\to\infty}\Psi(q)=0 
1124: \end{equation*}
1125: will be called an \emph{approximation function}; without loss of
1126: generality we can take $\Psi(q)\leqslant 1/(2q)$ and later we shall
1127: also assume that $q\Psi(q)$ is decreasing (by decreasing we mean
1128: non-increasing here and subsequently). A real 
1129: number $\alpha$ is said to be \emph{$\Psi$-approximable} if $\alpha$
1130: satisfies the inequality
1131: \begin{equation}
1132:   \label{eq:1d}  
1133:   \left\vert \alpha-\dfrac{p}{q} \right\vert  < \Psi(q)
1134: \end{equation}
1135: for infinitely many $p/q\in\mathbb{Q}$ (there should be no confusion
1136: with $v$-approximable numbers defined above). Note that this should
1137: not be confused with the inequality $\vert q\alpha-p\vert <\psi(q)$
1138: which is often considered, particularly in higher dimensions and that
1139: there are other definitions depending on the form of~\eqref{eq:1d},
1140: see for example \cite{bkm01}. The set of $\Psi$-approximable numbers
1141: in $\mathbb{R}$ will be denoted $W(\Psi)$.
1142: 
1143: Since for each $k\in \mathbb{Z}$, $(p+kq)/q \in \mathbb{Q}$ and 
1144: \begin{equation*}
1145:   \left\vert \alpha+k-\dfrac{p+kq}{q} \right\vert =\left\vert
1146:   \alpha-\dfrac{p}{q} \right\vert ,
1147: \end{equation*}
1148: it follows that $ W(\Psi)\cap[k,k+1)=(W(\Psi)\cap[0,1))+k$, so that
1149: $W(\Psi)$ can be decomposed into a union over unit intervals:
1150: \begin{displaymath}
1151:   W(\Psi)=\bigcup_{k\in\mathbb{Z}} (W(\Psi)\cap[k,k+1))
1152:   =\bigcup_{k\in\mathbb{Z}}  (W(\Psi)\cap[0,1))+k .
1153: \end{displaymath}
1154: 
1155: As with the other types of approximation, the definitions of very well
1156: approximable, $v$-approximable and $\Psi$-approximable numbers extend
1157: naturally to systems of linear forms, $p$-adic numbers, formal power
1158: series and to the hyperbolic setting. We will be interested in the
1159: case of a single linear form: we say with some abuse of notation that 
1160: a vector $\boldsymbol{\alpha}\in\mathbb{R}^n$ is $v$-approximable if
1161: \begin{equation}
1162:   \label{eq:valf}
1163:   \vert \mathbf{q}\cdot \boldsymbol{\alpha}-p\vert 
1164:   <\dfrac{1}{\vert \mathbf{q}\vert_\infty^{v}} 
1165: \end{equation}
1166: holds for infinitely many $\mathbf{q}\in\mathbb{Z}^n$ and
1167: $p\in\mathbb{Z}$; $\boldsymbol{\alpha}$ is very well approximable if
1168: $v>n$.  In the hyperbolic space setting, a point $\boldsymbol{\alpha}$
1169: in $\mathbb{H}^{\,n}$ is $\Psi$-approximable with respect to $G$ and
1170: $\mathfrak{p}$ if
1171: \begin{displaymath}
1172:   \left\vert\boldsymbol{\alpha}-g(\mathfrak{p})\right\vert_2
1173:   <\Psi(\lambda_g)
1174: \end{displaymath}
1175: for infinitely many $g\in G$. These and the other analogues of the
1176: real one-dimensional case will be discussed further when they arise
1177: below.
1178: 
1179: \section{Khintchine's  theorem and metrical Diophantine approximation}
1180: \label{sec:ktamda}
1181: Following earlier work of Borel~\cite{Borel12}, Khintchine gave an
1182: almost complete answer to the solubility of~\eqref{eq:1d} in terms of
1183: the measure of $W(\Psi)$.  In a series of papers in the 1920's on the
1184: Lebesgue measure of the sets $W(\Psi)$ and
1185: $\mathfrak{B}$~\cite{Kh24,Kh25,Kh25a,Kh26}, he laid the foundations of
1186: metrical Diophantine approximation.  This theory, which is closely
1187: related to probability, measure and ergodic theory, considers sets of
1188: solutions to Diophantine inequalities in terms of Lebesgue and other
1189: measures. As a result, $0$-$1$ laws are a feature of the Lebesgue part
1190: of the theory, as in Khintchine's theorem below (see
1191: also~\cite[\S2.2]{HarmanMNT}). In addition, because an exceptional set
1192: for which a result is invalid can be of measure zero, this can lead to
1193: theorems having a strikingly simple yet general character.
1194: \begin{theorem}[Khintchine]
1195:   \label{thm:K1}
1196:   \begin{equation*}
1197:     \vert W(\Psi)\cap[0,1)\vert = 
1198:     \begin{cases} 
1199:       0 , &\text{if }  \sum_{k=1}^\infty k\Psi(k)<\infty, \\
1200:       1, &\text{if } \ k\Psi(k) \ \text {is decreasing and }
1201:       \sum_{k=1}^\infty k\Psi(k)=\infty. 
1202:     \end{cases}
1203:   \end{equation*}
1204: \end{theorem}
1205: Thus $W(\Psi)$ is null when the series $\sum_k k\Psi(k)$ converges and
1206: is full when $k\Psi(k)$ decreases and the sum diverges. 
1207: Subsequently Khintchine extended the result to simultaneous
1208: Diophantine approximation~\cite{Kh26} (see also~\cite{Casselshort})
1209: and Groshev extended it to systems of linear forms~\cite{Sprindzuk}.
1210: In particular the measure of the set of points $(\alpha_1, \alpha_2)
1211: \in [0,1)^2$ such that 
1212: \begin{displaymath}
1213:   \max\left\{ \abs{\alpha_1 - \dfrac{p_1}{q}}, \abs{\alpha_2 -
1214:       \dfrac{p_2}{q}} \right\} < \Psi(q) 
1215: \end{displaymath}
1216: for infinitely many $p_1, p_2\in \mathbb{Z}$, $q \in \mathbb{N}$ is
1217: $0$ or $1$ accordingly as $\sum k^2 \Psi(k)^2$ converges or as $k
1218: \Psi(k)$ decreases and the sum diverges (\emph{cf.} Theorem
1219: \ref{thm:complexKT} below).
1220: 
1221: The convergence case follows readily from the Borel-Cantelli
1222: lemma~\cite{KandT} and the argument is closely related to that of
1223: Lemma~\ref{lem:HCLemma}. However, the case of divergence is much more
1224: difficult and relies on a crucial `pairwise quasi-independence' result
1225: combined with `mean and variance' ideas or with results from ergodic
1226: theory.  For further details of this remarkable improvement of the
1227: Borel-Cantelli lemma,
1228: see~\cite{Chung,ChungErdos52,HarmanMNT,KS64,RenyiPT,Sprindzuk,Sullivan84}
1229: (the books~\cite{HarmanMNT,Sprindzuk} also include accounts of
1230: W.~M.~Schmidt's important quantitative extension of Khintchine's
1231: theorem~\cite{Schmidt60}).  In terms of probability, we recall that
1232: the events $E_j$, $j=1,\dots,\infty$, with corresponding probabilities
1233: $P(E_j)$ are pairwise independent if for any $j\ne k$, the probability
1234: of the two events occurring is given by
1235: \begin{displaymath}
1236:    P(E_j\cap E_k)=P(E_j)P(E_k).
1237: \end{displaymath}
1238: As a result, the divergence half of the Borel-Cantelli Lemma holds for
1239: pairwise independence.  Total independence, where the probability of
1240: any finite sequence of events is given by the product of the
1241: individual probabilities, is not necessary. Recently it has been shown
1242: that the divergence case is also related to the lower bound for
1243: Hausdorff dimension (see \S\ref{sec:cJBt}).
1244: 
1245: There has been dramatic progress in the metrical theory over the last
1246: decades, particularly in the theory of `dependent variables' where the
1247: point $\boldsymbol{\alpha}$ lies on a manifold, so that coordinates
1248: of the point are functionally related. Sprind{\v z}uk's
1249: solution~\cite{Spr80} of Mahler's conjecture in transcendence theory
1250: in terms of the Diophantine approximation of points on the Veronese
1251: curve $\{(x,x^2,\dots,x^n)\colon x\in\mathbb{R}\}$ gave this topic an
1252: enormous impetus which has seen the recent proof of Sprind{\v z}uk's
1253: conjectures and further results~\cite{bbdd99,bkm01,KM98a}. 
1254: 
1255: The $p$-adic analogue of Khintchine's theorem was obtained by
1256: Jarník~\cite{Ja45} and extended to systems of linear forms by
1257: Lutz~\cite{Lutz}. For fields of formal power series, the analogue was
1258: obtained by de~Mathan~\cite{deMa70} and has recently been extended to
1259: systems of linear forms~\cite{kristensen02}.  The complex analogue of
1260: Khintchine's theorem, in which the approximation is by ratios of
1261: Gaussian integers, is discussed in~\S\ref{sec:DAGI} below.
1262: 
1263: Khintchine's theorem corresponds to our intuition since if the
1264: approximation function $\Psi$ is large, then there is a better chance
1265: of the inequality being satisfied. In particular, the set $W_v$ is
1266: null for $v>1$, since the series $\sum_k k^{-v}$ converges, and full for
1267: $v\leqslant 1$ (when $W_v=[0,1)$ by Dirichlet's theorem).
1268: \begin{cor}
1269:   \begin{equation*}
1270:     \vert W_v\cap[0,1)\vert = \vert R_v\cap[0,1)\vert =\vert
1271:     E_v\cap[0,1)\vert = 
1272:     \begin{cases}
1273:       0, & \text{when } \ v>1,\\
1274:       1, & \text{when }\ v\leqslant 1.
1275:     \end{cases}
1276:   \end{equation*}
1277: \end{cor}
1278: The result for the other sets follows from~\eqref{eq:UEWincl}.  Less
1279: obviously, the theorem shows the Lebesgue measure of the set of
1280: $\alpha\in [0,1)$ such that~\eqref{eq:1d} has infinitely many
1281: solutions is $1$ when $\Psi(k) = 1/(k^2 \log k)$ and $0$ when $\Psi(k)
1282: = 1/(k^2(\log k)^{1+\varepsilon})$ for any positive $\varepsilon$.
1283: 
1284: As a consequence of Jarník's theorem on simultaneous Diophantine
1285: approximation to be discussed in~\S\ref{sec:JBt}, none of the sets
1286: $W_v,R_v,E_v$ is an $s$-set (see~\eqref{eq:3} for the definition).
1287: However, using Lemma~\ref{lem:qis} and invariance under rational
1288: translates of $R_v$, Jarník had shown earlier that the set $R_v$ of
1289: numbers rationally approximable to order $v$ is not an $s$-set and
1290: obeys a `$0$-$\infty$' law~\cite{Ja30a,Ja30b}. Although superseded by
1291: the above theorem, the argument is very nice but the papers cited are
1292: not readily available, so the proof is repeated here.  Let $k$ be a
1293: positive integer and let $\alpha\in R_v\cap[0,1/k]$, so that for some
1294: $K=K(\alpha)>0$,
1295: \begin{displaymath}
1296:   \left\vert \alpha - \dfrac{p}{q}\right\vert<\dfrac{K}{q^{v+1}}
1297: \end{displaymath}
1298: for infinitely many rationals $p/q$.  For each $j/k$, $1\leqslant
1299: j \leqslant k-1$, 
1300: \begin{equation*}
1301:   \left\vert\alpha - \dfrac{p}{q}\right\vert=  
1302:   \left\vert \alpha+\dfrac{j}{k} - \dfrac{p}{q}-\dfrac{j}{k}\right\vert 
1303:   < \dfrac{K k^{v+1}}{(kq)^{v+1}}.
1304: \end{equation*}
1305: Thus $\alpha+j/k\in R_v\cap[j/k,(j+1)/k]$ and it follows that for each
1306: $j=0,1,\dots,k-1$,
1307: \begin{displaymath}
1308:   R_v\cap \left[\dfrac{j}{k},\dfrac{j+1}{k}\right] = 
1309:   R_v\cap \left[0,\dfrac{1}{k} \right]+\dfrac{j}{k}.  
1310: \end{displaymath}
1311: Moreover since Hausdorff measure is translation invariant,
1312: \begin{displaymath}
1313:   \mathcal{H}^s\left(R_v\cap [0,1]\right)
1314:   =\sum_{j=0}^{k-1}\mathcal{H}^s\left(R_v\cap\left[
1315:   \dfrac{j}{k},\dfrac{j+1}{k} \right]\right)=
1316:   k \mathcal{H}^s\left(R_v\cap\left[0,\dfrac{1}{k}\right]\right),
1317: \end{displaymath}
1318: whence for any $j,k$ with $1\leqslant j \leqslant k-1$,
1319: \begin{displaymath}
1320:   \mathcal{H}^s\left(R_v\cap\left[\dfrac{j}{k},\dfrac{j+1}{k}
1321:     \right]\right)
1322:   =\dfrac{1}{k}\mathcal{H}^s\left(R_v\cap[0,1])\right).
1323: \end{displaymath}
1324: Now every open interval $(a,b)$ can be represented as a union of a
1325: countable set of intervals $[j/k,(j+1)/k]$, \emph{i.e.},
1326: \begin{displaymath}
1327:   (a,b) = \bigcup_{j,k} \left[\dfrac{j}{k},\dfrac{j+1}{k}\right],  
1328: \end{displaymath}
1329: so that $b-a = \sum_{j,k} 1/k$.  Hence
1330: \begin{displaymath}
1331:   \mathcal{H}^s((a,b)) = \mathcal{H}^s\left(\bigcup_{j,k} 
1332:   \left[\dfrac{j}{k},\dfrac{j+1}{k}\right]\right)  
1333: \end{displaymath}
1334: and since $\mathcal{H}^s(\cdot)$ is an outer measure,
1335: \begin{align*}
1336:   \mathcal{H}^s\left((a,b)\cap R_v\cap[0,1]\right) &= 
1337:   \mathcal{H}^s\left(\bigcup_{j,k} 
1338:   R_v\cap \left[\dfrac{j}{k},\dfrac{j+1}{k}\right]\right)
1339:   \leqslant \sum_{j,k} 
1340:   \mathcal{H}^s\left(R_v\cap\left[\dfrac{j}{k},\dfrac{j+1}{k}
1341:     \right]\right) \\ 
1342: &  \leqslant \mathcal{H}^s(R_v\cap[0,1])\sum_{j,k} \dfrac{1}{k}
1343:   =  (b-a)\mathcal{H}^s(R_v\cap[0,1]).
1344: \end{align*}
1345: Thus the hypotheses of Lemma~\ref{lem:qis} are satisfied and so
1346: $\mathcal{H}^s(R_v\cap[0,1])=\mathcal{H}^s(R_v)$ is $0$ or
1347: $\infty$. The argument can be extended to show that the set of numbers 
1348: approximable to order $v$ by algebraic irrationals is not an
1349: $s$-set. The measure at the critical exponent was shown to be $\infty$
1350: by Bugeaud \cite{MR1931199}.
1351: 
1352: Khintchine's theorem also implies that the set $\mathfrak{B}$ of badly
1353: approximable numbers is null. For given any $K>0$, the sum $\sum_q
1354: (K/q)$ diverges and so by Khintchine's theorem the set of real numbers
1355: $\alpha$ satisfying $ \vert \alpha-p/q\vert  < K/q^2 $ for infinitely
1356: many $p/q\in \mathbb{Q}$ is full.  Thus the complementary set $F(K)$
1357: of the set of $\alpha$ such that $\vert \alpha-p/q\vert  \geqslant
1358: K/q^2$ for all but finitely many $p/q$ is null and evidently increases
1359: as $K$ decreases. From its definition,
1360: \begin{equation*}
1361:   \mathfrak{B}\subset \bigcup_{K>0} F(K) = \bigcup_{N=1}^\infty
1362:   F(1/N),
1363: \end{equation*}
1364: a countable union of null sets, whence $\mathfrak{B}$ is null. 
1365: 
1366: Since the set $W_v$ of very well approximable numbers is null, the
1367: inclusions~\eqref{eq:UEWincl} imply that $\mathcal{D}_v$ is full for
1368: $v>1$, \emph{i.e.}, that almost all real numbers are of Diophantine
1369: type $(K,v)$ for some positive $K$.  Thus almost all numbers are
1370: neither well or badly approximable. It is a remarkable fact, proved by
1371: Jarník~\cite{Ja28} in 1928, the same year as Besicovitch's first paper
1372: on Hausdorff measure and dimension (on planar 1-sets)~\cite{Bes28},
1373: that although the set $\mathfrak{B}$ is null, $\hdim \mathfrak{B}=1$,
1374: \emph{i.e.}, its Hausdorff dimension is maximal in the sense that it
1375: coincides with that of the ambient space $\mathbb{R}$.  This we now
1376: discuss.
1377: 
1378: \subsection{Jarn\'\i k's theorem for badly approximable 
1379:   numbers}
1380: \label{sec:Jathmba}
1381: 
1382: Let $\theta\in (0,1)$ and let $a_n$, $n=1,2,\dots$, denote the partial
1383: quotients of the continued fraction expansion for $\theta$.  For each
1384: $N\in\mathbb{N}$, define
1385: \begin{equation*}
1386:   M_N:=\{\theta\colon a_n\leqslant N\}.
1387: \end{equation*}
1388: Now a number is badly approximable if and only if it has bounded
1389: partial quotients $a_n$~\cite[\S11.4]{HW}, so that
1390: $\mathfrak{B}=\lim_{N\to\infty} M_N$.  In a pioneering paper that was 
1391: the first on Hausdorff dimension in Diophantine approximation,
1392: Jarník~\cite{Ja28} showed that for each $N\geqslant 8$,
1393: \begin{equation*}
1394:   1-\dfrac{4}{N\log 2}\leqslant \hdim M_N \leqslant 1-\dfrac{1}{8N\log
1395:     N}.
1396: \end{equation*}
1397: It is of course immediate that $\hdim \mathfrak{B}\leqslant 1$ since
1398: $\mathfrak{B}\subset \mathbb{R}$. We now state Jarník's theorem for
1399: badly approximable numbers, which follows from the above.
1400: \begin{theorem}[Jarník]
1401:   \begin{displaymath}
1402:     \hdim \mathfrak{B} = 1.
1403:   \end{displaymath}
1404: \end{theorem}
1405: 
1406: Using $(\alpha,\beta)$ games, W.~M.~Schmidt~\cite{Schmidt66,Schmidt69}
1407: proved much more, extending Jarník's theorem to higher dimensions, so
1408: that the set of simultaneously badly approximable points in the plane
1409: has Hausdorff dimension $2$. In fact, $\mathfrak{B}$ is a \emph{thick}
1410: set.  This is a `local' property in the sense that for each open
1411: interval $I$, $\mathfrak{B}$ has `full' Hausdorff dimension,
1412: \emph{i.e.}, $\hdim \mathfrak{B} \cap I=1$. A very general
1413: inhomogeneous analogue has been proved using quite different ideas
1414: from dynamical systems~\cite{Kleinbock99}. 
1415: 
1416: The game in one dimension involves two players A and B, a non-empty
1417: set $S\subset \mathbb{R}$ and two parameters $\alpha\in (0,1)$, given
1418: to the player A, and $ \beta\in (0,1)$, given to the player B.  Player
1419: B begins by picking a closed interval $B_1$. Then A chooses a closed
1420: subinterval $A_1\subset B_1$ with $\vert A_1\vert =\alpha\vert
1421: B_1\vert $.  Then B picks an interval $B_2\subset A_1$ with $\vert
1422: B_2\vert =\beta\vert A_1\vert =\beta\alpha \vert B_1\vert $ 
1423: and then A chooses another subinterval $A_2\subset B_2$ with
1424: $\vert A_2\vert =\alpha \vert B_2\vert =\alpha^2\beta\vert B_1\vert $
1425: and so on.  Clearly the intervals $B_1, A_1, B_2, A_2, \dots,$ form a
1426: decreasing nested sequence so that their intersection is a point,
1427: $\omega$ say, in $B_1$.  Player A is called the winner if $\cap_j 
1428: A_j=\{\omega\}\subseteq S$, otherwise B wins.
1429: 
1430: A fuller account of the game in $\mathbb{C}$ will be given in
1431: \S\ref{sec:cba}, so we will simply say that Schmidt showed that when
1432: $S=\mathfrak{B}$, A can force $\omega$ to be badly approximable, even
1433: though $\mathfrak{B}$ is null, and deduced that $\hdim
1434: \mathfrak{B}\geqslant 1$.  Since $\mathfrak{B}\subset \mathbb{R}$,
1435: $\hdim \mathfrak{B}=1$.  Note that since
1436: $\mathcal{H}^1(\mathfrak{B})=\vert \mathfrak{B}\vert $, the Lebesgue
1437: measure of $\mathfrak{B}$ and since $\vert \mathfrak{B}\vert =0$
1438: (Khintchine's theorem), it follows that the Hausdorff measure of
1439: $\mathfrak{B}$ vanishes at $s=\hdim \mathfrak{B}$ and
1440: \begin{equation*}
1441:   \mathcal{H}^s(\mathfrak{B}) = 
1442:   \begin{cases}
1443:     \infty, & \   0 \leqslant  s < 1,\\ 
1444:     0,  &    \  s\geqslant1.
1445:   \end{cases}
1446: \end{equation*} 
1447: Thus $\mathfrak{B}$ is not an $s$-set (this can also be proved using
1448: Jarník's lemma in \S\ref{sec:HD} above).
1449: 
1450: Jarník's theorem has also been extended to $p$-adic
1451: fields~\cite{AA95a}, to fields of formal power 
1452: series~\cite{kristensen03} and to hyperbolic space.  It follows from
1453: the hyperbolic space counterpart of Khintchine's theorem that for
1454: geometrically finite Kleinian groups $G$, the set
1455: $\mathfrak{B}(G,\mathfrak{p})$ of the hyperbolic analogue of badly
1456: approximable points, has zero Patterson measure and the analogue of
1457: Jarník's theorem holds~\cite{FM95,Patterson76a,Patterson89},
1458: \emph{i.e.},
1459: \begin{equation*}
1460:   \hdim \mathfrak{B}(G,\mathfrak{p}) =\hdim \Lambda(G).
1461: \end{equation*}
1462: In addition, the exponent of convergence of $G$,
1463: \begin{equation*}
1464:   \delta(G):=  \inf\{s>0\colon \sum_{g\in 
1465:     G}\lambda_g^{-s}<\infty\}=
1466:   \hdim \Lambda(G)
1467: \end{equation*}
1468: \cite{BJ97,Patterson76b,Sullivan84}.  In view of
1469: Lemma~\ref{lem:HCLemma}, this is perhaps not so surprising.  In a
1470: striking parallel with continued fractions, badly approximable
1471: points correspond to bounded orbits of flows on
1472: manifolds~\cite{BJ97,Dani86a,DaniNTDS,FM95,Series85}.
1473: 
1474: 
1475: \subsection{Jarník--Besicovitch theorem}
1476: \label{sec:JBt}
1477: 
1478: When $v>1$, the set $\mathcal{D}_v$ is complementary to the set $R_v$
1479: of points approximable to exponent $v$, which is related to the set
1480: $W_v$ of $v$-approximable numbers (these statements also hold for the
1481: higher dimensional analogues). The Hausdorff dimension of $W_v$ was
1482: determined by Jarník in 1929~\cite{Ja29} and independently by
1483: Besicovitch in 1934~\cite{Bes34}.
1484: \begin{theorem}[Jarník--Besicovitch] 
1485:   \label{thm:JBthm}
1486:   When $v\geqslant 1$, 
1487:   \begin{equation*}
1488:     \hdim W_v = \dfrac{2}{v+1},
1489:   \end{equation*}
1490:   and when $v \leq 1$, $W_v = \mathbb{R}$.
1491: \end{theorem}
1492: Establishing the upper bound is not difficult, since $W_v$ is a limsup
1493: set. There is no loss of generality in working with the more
1494: convenient set $W_v\cap[0,1]$, as 
1495: $W_v=\bigcup_{k\in\mathbb{Z}}(W_v\cap[0,1] +k)$.  Consider the limsup set
1496: \begin{displaymath}
1497:   W_v\cap[0,1]=\bigcap_{N=1}^\infty \bigcup_{q=N}^\infty 
1498:   \bigcup_{p=0}^q
1499:   B(p/q,q^{-v-1}),
1500: \end{displaymath}
1501: where $B(p/q,\varepsilon)=\{x\in[0,1]\colon
1502: \abs{x-p/q}<\varepsilon\}$, so that
1503: $\di(B(p/q,\varepsilon)) \leqslant 2\varepsilon$. Then 
1504: \begin{displaymath}
1505:   \sum_{q=1}^\infty \sum_{p=0}^q \di(B(p/q,q^{-v-1}))^s \leqslant
1506:   2^s \sum_{q=1}^\infty  q^{1-s(v+1)}<\infty
1507: \end{displaymath}
1508: when $s>2/(v+1)$.  Hence by Lemma~\ref{lem:HCLemma} and the properties 
1509: of Hausdorff dimension, $\hdim W_v\leqslant 2/(v+1)$ when $v>1$.  
1510: When $v \leq 1$, the theorem follows from Dirichlet's theorem.
1511: 
1512: Establishing the correct lower bound is much harder. Jarník's lengthy
1513: and complicated proof involved continued fractions and arithmetic
1514: arguments.  Besicovitch's proof was simpler and more geometric and is
1515: the basis of regular and ubiquitous systems which have turned out to
1516: be very effective techniques in determining the Hausdorff dimension of
1517: a variety of sets~\cite{BS,bbdbaker,DRV90a}. Indeed ubiquity can imply
1518: Khintchine's theorem~\cite{BDV03}, so ideas developed for the study of
1519: the the Hausdorff dimension of the null sets also contribute
1520: substantially to our understanding of this theorem. The
1521: Jarník--Besicovitch theorem has been extended considerably, to higher
1522: dimensions, hyperbolic space and local fields.  Jarník himself proved
1523: the Hausdorff measure analogue of Khintchine's theorem for
1524: simultaneous Diophantine approximation~\cite{Ja31} and deduced that
1525: the set of points in $\mathbb{R}^n$ satisfying
1526: \begin{displaymath}
1527:   \abs{\boldsymbol{\alpha} - \dfrac{\mathbf{p}}{q}} < q^{-v-1}
1528: \end{displaymath}
1529: for infinitely many $\mathbf{p} \in \mathbb{Z}^n$, $q \in \mathbb{N}$
1530: has Hausdorff dimension $(n+1)/(v+1)$ when $v \geqslant 1/n$ and $n$
1531: otherwise. He also showed that the Hausdorff $s$-measure of $W_v$ at
1532: the critical dimension is infinite. In view of the
1533: inclusions~\eqref{eq:UEWincl}, it follows that the same holds for
1534: $E_v$, $R_v$.  
1535: The points in $W_v$ form an uncountable totally disconnected subset of
1536: the line and so are in Mandelbrot's picturesque language `fractal
1537: dust', as is $\mathfrak{B}$. 
1538: 
1539: We have seen that the notion of a very well approximable point extends
1540: naturally to systems of real and $p$-adic linear forms and to
1541: hyperbolic space (where they can be interpreted in terms of geodesic
1542: excursions~\cite{Dani85}).  The Jarník--Besicovitch theorem and the
1543: Hausdorff measure analogue of Khintchine's theorem have been
1544: established in the real case~\cite{HDSV97,mmd92}, the $p$-adic
1545: case~\cite{AA95,ddj99}, the formal power series
1546: case~\cite{kristensen02} and the hyperbolic case~\cite{HVJBGF,MP93}.
1547: Other generalisations are to restricted sequences~\cite{Rynne92b},
1548: inhomogeneous Diophantine approximation~\cite{JL98} and to small
1549: linear forms~\cite{HD93}.  A further generalisation to `shrinking
1550: targets' has revealed some unexpected connections with complex
1551: dynamics and ergodic theory~\cite{HV95}.
1552: 
1553: We will be interested in the case of a single real linear form
1554: in~\S\ref{sec:applns}.  The set of $v$-approximable points
1555: $\boldsymbol{\alpha}\in\mathbb{R}^n$ (see \eqref{eq:valf}) will be
1556: denoted $L_v$, \emph{i.e.},
1557: \begin{equation}
1558:   \label{eq:Lv}
1559:   L_v=\{\boldsymbol{\alpha}\colon
1560:   \abs{\mathbf{q}\cdot\boldsymbol{\alpha}-p}<\abs{\mathbf{q}}_\infty^{-v}
1561:   \text{ for infinitely many } \ \mathbf{q}\in\mathbb{Z}^n,
1562:   p\in\mathbb{Z}\,\}
1563: \end{equation}
1564: and by the dual or linear form version of the 
1565: Jarník--Besicovitch theorem~\cite{BD86},
1566: \begin{equation}
1567:   \label{eq:dimLv}
1568:   \hdim L_v=
1569:   \begin{cases}
1570:     n-1+\dfrac{n+1}{v+1} &\text{when } \ v>n, \\
1571:     n                  &\text{when }\  v\leqslant n. \\
1572:   \end{cases}
1573: \end{equation}
1574: The $n-1$ term arises from the dimension of the resonant hyperplanes
1575: \begin{displaymath}
1576:   \{\mathbf{x}\in\mathbb{R}^n\colon  \mathbf{q}\cdot \mathbf{x}=p\}  
1577: \end{displaymath}
1578: sets while the other represents `fractal dust' normal to the
1579: hyperplanes.
1580: 
1581: \subsection{Approximation by ratios of Gaussian integers}
1582: \label{sec:DAGI}
1583: Recall that the Gaussian integers are defined as the set
1584: $\mathbb{Z}[i] = \{p_1 + i p_2 \in \mathbb{C} : p_1, p_2 \in
1585: \mathbb{Z}\}$. These form a ring (in fact a unique factorisation
1586: domain). The Gaussian rationals are defined as the set $\mathbb{Q}(i)
1587: = \{a/b + i c/d \colon a/b, c/d \in \mathbb{Q}\}$.
1588: Approximation by Gaussian rationals decouples into the independent
1589: approximation of the real and imaginary part respectively. Here we
1590: study the more interesting problem of approximation by ratios of
1591: Gaussian integers.
1592: 
1593: Approximation of complex numbers by ratios of Gaussian integers was
1594: studied by Hermite and Hurwitz in the 19th
1595: century~\cite[IV,~\S4]{Koksma} but, unlike in the real case, a
1596: continued fraction approach did not give the best possible analogue of
1597: Dirichlet's theorem.  This was obtained in 1925 by
1598: Ford~\cite{Ford25}, who used additional geometrical ideas based
1599: on the Picard group $\SL(2,\mathbb{Z}[i])$. The Gaussian analogues of
1600: the theorems of Dirichlet, Khintchine, Jarn\'ik and Besicovitch
1601: treated below should be compared with those corresponding to
1602: simultaneous Diophantine approximation in the real plane
1603: $\mathbb{R}^2$. 
1604: 
1605: In 1967 A.~Schmidt introduced a theory of regular and dually regular
1606: chains for continued fractions, to treat approximation problems in
1607: complex numbers~\cite{ASchmidt67,ASchmidt75a}.  Schmidt was concerned
1608: with the study of complex quadratic irrationals and the complex
1609: version of Pell's equation and with extensions to
1610: groups~\cite{ASchmidt69,ASchmidt74}.  Our interest, however, is with
1611: complex or Gaussian rational analogues of Dirichlet's theorem and with
1612: the observation that the extended complex plane is the limit set of
1613: the Picard group.  In this connection, Patterson established analogues
1614: of Dirichlet's theorem for the less general Fuchsian
1615: groups~\cite{Patterson76a,Velani93}.  Later Stratmann and Velani
1616: obtained versions for Kleinian
1617: groups~\cite{Stratmann94,SV95,VelaniKTHD} and so for the Picard group.
1618: These results can be translated into complex versions of Dirichlet's
1619: theorem with an undetermined constant.  Nevertheless, for
1620: completeness, a short geometry of numbers proof of the complex version
1621: of Dirichlet's theorem is given below.  Although the constant here is
1622: not best possible, the result is all we need.  Proofs of the complex
1623: analogues of Jarník's theorem on badly approximable numbers and the
1624: Jarník--Besicovitch theorem that do not use the hyperbolic space
1625: framework will be given in \S\ref{sec:cba} and \S\ref{sec:cJBt} below.
1626: 
1627: Complex Diophantine approximation has also been investigated from the
1628: point of view of the distribution of the values of polynomials with
1629: real integer coefficients but with complex variable $z$~\cite{bd99};
1630: for another complex analogue see~\cite{Harman02}. \emph{For the rest
1631:   of this section, $p=p_1+ip_2$, $q=q_1+iq_2$ will denote Gaussian
1632:   integers with $q\ne 0$}.
1633: \begin{theorem}
1634:   Given any $z=x+iy\in \mathbb{C}$ and $N\in \mathbb{N}$, there exist
1635:   Gaussian integers $p=p_1+ip_2$, $q=q_1+iq_2$ with $0<\vert q\vert
1636:   \leqslant N$ such that
1637:   \begin{equation}
1638:     \label{eq:CDT2}
1639:     \left\vert z-\dfrac{p}{q}\right\vert  < \dfrac{2}{\vert q\vert N}.
1640:   \end{equation}
1641:   Moreover for infinitely many $p, q\in \mathbb{Z}[i]$,
1642:   \begin{equation}
1643:     \label{eq:Dio}
1644:     \left\vert z-\dfrac{p}{q}\right\vert  <  \dfrac{2}{\vert q\vert
1645:       ^2}. 
1646:   \end{equation}
1647: \end{theorem}
1648: 
1649: \begin{proof} 
1650:   The inequality~\eqref{eq:CDT2} holds if and only if the inequality
1651:   \begin{equation}
1652:     \label{eq:CDT}
1653:     \left\vert x+iy-\dfrac{p_1+ip_2}{q_1+iq_2}\right\vert  
1654:     < \dfrac{2}{\vert q_1+iq_2\vert N} 
1655:   \end{equation}
1656:   holds, \emph{i.e.}, if and only if
1657:   \begin{equation*}
1658:     \label{eq:CDT1}
1659:     \left\vert (q_1x-q_2y-p_1)+i(q_2x+q_1y-p_2)\right\vert 
1660:     <\dfrac{2}{N}
1661:   \end{equation*}
1662:   holds, which is the case if 
1663:   \begin{equation}
1664:     \label{eq:CDTht}
1665:     \max\{\vert q_1x-q_2y-p_1\vert , \vert q_2x+q_1y-p_2\vert
1666:     \}< \dfrac{\sqrt{2}}{N}. 
1667:   \end{equation}
1668:   By Minkowski's linear forms theorem, the system of inequalities
1669:   \begin{eqnarray*}
1670:     \vert q_1x-q_2y-p_1\vert  & < & 2^{1/2} \, N^{-1} \\
1671:     \vert q_2x+q_1y-p_2\vert  & < &  2^{1/2} \, N^{-1} \\
1672:     \vert q_1\vert  & \leqslant &  2^{-1/2} N \\
1673:     \vert q_2\vert  & \leqslant &  2^{-1/2} N
1674:   \end{eqnarray*}
1675:   has a non-zero solution in integers $p_1,p_2,q_1,q_2$.
1676:   Hence~\eqref{eq:CDTht} has a solution with $\vert q\vert =\vert
1677:   q_1+iq_2\vert \leqslant N$, as claimed.
1678: \end{proof}
1679: Since the Gaussian rationals $p/q$ are not required to be on lowest
1680: terms, \eqref{eq:Dio} holds infinitely often. For if $p/q$
1681: fails to satisfy~\eqref{eq:Dio}, then \emph{a fortiori}, $(\kappa
1682: p)/(\kappa q)$, where $\kappa$ is a non-zero Gaussian integer, will
1683: also fail. And if $p/q$ satisfies~\eqref{eq:Dio}, only a finite number
1684: of the fractions $(\kappa p)/(\kappa q)$ can also satisfy it.
1685: 
1686: This result should be compared to \eqref{eq:5}. As in the real case,
1687: complex numbers for which Dirichlet's theorem cannot be significantly
1688: improved are called badly approximable and numbers for which it can
1689: are called very well approximable. More precisely, a complex number
1690: $z$ is \emph{badly approximable} if there exists a constant $K=K(z)$
1691: such that for all $p,q \in \mathbb{Z}[i]$, $q \neq 0$, 
1692: \begin{equation*}
1693:   \left\vert z-\dfrac{p}{ q}\right\vert \geqslant \dfrac{K}{\vert
1694:     q\vert ^2}. 
1695: \end{equation*}
1696: Ford~\cite{Ford25} showed that the complex quadratic irrationals
1697: $(1\pm i\sqrt3)/2$ are the worst approximable numbers with
1698: $K=1/\sqrt3$ and thus correspond to the golden ratio $(\sqrt5+1)/2$ in
1699: the real case.  Badly approximable numbers have been studied by
1700: A.~Schmidt from the viewpoint of the Markoff
1701: spectrum~\cite{ASchmidt84}.
1702: 
1703: As in the real case, a complex number $z$ is 
1704: \emph{$v$-approximable} if
1705: \begin{equation}
1706: \label{eq:vapprox}
1707:   \left\vert z-\dfrac{p}{q}\right\vert < \dfrac{1}{\vert q\vert ^{v+1}}
1708: \end{equation}
1709: for infinitely many $p, q \in \mathbb{Z}(i)$ and will be called
1710: \emph{$\Psi$-approximable} if 
1711: \begin{equation}
1712:   \label{eq:ineq1}
1713:   \left\vert z-\dfrac{p}{q}\right\vert 
1714:   <\Psi(\vert q\vert),
1715: \end{equation} 
1716: where $\Psi: [1, \infty) \rightarrow \mathbb{R}^+$. The set of
1717: $\Psi$-approximable $z$ will be written $W^*(\Psi)$. Thus $W^*(\Psi)$
1718: is the set of points $z\in\mathbb{C}$ for which the
1719: inequality~\eqref{eq:ineq1} holds for infinitely many Gaussian 
1720: rationals $p/q$. The set of $v$-approximable complex numbers will be
1721: written $W_v^*$.
1722: 
1723: \subsection{Khintchine's theorem for complex numbers}
1724: 
1725: The complex analogue of Khintchine's theorem was proved in 1952 by
1726: LeVeque~\cite{leveque1952} who combined Khintchine's continued
1727: fraction approach with ideas from hyperbolic geometry.  In 1976,
1728: Patterson proved slightly less sharp versions of Khintchine's theorem
1729: for Fuchsian groups acting on hyperbolic
1730: space~\cite{Patterson76a,Velani93}.  A little later,
1731: Sullivan~\cite{Sullivan82} used Bianchi groups and some powerful
1732: hyperbolic geometry arguments to prove more general Khintchine
1733: theorems for real and for complex numbers.  In the latter case, the
1734: result includes approximation of complex numbers by ratios $a/b$ of
1735: integers $a,b$ from the imaginary quadratic field
1736: $\mathbb{R}(i\sqrt{d})$, where $d$ is a squarefree natural number. The
1737: case $d=1$ corresponds to the Picard group and approximation by ratios
1738: of Gaussian integers. Stratmann and Velani extended Patterson's
1739: results with similar minor technical restrictions to Kleinian
1740: groups~\cite{SV95,VelaniKTHD}. These include the Bianchi groups and
1741: give a less precise and differently formulated version of Sullivan's
1742: result. We now state LeVeque's result in our notation.
1743: \begin{theorem}
1744:   \label{thm:complexKT}
1745:   Suppose $k^2\Psi^(k)$ is decreasing. Then the Lebesgue measure of
1746:   $W^*(\Psi)$ is null or full according as the sum
1747:   \begin{equation}
1748:     \label{eq:Ksum}
1749:     \sum_{k=1}^\infty k^3\Psi(k)^2
1750:   \end{equation}
1751:   converges or diverges.
1752: \end{theorem}
1753: As well as being more general, Sullivan's result is more precise as
1754: the growth condition for the function $\Psi$ is weakened to a
1755: `comparability condition'. Instead of the sum, Sullivan and LeVeque
1756: use the equivalent integral $\int x^3\Psi(x)^2\,dx$ (their integrands
1757: are different owing to different forms of~\eqref{eq:ineq1}).  It is
1758: readily verified that $W^*(\Psi)$ is invariant under translations by
1759: Gaussian integers $p=p_1+ip_2$, so that
1760: \begin{equation}
1761:   \label{eq:2}
1762:   W^*(\Psi)= \bigcup_{p\in\mathbb{Z}[i]} V^*(\Psi)+p = 
1763:   \bigcup_{p_1,p_2\in\mathbb{Z}}
1764:   V^*(\Psi)+p_1+ip_2,
1765: \end{equation}
1766: where $V^*(\Psi)=W^*(\Psi)\cap I^2$ and $I^2=[0,1)^2=\{x+iy\colon
1767: 0\leqslant x,y < 1 \}$.  We will work in the more convenient unit
1768: square $I^2$ and consider the set $V^*(\Psi)$.
1769: 
1770: Before continuing, we need some more definitions and notation.
1771: 
1772: \subsection{Resonant sets  and balls in $I^2$}
1773: 
1774: Let $q=q_1+iq_2\in\mathbb{Z}[i]\setminus\{0\}$. The set
1775: $R(q)=R(q_1,q_2)\subset I^2$ where
1776: \begin{align*}
1777:   \label{eq:resetal}
1778:   R(q)&=\left\{\dfrac{p}{q}\colon p\in\mathbb{Z}[i]\right\}\cap I^2
1779:   =
1780:   R(q_1,q_2)
1781:   \\&=
1782:   \left\{\left(\dfrac{p_1q_1+p_2q_2}{q_1^2+q_2^2}, 
1783:       \dfrac{p_2q_1-p_1q_2}{q_1^2+q_2^2}\right)
1784:     \colon p_1,p_2\in\mathbb{Z}\right\}\cap I^2
1785: \end{align*}
1786: is called a \emph{resonant set}. This set is the analogue in the
1787: complex plane of the set $\{p_1/q_1\colon 0\leqslant p_1<q_1\}$ in the
1788: real line.  The points in the resonant set form a lattice inclined at
1789: an angle $\tan^{-1} (q_1/q_2)$ to the real axis and in which the side
1790: length of the fundamental region is $\vert q\vert
1791: =(q_1^2+q_2^2)^{-1/2}$.  Area and congruence considerations give that
1792: the number of points of $ R(q_1,q_2)$ in $I^2$ is
1793: \begin{equation}
1794:   \label{eq:pts}
1795:   \# R(q)= \# R(q_1,q_2)=\vert q\vert ^2 
1796:   = q_1^2+q_2^2.  
1797: \end{equation}
1798: 
1799: \begin{figure}[htbp]
1800:   \centering
1801:   \includegraphics[width=6cm,keepaspectratio]{lattice-print.eps}
1802:   \caption{The lattice in $I^2$ corresponding to $q_1=5$, $q_2=3$.} 
1803:   \label{fig:latt}
1804: \end{figure}
1805: 
1806: The disc 
1807: \begin{equation}
1808:   \label{eq:disc}
1809:   D(p/q;\varepsilon)
1810:   =\left\{z\in \mathbb{C}\colon
1811:     \left\vert z-p/q\right\vert < \varepsilon
1812:   \right\},
1813: \end{equation}
1814: with radius $\varepsilon$ and centred at $p/q$ where
1815: \begin{equation*}
1816:   \dfrac{p}{q}=\dfrac{p_1+ip_2}{q_1+iq_2}=
1817:   \dfrac{(p_1+ip_2)(q_1-iq_2)}{q_1^2+q_2^2}=
1818:   \dfrac{p_1q_1+p_2q_2}{q_1^2+q_2^2} +
1819:   i\dfrac{p_2q_1-p_1q_2}{q_1^2+q_2^2} 
1820: \end{equation*}
1821: has area $\pi \varepsilon^2$. The set
1822: \begin{equation}
1823:   \label{eq:Bcd}
1824:   B(q,\varepsilon)=\left\{z\in I^2 \colon
1825:     \left\vert z-\dfrac{p}{q}\right\vert <\varepsilon \text{ for some } p\in
1826:     \mathbb{Z}[i]\right\}=\bigcup_{p\in\mathbb{Z}[i]}
1827:   D(p/q,\varepsilon)
1828: \end{equation}
1829: can be regarded as a neighbourhood of a resonant set and its measure
1830: \begin{equation}
1831:   \label{eq:|Bcd|}
1832:   \vert  B(q,\varepsilon)\vert = \pi \varepsilon^2 \left(\vert q\vert ^2+O(\vert q\vert )\right). 
1833: \end{equation}
1834: Now $V^*(\Psi)$ can be expressed in the form
1835: \begin{equation}
1836:   \label{eq:limsupW}
1837:   V^*(\Psi)= \bigcap_{N=1}^\infty\bigcup_{k=N}^\infty 
1838:   \bigcup_{\vert q\vert =k}
1839:   B(q,\Psi(\vert q\vert)) =\limsup_{\vert q\vert \to\infty}
1840:   B(q,\Psi(\vert q\vert)).
1841: \end{equation}
1842: Thus $V^*(\Psi)$ is a limsup set and for each $N=1,2,\dots$  has a 
1843: natural cover
1844: \begin{equation}
1845:   \label{eq:cover}
1846:   \mathcal{C}_N(V^*(\Psi))=  
1847:   \{ B(q,\Psi(\vert q\vert)\colon \vert q\vert \geqslant N\}.
1848: \end{equation}
1849: Hence for each $N=1,2,\dots$, the measure of $V^*(\Psi)$ satisfies
1850: \begin{displaymath}
1851:   \vert V^*(\Psi) \vert \leqslant \sum_{k=N}^\infty \sum_{k \leqslant
1852:     \abs{q} < k+1} \vert B(q,\Psi(\vert q\vert)\vert \ll
1853:   \sum_{k=N}^\infty k^2 \Psi(k)^2 \sum_{k \leqslant \abs{q} < k+1} 1. 
1854: \end{displaymath}
1855: Now, $\sum_{1 \leqslant \abs{q} \leqslant k} 1$ is the number of
1856: lattice points in the closed disc $D(0,k)$ of radius $k$. By
1857: \cite{MR27:4799}, given any $\varepsilon > 0$, 
1858: \begin{displaymath}
1859:   \sum_{1 \leqslant \abs{q} \leqslant k} 1 = \pi k^2 + O\left(k^{12/37
1860:       + \varepsilon}\right).
1861: \end{displaymath}
1862: Hence 
1863: \begin{equation}
1864:   \label{eq:7}
1865:   \sum_{k \leqslant \abs{q} < k+1} 1 = \pi (k+1)^2 +
1866:   O\left(k^{1/3}\right) - \pi k^2 + O\left(k^{1/3}\right) = 2 \pi k
1867:   + O\left(k^{1/3}\right).
1868: \end{equation}
1869: Therefore
1870: \begin{displaymath}
1871:   \abs{V^*(\Psi)} \ll \sum_{k=N}^\infty k^3\Psi(k)^2.
1872: \end{displaymath}
1873: Since $N$ is arbitrary, the  convergence of the series $\sum_k
1874: k^3\Psi(k)^2$ implies that
1875: \begin{displaymath}
1876:   \abs{V^*(\Psi)}=\abs{W^*(\Psi)}=0,
1877: \end{displaymath}
1878: which is the convergence part of the complex analogue of Khintchine's
1879: theorem.  The much more difficult case of divergence requires deeper
1880: arguments and the reader is referred to Sullivan's bold and highly
1881: geometrical paper~\cite{Sullivan82}.
1882: 
1883: \section{Badly approximable complex numbers and Jarník's theorem}
1884: \label{sec:cba}
1885: The results of~\cite{BJ97} and of~\cite{FM95} on badly approximable
1886: points arising with the action of Kleinian groups and on bounded
1887: geodesics on Riemann surfaces respectively, specialised to the case of
1888: the Picard group, could be translated to give Jarník's theorem for
1889: complex numbers badly approximable by ratios of Gaussian integers.
1890: However, we will give a self-contained proof using an extension to the
1891: complex numbers of the $(\alpha,\beta)$-game introduced by
1892: W.~M.~Schmidt~\cite{Schmidt66}. In the setting of the complex plane,
1893: $S \subset \mathbb{C}$, $\alpha, \beta \in (0,1)$ and discs replace
1894: intervals. Thus, the game begins with player B choosing a disc $B_1 =
1895: \{ z \in \mathbb{C} : \abs{z-b_1} \leqslant \rho_1\}$. Next A chooses
1896: a disc $A_1 \subset B_1$ with radius $\alpha \rho_1$. Then B chooses a
1897: disc $B_2 \subset A_2$ with radius $\beta\alpha\rho_1$ and so on ad
1898: infinitum, such that $B_{n+1}$ has radius $(\alpha \beta)^n \rho_1$
1899: for any $n \geq 0$.
1900: 
1901: The discs $B_1, A_1, B_2, A_2, \dots$ form a nested decreasing
1902: sequence of closed sets, so there is a unique intersection point,
1903: $\omega$ say. Player A wins if this point is an element of the set
1904: $S$, \emph{i.e.}, if $\bigcap_j A_j = \{\omega\} \subset S$.
1905: Otherwise, B wins. A set $S$ is said to be $(\alpha, \beta)$-winning
1906: if A can win the game for the parameters $\alpha$ and $\beta$ no
1907: matter how well B plays. If for some $\alpha \in (0,1)$, A can win the
1908: game for any $\beta \in (0,1)$, the set $S$ is said to be
1909: $\alpha$-winning.
1910: 
1911: Player A benefits from $\alpha$ being small. Indeed, when $\alpha$
1912: gets smaller, A can limit the amount of choice B has in the next
1913: move. As in the real case, it may be shown that if $\alpha' < \alpha$
1914: and $S$ is $\alpha$-winning, then $S$ is $\alpha'$-winning.  Hence,
1915: given $S$, there is a largest value of $\alpha$ for which $S$ is
1916: $\alpha$-winning so we may define
1917: \begin{displaymath}
1918:   \alpha^*(S) := \sup \left\{\alpha \in (0,1) : S \text{ is
1919:       $\alpha$-winning} \right\}.
1920: \end{displaymath}
1921: 
1922: In fact, for any $S \neq \mathbb{C}$, we easily see that $\alpha^*(S)
1923: \leqslant \tfrac{1}{2}$. To see this, note that for the parameters
1924: $\alpha > \tfrac{1}{2}$ and $\beta \in (0, 2\alpha -1)$, B may ensure
1925: that the centres of all the $B_i$ are the same. Hence, by choosing the
1926: first disc $B_1$ with centre $b_1 \notin S$, we have the result.
1927: 
1928: When $S = \mathfrak{B}$, A wins if she can force $\omega$ to be badly
1929: approximable, \emph{i.e.}, if $\omega$ is in the set
1930: \begin{displaymath}
1931:   \mathfrak{B} = \left\{ z \in \mathbb{C} : \exists K > 0 \: \forall 
1932:     p,q \in
1933:     \mathbb{Z}[i]: \abs{z - \dfrac{p}{q}} > 
1934:     \dfrac{K}{\abs{q}^2}\right\}.
1935: \end{displaymath}
1936: Since $\mathfrak{B}$ is null, this seems unlikely to be the case, but in
1937: fact A can win the game whenever $2\alpha < 1 + \alpha\beta$.  It
1938: immediately follows that
1939: \begin{equation}
1940:   \label{seq:5}
1941:   \alpha^*(\mathfrak{B}) = 1/2,
1942: \end{equation}
1943: so in fact A may win this game almost as easily as she could win the
1944: game where $S$ is the entire complex plane with one point removed. We
1945: take some time to prove \eqref{seq:5}.
1946: 
1947: We have already seen that $\alpha^*(\mathfrak{B}) \leqslant 1/2$, so
1948: we will only consider $\alpha \leqslant 1/2$, as this will simplify
1949: the proof. Let $\alpha \in (0,1/2]$ and $\beta \in (0,1)$ be
1950: fixed. Note that $\gamma := 1 + \alpha \beta - 2 \alpha > 0$. We may
1951: assume without loss of generality that the radius of the initial ball
1952: $B_1$ satisfies $\rho_1 \leqslant
1953: \alpha\beta\tfrac{\gamma}{8}$. Indeed, otherwise we would let the game
1954: continue in an arbitrary fashion until reaching a $B_j$ for which
1955: $\rho_j \leqslant \alpha\beta\tfrac{\gamma}{8}$ and then take this to
1956: be our starting point. The constant $K$ in the definition of
1957: $\mathfrak{B}$ will be $\delta = \tfrac{\gamma}{4} \min(\rho, \alpha^2
1958: \beta^2 \tfrac{\gamma}{8})$, where $\rho = \rho_1$.
1959: 
1960: Let $t \in \mathbb{N}$ be such that $\alpha \beta \gamma \leqslant 2
1961: (\alpha\beta)^t < \gamma $. Let $R = (\alpha \beta)^{-t/2}$. Clearly,
1962: it suffices to prove for any $n \in \mathbb{N}$ that if
1963: \begin{subequations}
1964:   \begin{eqnarray}
1965:     \label{seq:3}
1966:     \gcd(p,q) = 1, \\ z \in B_{nt+1}, \\
1967:     \label{seq:4}
1968:     0 < \abs{q} < R^n,
1969:   \end{eqnarray}
1970: \end{subequations}
1971: then $\abs{z - \tfrac{p}{q}} > \tfrac{\delta}{\abs{q}^2}$.  This may
1972: be done using induction.
1973: 
1974: For $n = 0$, there is nothing to prove, as \eqref{seq:4} leaves us no
1975: $q$ to consider. Hence, we may assume that we have $B_1, \dots,
1976: B_{(k-1)t+1}$ such that the above holds for $0 \leqslant n \leqslant
1977: k-1$. In subsequent play, A thus only needs to worry about
1978: $\tfrac{p}{q}$ with $R^{k-1} \leqslant \abs{q} < R^k$, as the
1979: remaining problematic fractions have been sorted out in the preceding
1980: steps of the game.
1981: 
1982: In fact, there can be at most one such $\tfrac{p}{q}$. Indeed, suppose
1983: that there exists $p,p',q,q' \in \mathbb{Z}[i]$ and $z, z' \in B_{(k-1)t
1984:   + 1}$ with
1985: \begin{displaymath}
1986:   \abs{z - \dfrac{p}{q}} \leqslant \dfrac{\delta}{\abs{q}^2} \quad
1987:   \text{and} \quad \abs{z' - \dfrac{p'}{q'}} \leqslant 
1988:   \dfrac{\delta}{\abs{q'}^2}
1989: \end{displaymath}
1990: with $\abs{q}, \abs{q'} \in [R^{k-1}, R^k)$. Then
1991: \begin{multline*}
1992:   \abs{\dfrac{p}{q} - \dfrac{p'}{q'}} \leqslant \abs{z- 
1993:     \dfrac{p}{q}} +
1994:   \abs{z' - \dfrac{p'}{q'}} + \abs{z-z'} \\ \leqslant
1995:   \dfrac{\delta}{\abs{q}^2} + \dfrac{\delta}{\abs{q'}^2} +
1996:   2\rho(\alpha\beta)^{(k-1)t} \leqslant 2 \delta R^{2-2k} + 2 \rho 
1997:   R^{2-2k}
1998:   \\ \leqslant 4 \rho (\alpha\beta)^{-t} R^{-2k} \leqslant 4 \tfrac{1}{8} 
1999:   \alpha
2000:   \beta \gamma \dfrac{2}{\alpha \beta \gamma} R^{-2k} = 
2001:   R^{-2k}.
2002: \end{multline*}
2003: On the other hand, since $\gcd(p,q) = \gcd(p',q') = 1$ and since
2004: $\mathbb{Z}[i]$ is a unique factorisation domain,
2005: \begin{displaymath}
2006:   \abs{\dfrac{p}{q} - \dfrac{p'}{q'}} = \abs{\dfrac{pq' - 
2007:       p'q}{q q'}}
2008:   \geqslant \dfrac{1}{\abs{q}\abs{q'}} > R^{-2k}
2009: \end{displaymath}
2010: whenever $p\neq p'$ and $q \neq q'$. Hence, there can be at most one
2011: problematic point.
2012: 
2013: Note that we have used the property of the ring of Gaussian integers
2014: being a unique factorisation domain. For other rings, this may not be
2015: the case and stronger tools are needed for proving the analogous
2016: result. However, for clarity of this exposition, we will take the
2017: simple route and use the unique factorisation property. Note also
2018: that for the corresponding result in simultaneous Diophantine
2019: approximation, the module $\mathbb{Z}^2$ takes the place of the
2020: Gaussian integers. Even though the underlying sets are the same, the
2021: difference in algebraic structure prevents the method of this proof  
2022: from working in the case of simultaneous Diophantine approximation.
2023: 
2024: As there can be at most one point in $B_{(k-1)t+1}$ suitably close to
2025: a Gaussian rational $p/q$, we may devise strategies that avoids a disc
2026: around this point $p/q$ of a suitable radius. It is clear that we need
2027: to avoid $C=B(\tfrac{p}{q}, \tfrac{\delta}{\abs{q}^2})$. We examine
2028: two possibilities in turn.
2029: 
2030: \begin{figure}[htbp]
2031:   \centering 
2032:   \includegraphics[width=6cm,keepaspectratio]{easywin.eps}
2033:   \caption{A avoids $C$ in one move}
2034:   \label{fig:easywin}
2035: \end{figure}
2036: First, consider the case when $\abs{\tfrac{p}{q} - b_{(k-1)t+1}} >
2037: \delta R^{2-2k}$. In this case,
2038: \begin{displaymath}
2039:   \abs{\dfrac{p}{q} - b_{(k-1)t+1}} > \delta R^{2-2k} \geqslant
2040:   \dfrac{\delta}{\abs{q}^2}. 
2041: \end{displaymath}
2042: Hence, $b_{(k-1)t+1} \notin C$, so we are in the situation of figure
2043: \ref{fig:easywin}. As $\alpha \leqslant 1/2$, we see that A can choose her
2044: next disc in such a way that it does not intersect with $C$.
2045: 
2046: This leaves the final case when
2047: \begin{displaymath}
2048:   \abs{\dfrac{p}{q} - b_{(k-1)t+1}}\leqslant \delta R^{2-2k}.  
2049: \end{displaymath}
2050: This corresponds to the case when $b_{(k-1)t+1} \in C$ (figure
2051: \ref{fig:difficult}).
2052: \begin{figure}[htbp]
2053:   \centering
2054:   \includegraphics[width=6cm,keepaspectratio]{difficult.eps}
2055:   \caption{A avoids $C$ in $t$ steps}
2056:   \label{fig:difficult}
2057: \end{figure}
2058: In this case, we clearly need to work harder to get an answer, as the
2059: right strategy of A is not immediately obvious. However, it turns out
2060: that picking a fixed direction and moving as far as possible in this
2061: direction at each turn will cause the disc chosen by B after $t$ steps
2062: to have empty intersection with $C$. The following lemma formalises
2063: this.
2064: 
2065: \begin{lemma}
2066:   \label{lem:fundamental}
2067:   Let $\alpha, \beta \in (0,1)$ with $\gamma = 1 + \alpha\beta -
2068:   2\alpha >0$. Let $t \in \mathbb{N}$ be such that $(\alpha\beta)^t <
2069:   \gamma/2$. Suppose that the disc $B_k = (b_k, \rho_k)$ occurs at
2070:   some stage in an $(\alpha, \beta)$-game. Then A can play in such a
2071:   way that
2072:   \begin{displaymath}
2073:     B_{k+t} \subseteq \left\{z \in \mathbb{C} : \abs{z-b_k} > 
2074:       \rho_k \tfrac{\gamma}{2} \right\}.
2075:   \end{displaymath}
2076: \end{lemma}
2077: 
2078: \begin{proof}
2079:   We define a strategy for A in the following way: Suppose that the
2080:   last disc chosen by B was $B(c,\rho)$ for some $\rho > 0$. Choose
2081:   some $\hat{c} \in \mathbb{C}$ with  $\abs{\hat{c}} =
2082:   \rho(1-\alpha)$. We define a legal move for A by the map
2083:   \begin{displaymath}
2084:     B(c,\rho) \mapsto B(c+\hat{c}, \alpha \rho).
2085:   \end{displaymath}
2086:   We need to convince ourselves that this move is in fact a legal
2087:   move. First, we note that the radius is the right one. Hence, we
2088:   need only prove the inclusion $B(c+\hat{c}, \alpha \rho) \subseteq
2089:   B(c,\rho)$. But this follows since for any $z \in B(c+\hat{c},
2090:   \alpha \rho)$,
2091:   \begin{displaymath}
2092:     \abs{z - c} = \abs{z - \hat{c} + \hat{c} - c} \leqslant \abs{z -
2093:       (c + \hat{c})} + \abs{\hat{c}} \leqslant \alpha \rho + \rho
2094:     (1-\alpha) = \rho.
2095:   \end{displaymath}
2096:   The strategy of player A will be to use the above move, no matter
2097:   how B plays the game. 
2098:   
2099:   We denote the discs chosen by B by $B_k = (b_k, \rho_k)$, and the
2100:   discs chosen by A by $A_k = (a_k, \alpha \rho_k)$. Note that
2101:   \begin{equation}
2102:     \label{seq:1}
2103:     \abs{a_k - b_k} = \abs{b_k + \hat{c} - b_k} = 
2104:     \rho_k(1-\alpha).
2105:   \end{equation}
2106:   \begin{figure}[htbp]
2107:     \centering 
2108:     \includegraphics[width=6cm,keepaspectratio]{figure.eps}
2109:     \caption{The point $b_{k+1}$ must be chosen in the shaded 
2110:       area.}
2111:     \label{fig:Blacks_choice}
2112:   \end{figure}  
2113:   Also, from figure \ref{fig:Blacks_choice}, we see that
2114:   \begin{equation}
2115:     \label{seq:2}
2116:     \abs{b_{k+1} - a_k} \leqslant \alpha \rho_k - \alpha \beta \rho_k.
2117:   \end{equation}
2118:   Hence, by \eqref{seq:1} and \eqref{seq:2},
2119:   \begin{multline*}
2120:     \abs{b_k - b_{k+1}} = \abs{(b_k - a_k) - (b_{k+1} - a_k)}
2121:     \geqslant \big\vert\abs{b_k - a_k} - \abs{b_{k+1} - a_k}
2122:     \big\vert\\ 
2123:     = \big\vert \rho_k (1-\alpha) - \abs{b_{k+1} - a_k}\big\vert >
2124:     \abs{\rho_k (1-\alpha) - \alpha \rho_k + \alpha \beta \rho_k} =
2125:     \rho_k \gamma. 
2126:   \end{multline*}
2127:   Continuing as above, at each step choosing $\hat{c}$ to point in the
2128:   direction of the first one chosen, we obtain,
2129:   \begin{displaymath}
2130:     \abs{b_{k+t} - b_k} > \rho_k \gamma.
2131:   \end{displaymath}
2132:   But since $\rho_{k+t} = (\alpha \beta)^t \rho_k < \rho_k
2133:   \tfrac{\gamma}{2}$, we have for any $z \in B_{k+t}$,
2134:   \begin{displaymath}
2135:     \abs{z - b_k} \geqslant \big\vert\abs{z - b_{k+t}} - \abs{b_{k+t}
2136:       -  b_k}\big\vert > \rho_k \tfrac{\gamma}{2}.
2137:   \end{displaymath}
2138:   This completes the proof.
2139: \end{proof}
2140: 
2141: With the induction step done, we have shown that A can play in such a
2142: way that \eqref{seq:5} holds for the final intersection point.  In
2143: fact, we have found an explicit lower bound on the constant $K$ for
2144: which \eqref{seq:5} holds.
2145: 
2146: From \eqref{seq:5}, we can get the Hausdorff dimension of the set
2147: $\mathfrak{B}$. By considering the amount of choice B has in the game
2148: and using this to find a lower bound on the $s$-length of appropriate
2149: covers, we obtain a lower bound on the dimension of any $(\alpha,
2150: \beta)$-winning set. This construction was carried out in considerable
2151: generality by W. M. Schmidt \cite{Schmidt66}, who obtained a lower
2152: bound for the Hausdorff dimension of $(\alpha,\beta)$-winning sets in
2153: real Hilbert spaces. We sketch an approach to obtaining the dimension
2154: from the number $\alpha^*(\mathfrak{B})$. Because of the geometric
2155: nature of the problem, $\mathfrak{B}$ will be regarded as a subset in
2156: $\mathbb{R}^2$. We will compute the usual planar Hausdorff dimension
2157: here and subsequently.
2158: 
2159: To obtain the Hausdorff dimension of $\mathfrak{B}$, we consider the
2160: game from B's point of view. Assume first without loss of generality
2161: that $B_0$, the first disc chosen, has radius $1$. At any point of the
2162: game, B can choose to direct the game into a number of disjoint
2163: discs. While there may be a variety of ways in which these discs can
2164: be chosen, the maximum number $N(\beta)$ is roughly equal to
2165: $1/\beta^2$, \emph{i.e.}, $N(\beta) \asymp 1/\beta^2$.
2166: 
2167: We limit B's choice to these $N(\beta)$ possible moves and assume that
2168: A is playing to win the game. This gives us a parametrisation of the
2169: sequence of discs chosen by B, so that $B_k=B_k(j_1, \dots, j_k)$ with
2170: the $j_i \in \{0, \dots, N(\beta)-1\}$ for $i = 1, \dots, k$. For
2171: later use, note that the radius of each these discs is $\rho_k =
2172: (\alpha\beta)^k$. By simultaneously considering all the different
2173: ways, B may play the game, we obtain a function
2174: \begin{displaymath}
2175:   f: \left\{0, \dots, N(\beta)-1\right\}^{\mathbb{N}} \rightarrow
2176:   \mathfrak{B}, \quad (\lambda_k)_{k \in \mathbb{N}} \mapsto
2177:   \bigcap_{k \in \mathbb{N}} B_k(\lambda_1, \dots, \lambda_k) =
2178:   \{x(\lambda)\}.
2179: \end{displaymath}
2180: We define the set $\mathfrak{B}^* \subseteq \mathfrak{B}$ to be the
2181: range of $f$. As every number in the interval $[0,1]$ has at least one
2182: expansion in base $N(\beta)$, we may map this set onto the unit
2183: interval by the map
2184: \begin{displaymath}
2185:   g: \mathfrak{B}^* \rightarrow [0,1], \quad x(\lambda) \mapsto
2186:   0.\lambda_1 \lambda_2 \dots.
2187: \end{displaymath}
2188: Note that these functions could well be multivalued, but this is of no
2189: concern to us. All we need is a cover of $[0,1]$. We extend this
2190: function to subsets of the complex plane by defining $g(Z) = g(Z \cap
2191: \mathfrak{B}^*)$ for any $Z \subseteq \mathbb{C}$, where by convention
2192: $B(\emptyset) = 0$.
2193: 
2194: Now, take some cover $\mathcal{C} = \{C_l\}_{l \in \mathbb{N}}$ of
2195: $\mathfrak{B}$ with discs of radius $\rho(C_l) = \rho_l$. We wish to
2196: find a lower bound on the $s$-length of this cover for appropriate $s$
2197: as $\rho_l$ becomes smaller. This is where the function $g$ comes in
2198: handy. As $\mathcal{C}$ covers $\mathfrak{B}^*$, we see that
2199: $g(\mathcal{C})$ covers $[0,1]$. We let $\overline{\mu}$ denote the
2200: outer Lebesgue measure. By sub-additivity,
2201: \begin{equation}
2202:   \label{seq:10}
2203:   \sum_{l=1}^\infty \overline{\mu} (g(C_l)) \geqslant
2204:   \overline{\mu}\left(\bigcup_{l=1}^\infty g(C_l)\right) \geqslant 1.
2205: \end{equation}
2206: 
2207: Now let $\omega>0$ be sufficiently small so that any disc of radius
2208: $\omega(\alpha\beta)^k$ intersects at most two of the discs $B_k(j_1,
2209: \dots, j_k)$. By \cite[Lemma 20]{Schmidt66}, $\omega = 2/\sqrt{3} -1 <
2210: 1$ has this property. We define integers
2211: \begin{displaymath}
2212:   k_l = \left[\dfrac{\log \left(2 \omega^{-1}
2213:         \rho_l\right)}{\log(\alpha\beta)}\right].
2214: \end{displaymath}
2215: 
2216: For $\rho_l$ sufficiently small, we see that $k_l > 0$ and $\rho_l <
2217: \omega (\alpha\beta)^{k_l}$. Hence, the disc $C_{l}$ intersects at
2218: most two of the discs $B_{k_l}(j_1, \dots, j_{k_l})$. As
2219: $g(B_{k_l}(j_1, \dots, j_{k_l}))$ is clearly an interval of length
2220: $N(\beta)^{-k_l}$, we have $\overline{\mu} (g(C_{l})) \leqslant 2
2221: N(\beta)^{k_l}$, so by \eqref{seq:10},
2222: \begin{displaymath}
2223:   1 \leqslant \sum_{l=1}^\infty \overline{\mu} (g(C_l)) \leqslant
2224:   \sum_{l=1}^\infty 2 N(\beta)^{k_l} \leqslant
2225:   2(2\omega^{-1})^{\tfrac{\log(N(\beta))}{\abs{\log(\alpha\beta)}}}
2226:   \sum_{l=1}^\infty
2227:   \rho_l^{\tfrac{\log(N(\beta))}{\abs{\log(\alpha\beta)}}}.
2228: \end{displaymath}
2229: Thus, for $s = \log(N(\beta))/\abs{\log(\alpha\beta)}$, the $s$-length
2230: of the cover $\mathcal{C}$ is strictly positive, so the Hausdorff
2231: dimension of $\mathfrak{B}^*$ must be greater than this number.
2232: 
2233: Now, we fix $\alpha \in (0,1/2)$ and apply the above,
2234: \begin{displaymath}
2235:   \hdim(\mathfrak{B}) \geqslant \dfrac{\log \beta^{-2}}{\abs{\log
2236:       \alpha \beta}} = \dfrac{2\abs{\log \beta}}{\abs{\log
2237:       \alpha} + \abs{\log \beta}} \rightarrow 
2238:       2 
2239: \end{displaymath}
2240: as $\beta \to 0$. We have thus proved the analogue of Jarník's theorem
2241: for the complex numbers:
2242: \begin{theorem}
2243:   \label{thm:jarnik}
2244:   The set $\mathfrak{B}$ is thick, \emph{i.e.}, for any disc $B \subseteq
2245:   \mathbb{C}$,
2246:   \begin{displaymath}
2247:     \hdim(B \cap \mathfrak{B}) = 2.
2248:   \end{displaymath}
2249: \end{theorem}
2250: 
2251: \section{The complex Jarník--Besicovitch theorem}
2252: \label{sec:cJBt} 
2253: Let $W^*_v$ be the set of $v$-approximable complex numbers which
2254: satisfy the inequality~\eqref{eq:vapprox} for infinitely many Gaussian
2255: rationals $p/q$ (recall that $p,q \in \mathbb{Z}[i]$).  The general
2256: Jarn\'ik--Besicovitch theorems in~\cite{HVJBGF,MP93} can be
2257: specialised to  the Picard group to yield the Hausdorff dimension of
2258: $W^*_v$ (see~\cite[Corollary~2]{MP93}) but we give a more
2259: self-contained and direct proof. For convenience we consider
2260: \begin{equation*}
2261:   \label{eq:Vv}
2262:   V^*_v:=W^*_v\cap I^2=\left\{z\in I^2\colon \left|z-\frac{p}{q}\right|
2263:     <\frac1{|q|^{v+1}} \ {\text{ for
2264:         infinitely many }} \frac{p}{q} \right\}.
2265: \end{equation*}
2266: \begin{theorem}
2267:   \label{thm:complexJBt}
2268:   \begin{equation*}
2269:     \hdim V^*_v=\hdim W^*_v =
2270:     \begin{cases}
2271:       \dfrac{4}{v+1} & \text{ when } v\geqslant 1, \\
2272:       2  & \text{ when } v\leqslant 1.
2273:     \end{cases}
2274:   \end{equation*}
2275: \end{theorem}
2276: When $v\leqslant 1$, $V^*_v$ is full by the complex form of
2277: Khintchine's theorem so the second equality holds by (iii) in
2278: \S~\ref{sec:HD}. As usual, the proof for $v > 1$ falls into two
2279: parts. First, to obtain the upper bound for $\hdim V^*_v$, consider
2280: the cover ${\scr{C}}(V^*_v)$ given by~\eqref{eq:cover} with $N=1$ and
2281: $\Psi(k)=k^{-v-1}$. By \eqref{eq:7}, this has $s$-length
2282: \begin{eqnarray*}
2283:   \ell^s(\scr{C}(V^*_v)) & \ll & \sum_{k=1}^\infty \sum_{k \leqslant 
2284:     \abs{q} < k+1} \abs{q}^2 \left(\abs{q}^{-v-1}\right)^s \ll
2285:   \sum_{k=1}^\infty k^{2-(v+1)s} \sum_{k \leq \abs{q} < k+1} 1
2286:   \\
2287:   &\ll& \sum_{k=1}^\infty k^{3-s(v+1)} <\infty
2288: \end{eqnarray*}
2289: for $s>4/(v+1)$. It follows that when $v\geqslant 1$, 
2290: \begin{equation}
2291:   \label{eq:dimle}
2292:   \hdim V^*_v\leqslant \dfrac{4}{v+1}.  
2293: \end{equation}
2294: 
2295: To obtain the lower bound in the Jarník--Besicovitch theorem, we use
2296: ubiquity \cite{MDAM, DRV90a}. Let $S\subseteq I^2$ and let $\rho\colon
2297: \mathbb{N}\to (0,\infty)$ be a function.  Put
2298: \begin{equation*}
2299:   B(S;q,\varepsilon):=  
2300:   \left\{z\in S \colon \left\vert z-\dfrac{p}{q}\right\vert <\varepsilon
2301:     \ \text{ for some } \ p \in \mathbb{Z}[i]\right\}.
2302: \end{equation*}
2303: The set 
2304: \begin{equation*}
2305:   \mathcal{R}=\bigcup_{q} R(q)\subset \mathbb{C} ,  
2306: \end{equation*}
2307: where the union is over non-zero $q\in \mathbb{Z}[i]$, consists of
2308: discrete points and so has dimension 0.  Let $S$ be any open square in
2309: $I^2$.  By definition, $\mathcal{R}$ is \emph{ubiquitous} in
2310: $S\subseteq I^2 =[0,1)^2$ with respect to $\rho$ if
2311: \begin{equation*}
2312:   \left\vert \bigcup_{q} B(S;q,\rho(N))\right\vert  \to \vert S\vert 
2313: \end{equation*}
2314: as $N\to \infty$. Now by the complex analogue of Dirichlet's theorem,
2315: for each $N\in\mathbb{N}$,
2316: \begin{eqnarray*}
2317:   \left\{z\in S \colon \left\vert z-\dfrac{p}{q}\right\vert
2318:     <\dfrac{2}{\vert q\vert N}\ \text{ for some }
2319:     p,q\in\mathbb{Z}[i],1\leqslant\vert q\vert \leqslant N
2320:   \right\} \\
2321:  \qquad \qquad  =\bigcup_{q} B\left(q,\dfrac{2}{\vert q\vert N
2322:     }\right) = S,\qquad\qquad\qquad\qquad\qquad &
2323: \end{eqnarray*}
2324: so that 
2325: \begin{equation*}
2326:   \left\vert \bigcup_{q} B\left(q,\dfrac{2}{\vert q\vert N
2327:       }\right)\right\vert =\vert S\vert.
2328: \end{equation*}
2329: Consider the set $S(N)$ of $z\in S$ with `small denominators':
2330: \begin{alignat*}{2}
2331:   S(N) &= \left\{z\in S\colon \text{ there exist } p, q \text{ such 
2332:       that } \left\vert z-\dfrac{p}{q}\right\vert <\dfrac{2}{\vert
2333:       q\vert N}, 1\leqslant \vert q\vert <\dfrac{N}{\log N}\right\} \\ 
2334:   &= \bigcup_{1\leqslant \vert q\vert < N/\log 
2335:     N} B\left(q,\dfrac{2}{\vert q\vert N}\right).
2336: \end{alignat*}
2337: The measure $\vert S(N)\vert $ of $S(N)$ satisfies
2338: \begin{eqnarray*}
2339:   \abs{S(N)} &=&\sum_{1\leqslant \vert q\vert < N/\log N} \vert
2340:   B(q,2/\vert q\vert N)\vert  \ll \sum_{1\leqslant \vert q\vert <
2341:     N/\log N} (\vert q\vert N)^{-2}\vert q\vert ^2\\ 
2342:   &\ll & N^{-2}\sum_{1\leqslant \vert q\vert < N/\log N} 1\ll N^{-2} 
2343:   \left(N/\log N\right)^2 \ll (\log N)^{-2}\to 0
2344: \end{eqnarray*}
2345: as $N\to \infty$.  Choose $\rho(N)=2N^{-2}\log N$.  Then since
2346: $\vert q\vert >N/\log N$ implies $\rho(N)>2/\vert q\vert N$,
2347: \begin{equation*}
2348:   \bigcup_{N/\log N < \vert q\vert \leqslant N} B(q,\rho(N))\supset 
2349:   \bigcup_{N/\log N < \vert q\vert \leqslant N} B\left(q,\dfrac{2}{\vert
2350:     q\vert N}\right) \to S 
2351: \end{equation*}
2352: in measure as $N\to\infty$ and so $\mathcal{R}$ is ubiquitous with
2353: respect to $\rho(N)= 2\log N /N^2$ for any $S\subseteq I^2$ and so for
2354: $S=I^2$. But since $\Psi$ is decreasing, by~\cite{DRV90a},
2355: \begin{equation}
2356:   \label{eq:dimge}
2357:   \hdim V^*_v \geqslant \dim \mathcal{R} + \codim \mathcal{R}
2358:   \limsup_N \dfrac{\log \rho(N)}{\log \Psi (N)} = \dfrac{4}{v+1}, 
2359: \end{equation}
2360: where $\dim$ is the topological dimension, so that $\dim \mathcal{R}
2361: =0$, the codimension $\codim \mathcal{R}$ in $\mathbb{C}$ (regarded as
2362: $\mathbb{R}^2$) of $\mathcal{R}$ is $2$ and $\Psi(N)=N^{-v-1}$.
2363:                    
2364: The required result follows on combining the two complementary
2365: inequalities~\eqref{eq:dimle} and~\eqref{eq:dimge}.
2366: 
2367: In fact, since $S$ was an arbitrary open square, $\mathcal{R}$ is
2368: \emph{locally} ubiquitous and it has been shown that the local
2369: ubiquity of $\mathcal{R}$ also implies the divergence case of
2370: Khintchine's theorem~\cite{BDV03}.
2371: 
2372: Note that even though one might expect Diophantine approximation in
2373: $\mathbb{C}$ with respect to Gaussian integers to be similar to
2374: simultaneous Diophantine approximation in the real plane
2375: $\mathbb{R}^2$, this is not at all the case. The analogues of
2376: Dirichlet's theorem, Khintchine's theorem and the
2377: Jarn\'ik--Besicovitch theorem are quite different in the two cases.
2378: Indeed the complex Dirichlet's Theorem and the Jarn\'ik--Besicovitch
2379: theorem are closer to the real, one-dimensional case. Only the
2380: analogue of Jarn\'ik's theorem on badly approximable numbers remains
2381: unchanged and in this case there is a substantial difference in the
2382: proofs of the two theorems.
2383: 
2384: \section{Applications}
2385: \label{sec:applns}
2386: 
2387: The connection between the physical phenomenon of resonance and
2388: Diophantine equations can give rise to the notorious problem of small
2389: denominators in which solutions to a variety of questions contain
2390: denominators that can become arbitrarily small.  When these small
2391: denominators are related to very well approximable points, it is
2392: sometimes possible to impose appropriate Diophantine conditions which
2393: overcome the problem by excluding the offending denominators without
2394: significantly affecting the validity of the solution.  The techniques
2395: developed in the metrical theory of Diophantine approximation lend
2396: themselves to this and in particular the Jarník--Besicovitch theorem
2397: allows the determination of the Hausdorff dimension of the associated 
2398: exceptional sets.  Some examples of problems involving small
2399: denominators and the associated exceptional sets are now discussed.
2400: We begin with a very simple example.
2401: 
2402: \subsection{Partial differential equations}
2403: \label{sec:pdes}
2404: 
2405: Diophantine approximation has been applied to the wave equation
2406: (\cite{MR50:7828} and more recently \cite{MR95c:35030}), as well as to
2407: the Schrödinger equation \cite{kristensen:_dioph_schroed}. For an
2408: extensive treatment of Diophantine problems related to partial
2409: differential equations, the reader is referred to \cite{Ptashnik84}.
2410: 
2411: We will illustrate the applications of Diophantine approximation by
2412: Gaussian rationals by considering the following innocuous first-order
2413: linear complex partial differential equation,
2414: \begin{equation}
2415:   \label{eq:6}
2416:   \alpha \dfrac{\partial f(z,t)}{\partial t} + \beta
2417:   \dfrac{\partial f(z,t)}{\partial z} = g(z,t),  
2418: \end{equation}
2419: where $z \in \{x+iy \in \mathbb{C} : x,y \geqslant 0\}, t \geqslant 0$
2420: and $\alpha,\beta$ are non-zero complex numbers. That is, we are
2421: studying the partial differential equation on the interior of the set
2422: defined above under the additional assumption that the functions
2423: involved as well as all their derivatives may be extended to the whole
2424: set. Assume that $g(z,t)$ is smooth (\emph{i.e.}, $C^\infty$) and can
2425: be expressed in the form
2426: \begin{displaymath}
2427:   g(z,t) = \sum_{a,b,c,d \in \mathbb{Z}} g_{a,b,c,d} \exp\left((a+ib)z +
2428:     (c+id)t\right), \quad g_{a,b,d,c} \in \mathbb{C}.
2429: \end{displaymath}
2430: We seek smooth solutions to this equation of the same form, namely
2431: \begin{equation}
2432:   \label{eq:fzt}
2433:   f(z,t) = \sum_{a,b,c,d \in \mathbb{Z}} f_{a,b,c,d}
2434:   \exp\left((a+ib)z + (c+id)t\right), \quad f_{a,b,d,c}
2435:   \in \mathbb{C}.
2436: \end{equation}
2437: Thus, we are not just looking for solutions but rather trying to solve
2438: the partial differential equation subject to boundary conditions. 
2439: 
2440: As usual, we solve the problem formally by substituting these two
2441: expressions into~\eqref{eq:6} and identifying coefficients on either
2442: side of the equality. Isolating the coefficients of $f$, we get
2443: \begin{equation}
2444:   \label{eq:fabcd}
2445:   f_{a,b,c,d} = \dfrac{\frac{1}{\alpha}}{\frac{\beta}{\alpha}(a+ib)
2446:     + (c+id)} g_{a,b,c,d}. 
2447: \end{equation}
2448: We need the coefficients to decay fast enough so that both the series
2449: \eqref{eq:fzt} and its derivatives are convergent. Since $g$ is
2450: already smooth, the coefficients $g_{a,b,c,d}$ decay rapidly and so
2451: are not obstructing this convergence. However, the denominator of the
2452: fraction may become small and cause the the coefficients $f_{a,b,c,d}$
2453: to become large enough to pose a problem. In order to avoid this, we
2454: see that it is certainly sufficient for the denominator to be bounded
2455: from below by some polynomial in $a,b,c,d$, \emph{i.e.}, we require
2456: for some $K,v > 0$ and for all $(a,b,c,d) \in \mathbb{Z}^4 \setminus
2457: \{0\}$, 
2458: \begin{equation}
2459:   \label{eq:9}
2460:   \abs{\dfrac{\beta}{\alpha}(a+ib) + (c+id)} \geqslant K
2461:   \max\left\{\abs{a},\abs{b}, \abs{c},\abs{d}\right\}^{-v}.
2462: \end{equation}
2463: Since we are only concerned with small denominators, we can assume
2464: without loss of generality that $\abs{a+ib} \asymp \abs{c+id}$, so
2465: that after adjusting $K$ we can drop the dependence on $c$, $d$ on the
2466: right-hand side of \eqref{eq:9} and require
2467: \begin{displaymath}
2468:   \left\vert \dfrac{\beta}{\alpha}-\dfrac{p}{q}\right\vert \geqslant 
2469:   \dfrac{K}{\vert q\vert ^{v+1}},
2470: \end{displaymath}
2471: where $p=a+ib$ and $q=c+id$. Thus we require $\beta/\alpha$ to be of
2472: complex Diophantine type $(K,v)$ for some $K,v$.  The complement of
2473: this set is $ E^*=\cap_{v>1} E_v^*$, where
2474: \begin{equation*}
2475:   E_v^*=\left\{z\in\mathbb{C}\colon  {\text{ for any }} \ K>0, 
2476:     \left\vert z-\dfrac{p}{q}\right\vert  <
2477:     \dfrac{K}{\vert q\vert^{v+1}}  {\text{ for some }} \
2478:     p, q \in\mathbb{Z}[i] \right\}.
2479: \end{equation*}
2480: But it can be readily verified by an argument similar to that giving
2481: the inclusion~\eqref{eq:UEWincl}, that for each $\varepsilon>0$,
2482: \begin{equation*}
2483:   W^*_{v+\varepsilon}\subset E_v^* \subset W^*_v,
2484: \end{equation*}
2485: whence by the properties of Hausdorff dimension given in
2486: \S\ref{sec:HD} and the Jarník--Besicovitch theorem for Gaussian
2487: rational approximation (Theorem~\ref{thm:complexJBt}),
2488: \begin{displaymath}
2489:   \hdim E_v^*=\dfrac{4}{v+1}
2490: \end{displaymath}
2491: for $v\geqslant 1$ and so $\hdim E^*=\lim_{v\to\infty} \hdim E_v^*=0$.
2492: Thus the exceptional set associated with the inequality~\eqref{eq:9}
2493: failing to hold has Hausdorff dimension zero.
2494: 
2495: \subsection{The rotation number}
2496: \label{sec:rotno}
2497: 
2498: The rotation number $\rho(f)$ is a measure of how far `on average' a
2499: continuous, orientation preserving homeomorphism $f\colon
2500: \mathbb{S}^1\to \mathbb{S}^1$ moves a point round the circle. We will
2501: not give the fairly lengthy definition which is explained
2502: in~\cite{MDAM,mmdnewton,KatokHass,NiteckiDD} but content ourselves
2503: with the observation that the rotation number of a rotation $r_\alpha$
2504: by an angle $2\pi\alpha$, where $0\leqslant \alpha <1$, given by
2505: \begin{equation*}
2506:   r_\alpha(z)= ze^{2\pi i \alpha}  
2507: \end{equation*}
2508: is, naturally enough, $\alpha$. The rotation number is a nice example
2509: of how Diophantine properties can arise in analysis as it can be shown
2510: that $\rho(f)$ is irrational if and only if $f$ has no periodic points
2511: (see~\cite[Chap.~11,12]{KatokHass} or~\cite[Chap.~1]{NiteckiDD} for
2512: more details).
2513: 
2514: If $\rho(f)$ is irrational then for $z\in \mathbb{S}^1$, the closure
2515: $A$ of the orbit
2516: \begin{displaymath}
2517:   \omega(z)=\{f^n(z)\colon n\in\mathbb{N}\}
2518: \end{displaymath} 
2519: does not depend on $z$ and either $A$ is perfect and nowhere dense or
2520: $A=\mathbb{S}^1$.  In the latter case $f$ is transitive and is
2521: topologically conjugate to the rotation $r_{\rho(f)}\,$ by the
2522: rotation number $\rho(f)$ of $f$, \emph{i.e.}, there exists an
2523: orientation preserving homeomorphism $\varphi\colon \mathbb{S}^1\to
2524: \mathbb{S}^1$ such that
2525: \begin{equation*}
2526:   f=\varphi^{-1}\circ r_{\rho(f)} \circ \varphi, 
2527: \end{equation*}
2528: usually written $f\sim r_{\rho(f)}$.  This can be regarded as
2529: obtaining a normal form for $f$ and is analogous to 
2530: diagonalising a matrix. 
2531: 
2532: Denjoy showed that when $f$ is $C^2$ and $\rho(f)$ is irrational, then
2533: $f$ is topologically conjugate to the rotation by $\rho(f)$. More
2534: subtle aspects arise when additional differentiability conditions are
2535: imposed on the conjugation. For example, every $C^\infty$
2536: diffeomorphism $f$ of the the circle is $C^\infty$ conjugate to a
2537: rotation if and only if the rotation number $\rho(f)$ of $f$ is of
2538: Diophantine type~\cite{Yoccoz84}.  In the analytic case, the rotation
2539: numbers of real analytic diffeomorphisms form a set lying strictly
2540: between $\mathcal{D}$ and the set of Bruno
2541: numbers~\cite[p.~92]{Yoccoz95a}.  As in the preceding example, the
2542: Diophantine condition arises from the denominator $1-e^{2\pi i \rho
2543: k}$ in the coefficients for a Fourier series solution of a linearised
2544: auxiliary equation in an iterative Newton's tangent method argument,
2545: modified at each step to retain convergence. In order to guarantee
2546: convergence of the iterative argument and of the Fourier series, the
2547: inequalities
2548: \begin{equation*}
2549:   \vert 1-e^{2\pi i  \rho k}\vert 
2550:   \geqslant 2\left\vert \sin\left(\dfrac{\rho k-j}{2}\right)\right\vert 
2551:   \geqslant \dfrac{2\vert \rho k-j\vert}{\pi},
2552: \end{equation*}
2553: where $\rho k-j\in [0,2\pi)$, must be set against the very rapid decay
2554: of the corresponding Fourier coefficient $f_k$ ($\ll k^{-N}$ for any
2555: $N>0$) in the numerator.  It suffices that $\rho$ is of Diophantine
2556: type $(K,v)$ for some $K>0$, $v>1$, since then
2557: \begin{equation*}
2558:   \left\vert k\rho-j\right\vert \geqslant \dfrac{K}{k^{v}}, \
2559:   \emph{i.e.}, \left\vert \rho-\dfrac{j}{k}\right\vert \geqslant
2560:   \dfrac{K}{k^{v+1}}, 
2561: \end{equation*}
2562: for all $j/k\in\mathbb{Q}$.
2563: 
2564: Now we saw in~\S\ref{sec:Dt} that when $v>1$, almost all real numbers
2565: are of Diophantine type $(K,v)$ for some positive $K$, \emph{i.e.},
2566: the set
2567: \begin{equation*}
2568:   \mathcal{D}_v = \bigcup_{K>0} \{\alpha\in\mathbb{R}\colon \vert
2569:   \alpha-p/q\vert \geqslant Kq^{-1-v} \text{ for each }
2570:   p/q\in\mathbb{Q}\}  
2571: \end{equation*}
2572: is of full Lebesgue measure. Thus the complementary set $E_v$
2573: \begin{equation*}
2574:   E_{v}= \bigcap_{K>0} \{\alpha\in\mathbb{R}\colon \vert
2575:   \alpha-p/q\vert < Kq^{-2-v} {\text{ for some }} p/q\in\mathbb{Q}\}
2576: \end{equation*}
2577: is null for $v>1$ (see \S\ref{sec:Dt}). As in the preceding section,
2578: its Hausdorff dimension can be determined using the
2579: inclusions~\eqref{eq:UEWincl} and the Jarník--Besicovitch theorem
2580: (Theorem~\ref{thm:JBthm})
2581: \begin{equation*}
2582:   \hdim E_v=\hdim W_{v}=\dfrac{2}{v+1} 
2583: \end{equation*}
2584: for $v\geqslant 1$.  If the rotation number of the smooth circle
2585: function $f$ does not lie in $E=\bigcap_{v>1} E_v=\lim_{v\to\infty}
2586: E_v$ ($E_v$ decreases as $v$ increases), then $f$ is smoothly
2587: conjugate to a rotation. The Hausdorff dimension of the exceptional
2588: set is
2589: \begin{equation}
2590:   \label{eq:dimE0}
2591:   \hdim E= \lim_{v\to\infty} \dfrac{2}{v+1} 
2592:   = 0,
2593: \end{equation}
2594: \emph{i.e.}, the complement of $\mathcal{D}$ has Hausdorff dimension
2595: 0.
2596: 
2597: \subsection{The structure of Julia and Fatou sets}
2598: \label{sec:struct-julia-fatou}
2599: 
2600: Let $R(z) = P(z)/Q(z)$ be a rational map on the Riemann sphere
2601: $\mathbb{C}_\infty$. A famous result due to Sullivan
2602: \cite{MR99e:58145, MR87i:58103, MR87i:58104} (see also
2603: \cite{MR88e:00019}) states that the Fatou set $F_R = \mathbb{C}_\infty
2604: \setminus J_R$ of such a map has countably many periodic connected
2605: components. These connected components were further classified
2606: according to the type of periodic cycles they are associated with. As
2607: the Julia set is the complementary set of the Fatou set, this
2608: classification also deals with the structure of the Julia set on which
2609: the dynamics of $Q$ is chaotic.
2610: 
2611: Cycles may be classified according to the value of $dR/dz$ on the
2612: points of the cycle. By the chain rule, this value remains constant,
2613: $\lambda$ say. The cycle is attracting (resp.  repelling) as
2614: $\abs{\lambda} < 1$ (resp. $\abs{\lambda} > 1$). Repelling cycles are
2615: part of the Julia set $J_R$, so no parts of the Fatou set corresponds
2616: to this case. For attracting cycles, an associated periodic connected
2617: component of the Fatou set is in fact the immediate basin of
2618: attraction of this cycle, \emph{i.e.} the union of the connected
2619: components of the Fatou set containing the points of the cycle in
2620: question. 
2621: 
2622: When $\abs{\lambda}=1$, Diophantine properties of $\lambda$ determine
2623: the behaviour of the dynamics of $R$ and hence the structure of the
2624: Julia and Fatou sets. In this case, $\lambda = \exp(2 \pi i
2625: \alpha)$. When $\alpha$ is rational, the cycle is said to be
2626: parabolic and the associated component of the Fatou set is again the
2627: immediate basin of attraction.
2628: 
2629: In the case when $\alpha$ is irrational, one is interested in the
2630: situation when the domain of the Fatou set corresponding to the
2631: periodic point is a collection of Siegel discs associated with the
2632: cycle. That is, we are looking for sets on which the dynamics are
2633: topologically conjugate to a rotation of a disc. Such rotation numbers
2634: were studied in the preceding section, where the exceptional set
2635: associated with the failure of this condition was shown to have
2636: Hausdorff dimension zero.
2637: 
2638: \subsection{Linearising diffeomorphisms}
2639: \label{sec:lindiffeos}
2640: Suppose the complex analytic diffeomorphism $f:\mathbb{C}^{\,n}\to
2641: \mathbb{C}^{\,n}$ has a fixed point.  Without loss of generality, this
2642: can be taken to be the origin, so that $f(0)=0$.  If in a
2643: neighbourhood of $0$, $f$ is analytically (biholomorphically)
2644: conjugate to its linear part or Jacobian $Df\vert_0=A$ say, the function
2645: $f$ is said to be linearisable.  The linearising transformation $\phi$
2646: is given by the solution to the functional equation
2647: \begin{equation*}
2648:   f=\phi^{-1}\circ A  \circ  \phi,
2649: \end{equation*}
2650: known as Schr\"oder's equation when $n=1$.  Thus linearisation is
2651: similar to conjugating a circle map to a rotation, discussed above
2652: in~\S\ref{sec:rotno}.  And problems of small denominators arise when
2653: the eigenvalues $\alpha_1,\dots,\alpha_n$ of $A$ are close to being
2654: resonant in the sense that they are close to satisfying the equation
2655: \begin{equation*}
2656:   \alpha_k=\prod_{r=1}^n \alpha_r^{j_r}
2657: \end{equation*}
2658: for all $\mathbf{j}=(j_1,\dots,j_n)$ with $j_r \in \mathbb{N} \cup
2659: \{0\}$, $r=1,\dots,n$ and $\vert \mathbf{j}\vert_1=\sum_r \vert
2660: j_r\vert  \geqslant 2$. Linearisation is well understood when $n=1$
2661: and the diffeomorphism $f:\mathbb{C}\to \mathbb{C}$ with $f(0)=0$ can
2662: be linearised when $\vert (Df\vert_0)\vert =\vert f'(0)\vert  \not=1$.
2663: The interesting case when $\vert f'(0) \vert =1$ is closely related
2664: via lifts to the conjugacy of a circle map to a rotation, discussed
2665: in~\S\ref{sec:rotno} above, and necessary and sufficient conditions
2666: for the linearisation of a diffeomorphism
2667: $f\colon\mathbb{C}\to\mathbb{C}$ are known~\cite{Yoccoz95a}.  On the
2668: other hand, when $n \geqslant 2$, the problem of finding which 
2669: functions can be linearised is very difficult but Siegel's normal
2670: forms theorem~\cite{Siegel42,Siegel52} gives sufficient conditions on
2671: $Df\vert_0$ for the existence of a linearising transformation $\phi$.  The
2672: point $(\alpha_1,\dots,\alpha_n)$ in $\mathbb{C}^n$ is said to be of
2673: \emph{multiplicative type} $(K,v)$~\cite[p.~191]{ArnoldGM} if
2674: \begin{equation}
2675:   \label{eq:multype}
2676:   \left\vert \alpha_k-\prod_{r=1}^n \alpha_r^{j_r}\right\vert \geqslant 
2677:   K\vert \mathbf{j}\vert_1^{-v} 
2678: \end{equation}
2679: for all $\mathbf{j}\in (\mathbb{N}\cup \{0\})^n$ with
2680: $\vert \mathbf{j}\vert_1\geqslant 2$.  Siegel showed that if the vector
2681: $(\alpha_1,\dots,\alpha_n)$ of eigenvalues of $Df\vert_0$ is of
2682: multiplicative type $(K,v)$ for some $K>0$ and $v>0$, then $f$ can be
2683: linearised in a neighbourhood (further details are
2684: in~\cite{ArnoldGM,Herman87,MoserSRM}).  To stop Siegel's condition
2685: being too restrictive, one chooses $v>(n-1)/2$, since then the set of
2686: points of multiplicative type $(K,v)$ has full measure for any $K>0$.
2687: However, the neighbourhood of linearisation decreases as $v$ increases
2688: (it also depends on $K$ but less significantly) and so we do not want
2689: $v$ to be too large.
2690: 
2691: Let $\mathcal{E}_v$ denote the exceptional set of points in
2692: $\mathbb{C}^{\,n}$ (regarded as $\mathbb{R}^{2n}$) which for a given
2693: exponent $v$, fail to be of multiplicative type $(K,v)$ for any $K>0$
2694: and so fail to satisfy the conditions of Siegel's theorem and suppose
2695: if $v > (n-1)/2$.  Then $\mathcal{E}_v$ is null and its Hausdorff
2696: dimension is given by~\eqref{eq:dimLv}; namely,
2697: \begin{equation}
2698:   \label{eq:dimEv}
2699:   \dim \mathcal{E}_v = 2(n-1) + \dfrac{n+1}{v+1}.  
2700: \end{equation} 
2701: This result is established by means of an exponential map which
2702: preserves the Hausdorff dimension and allows $\mathcal{E}_v$ to be
2703: replaced by a set involving a more general kind of additive type (see
2704: next section) with a simpler structure~\cite{DRV94}.
2705: 
2706: \subsection{Lyapunov stability of vector fields}
2707: \label{sec:lypstab}
2708: Consider the differential equation
2709: \begin{equation}
2710:   \label{eq:de}
2711:   \dot{\mathbf{z}}=A\mathbf{z} + Q(\mathbf{z})\in\mathbb{C}^n,
2712: \end{equation}
2713: where $A$ is a $n\times n$ complex matrix and the holomorphic
2714: functions $Q\colon \mathbb{C}^n\to\mathbb{C}^n$ and $\partial
2715: Q_k/\partial z_j$ vanish at the origin $0$.  The obvious solution
2716: $\mathbf{z}_0(t)=0$ is said to be \emph{future (resp. past) stable} if
2717: points near $0$ remain there under evolution by~\eqref{eq:de} to the
2718: future (resp. past).  More precisely, the solution $\mathbf{z}(t)$ is
2719: future (resp. past) stable if for every neighbourhood $N$ of $0$,
2720: there exists a subneighbourhood $N'$ with $0\in N'\subset N$ such that
2721: $\mathbf{z}(0)\in N'$ guarantees that $\mathbf{z}(t)\in N$ for all
2722: $t>0$ (resp. $t<0$).  By a well known theorem of
2723: Lyapunov~\cite{MoserSRM}, the solution is future stable if the real
2724: parts of the eigenvalues of $A$ are at most $0$ and past stable if the
2725: real parts are at least $0$. Thus for the solution $\mathbf{z}(t)$ to
2726: be future and past stable or simply stable, the eigenvalues must have
2727: zero real part and so must be purely imaginary.  The stability of the
2728: solution $\mathbf{z}(t)$ is determined by a remarkable theorem due to
2729: Carath{\'e}odory and Cartan~\cite{MoserSRM} which asserts that
2730: stability is equivalent to $A$ being diagonalisable with purely
2731: imaginary eigenvalues and the vector field being holomorphically
2732: linearisable in a neighbourhood of the origin.   By Siegel's normal
2733: form theorem~\cite{Siegel42,Siegel52}, also used in the preceding
2734: section, this last condition holds if the eigenvalues
2735: $\gamma_k=i\lambda_k$, $\lambda_k\in\mathbb{R}$, satisfy for some
2736: $K$>0, $v >n-1$,
2737: \begin{equation}
2738:   \label{eq:addtype}
2739:   \abs{\lambda_k-\sum_{r=1}^n \lambda_r\, j_r}\geqslant 
2740:   K\abs{\mathbf{j}}_1^{-v}
2741: \end{equation}
2742: for all $\mathbf{j}\in (\mathbb{N}\cup \{0\})^n$ with
2743: $\vert \mathbf{j}\vert_1\geqslant 2$ (note that this is the additive
2744: form of~\eqref{eq:multype}).  The complement of this set of points
2745: $\boldsymbol{\lambda}\in\mathbb{R}^n$ of additive type is the set of
2746: $\boldsymbol{\alpha}\in\mathbb{R}^n$ such that for any $K>0$,
2747: \begin{equation}
2748:   \label{eq:notaddtype}
2749:   \abs{\alpha_k-\sum_{r=1}^n \alpha_r\, j_r}< K\abs{\mathbf{j}}_1^{-v}
2750: \end{equation}
2751: for some $\mathbf{j}\in\mathbb{Z}^n$. This set, $\widehat{E}_v$ say,
2752: is related to and has the same metrical character as the set
2753: \begin{equation}
2754:   \label{eq:hLv}
2755:   \widehat{L}_v=\{\boldsymbol{\alpha}\colon 
2756:   \abs{\mathbf{q}\cdot\boldsymbol{\alpha}}<\abs{\mathbf{q}}_\infty^{-v} 
2757:   \text{ for infinitely many } \ \mathbf{q}\in\mathbb{Z}^n\}
2758: \end{equation}
2759: and in fact by an argument similar to that giving~\eqref{eq:UEWincl},
2760: for any $\varepsilon>0$,
2761: \begin{equation*}
2762:   \widehat{L}_{v+\varepsilon}\subset \widehat{E}_v\subset
2763:   \widehat{L}_v  
2764: \end{equation*}
2765: (see~\cite[Sect.~7.5.2]{MDAM}).  This inclusion implies that the two
2766: sets $ \widehat{L}_v$ and $\widehat{E}_v$ have the same Hausdorff
2767: dimension.  The set $\widehat{L}_v $ is also related to $L_v$ and
2768: roughly speaking, has one degree of freedom less.  The Hausdorff
2769: dimension of $\widehat{L}_v $ is a special case of an `absolute value'
2770: analogue, proved by Dickinson~\cite{HD93}, of the general form of the
2771: Jarník--Besicovitch theorem:
2772: \begin{equation}
2773:   \label{eq:dimhL}
2774:   \dim \widehat{L}_v  = \dim \widehat{E}_v = n-1+ \dfrac{n}{v+1}
2775: \end{equation}
2776: when $v>n-1$ (see also~\cite{DV}); note that $
2777: \widehat{L}_v=\mathbb{R}^n$ otherwise.  Thus the exceptional set of
2778: eigenvalues for which the solution to~\eqref{eq:de} cannot be shown to
2779: be stable has Hausdorff dimension 0.
2780: 
2781: \subsection{Kolmogorov-Arnol'd-Moser theory}
2782: \label{sec:KAM}
2783: 
2784: The stability of the solar system is one of the oldest problems in
2785: mechanics~\cite{Moser78}.  It is of course a special case of the $N$
2786: body problem of understanding the motion of $N$ point masses subject
2787: only to gravitational attraction, with all other forces neglected.
2788: When $N=2$, the solution is well known and the periodic solutions in
2789: which the bodies move in an ellipse about their centre of mass persist
2790: forever.  For $N\geqslant 3$, however, the situation is
2791: extraordinarily complicated and is far from being fully understood,
2792: even for solar systems where the mass $m_N$ of the sun is much greater
2793: than the masses of the $n=N-1$ planets.  If, as a first approximation, 
2794: the centre of mass of the system is assumed to coincide with that of
2795: the sun and if the gravitational interactions between the planets and
2796: other effects are neglected, the system decouples into $n$ two-body
2797: problems, in which each planet describes an elliptical orbit around
2798: the sun, with period $T_j$ say and frequency $\omega_j=2\pi/T_j$,
2799: $j=1,\dots,n$.
2800: 
2801: For each vector $\omega=(\omega_1,\dots,\omega_n)$ of frequencies in
2802: the $n$-dimensional torus $\mathbb{T}^n=\mathbb{S}^1\times \dots\times
2803: \mathbb{S}^1$, the map $\varphi_{\omega} \colon \mathbb{R} \to
2804: \mathbb{T}^n$ given by
2805: \begin{equation*}
2806:   \varphi_{\omega}(t) = \varphi_{\omega}(0)+ t\omega
2807: \end{equation*}
2808: is a quasi-periodic flow on the torus.  
2809: 
2810: The case $n=1$ corresponds to uniform motion around a circle and so is
2811: periodic.  When the frequencies are all rational, the flow is
2812: periodic.  If the frequencies are not all rational, then by
2813: Kronecker's theorem~\cite{HW}, the flow winds round the torus, densely
2814: filling a subspace of dimension given by the number of rationally
2815: independent frequencies. Thus when the frequencies are independent,
2816: the closure of $\varphi_{\omega}(\mathbb{R})$ is the torus
2817: $\mathbb{T}^n$ and solutions will persist for ever.  Gravitational
2818: interactions between the planets are represented by a small
2819: perturbation of the original Hamiltonian describing the system.
2820: Stability then reduces to the solutions of the perturbed Hamiltonian
2821: system continuing to wind round a perturbed invariant torus.  Of
2822: course this model is idealised and takes no account of the final fate
2823: of the universe.
2824:            
2825: Details of the history of the solution to this problem are in
2826: \cite[Chap.~1]{MoserSRM}.  Siegel's success in overcoming the related
2827: `small denominator' problem in the linearisation of complex
2828: diffeomorphisms (see~\S\ref{sec:lindiffeos}, \S\ref{sec:lypstab}) was
2829: followed by Kolmogorov's conjecture that quasi-periodic solutions for
2830: a perturbed analytic Hamiltonian system not only existed but were
2831: relatively abundant in the sense that they formed a complicated Cantor
2832: type set of positive Lebesgue measure~\cite{Kolmogorov54}.  This was
2833: proved completely in 1962 by Arnol'd~\cite{Arnold63} and independently
2834: Moser proved an analogous result for sufficiently smooth `twist' maps
2835: \cite{Moser62,SiegelMoserCM}.  The results imply that for planets very
2836: much smaller than the sun and for the majority of initial conditions
2837: in which the orbits are close to co-planar circles, distances between
2838: the bodies will remain perpetually bounded, \emph{i.e.}, the planets
2839: will never collide, escape or fall into the sun.  Further details can
2840: be found in \cite{AKNIII} and \cite{KAMtut}.
2841: 
2842: 
2843: The differentiability and Diophantine conditions were relaxed
2844: substantially by R\"ussmann in \cite{Russmann83,Russmann91,Russmann01}
2845: (see also \cite[Sect.~6.3]{APIDS}, \cite{NTDS},
2846: \cite[Chap.~1]{MoserSRM}, \cite{Poschel82}).  Another approach is to
2847: use `averaging' methods~\cite{ArnoldGM}; this can involve Diophantine
2848: approximation on manifolds~\cite{DRV89a}, see also~\cite{Russmann01}.
2849: 
2850: As in the above examples, it turns out that in order to ensure
2851: convergence of a Fourier series and an infinite dimensional extension
2852: of Newton's iterative tangent method, the frequencies
2853: $\omega=(\omega_1,\dots,\omega_n)$ must satisfy a Diophantine
2854: condition, which in this case is
2855: \begin{equation}
2856:   \label{eq:kamda}
2857:   \abs{\mathbf{q}\cdot \boldsymbol{\omega}}=\abs{q_1 \omega_1 + \dots + 
2858:     q_n
2859:     \omega_n} \geqslant K\vert \mathbf{q}\vert_1^{-v}\,, 
2860: \end{equation}
2861: for some positive constants $K=K(\omega)$ and $v=v(\omega)$ for all
2862: non-zero $\mathbf{q}=(q_1,\dots, q_n)$ vectors in $\mathbb{Z}^n$.  The
2863: exponent $v$ is subject to two conflicting requirements.  It should be
2864: large enough ($v>n-1$) to ensure that the Diophantine condition above
2865: is not too restrictive, but small enough to ensure that the
2866: perturbation has physical significance and that the stability is
2867: robust.  The proof breaks down when the frequencies lie in the
2868: complementary exceptional set $E_v$, say, of frequencies which are
2869: close to resonance in the sense that, given any $K>0$, there exists a
2870: $\mathbf{q}\in\mathbb{Z}^n$ such that
2871: \begin{equation*}
2872:   \vert \mathbf{q}\cdot \boldsymbol{\alpha}\vert <K\vert
2873:   \mathbf{q}\vert_1^{-v}.
2874: \end{equation*}
2875: The set $E_v$ is related to the set $\widehat{L}_v$ discussed in the
2876: preceding section and allows us to deduce that
2877: \begin{displaymath}
2878:   \hdim E_v = \hdim \widehat{L}_v=n-1+\dfrac{n}{v+1}
2879: \end{displaymath}
2880: when $v>n-1$. 
2881: 
2882: Thus Hausdorff dimension plays its part in other branches of
2883: mathematics and mechanics, as well as in number theory. It has even
2884: been of use to Mandelbrot in arousing the interest of mathematicians
2885: in his profoundly original and imaginative
2886: ideas~\cite[Chapter~2]{Cherbit91}.
2887: 
2888: 
2889: \section{Acknowledgements}
2890: \label{sec:acknowledgements}
2891: 
2892: We thank Brent Everitt and Sanju Velani for pointing out the
2893: significance of the Bianchi groups. We also thank the referee
2894: for helpful comments.
2895: 
2896: \bibliographystyle{amsplain}
2897: \begin{thebibliography}{100}
2898: 
2899: \bibitem{AA95a}
2900: A.~G. Abercrombie, \emph{Badly approximable $p$-adic integers}, Proc.\ Indian
2901:   Acad. \ Sci. Math. \ Sci. \textbf{105} (1995), 123--134.
2902: 
2903: \bibitem{AA95}
2904: \bysame, \emph{The {H}ausdorff dimension of some exceptional sets of $p$-adic
2905:   matrices}, J.\ Number Theory \textbf{53} (1995), 311--341.
2906: 
2907: \bibitem{AhlforsMT}
2908:   L.~V. Ahlfors, \emph{M{\"o}bius Transformations in Several
2909:     Dimensions}, Ordway Professorship Lectures in Mathematics, School
2910:   of Mathematics, University of Minnesota, Minneapolis, 1981. 
2911: 
2912: \bibitem{AndersonHG}
2913: J.~W. Anderson, \emph{Hyperbolic Geometry}, Springer Undergraduate Lecture
2914:   Notes, Springer-Verlag, London, 1999.
2915: 
2916: \bibitem{Arnold63}
2917: V.~I. Arnol'd, \emph{Small denominators and problems of stability of motion in
2918:   classical and celestial mechanics}, Usp.\ Mat.\ Nauk \textbf{18} (1963),
2919:   91--192, English transl. in Russian Math.\ Surveys, {\bf{18}}\, (1963),
2920:   85--191.
2921: 
2922: \bibitem{ArnoldMM}
2923:   \bysame, \emph{Mathematical Methods of Classical Mechanics},
2924:   Springer-Verlag, New York, 1978, Translated by
2925:   K.~Vogtmann and A.~Weinstein. 
2926:   
2927: \bibitem{ArnoldGM} 
2928:   \bysame, \emph{Geometrical Methods in Ordinary
2929:     Differential Equations}, Springer-Verlag, New York, 1983,
2930:   Translated by J.~Sz{\"u}cs.
2931: 
2932: \bibitem{AKNIII}
2933:   V.~I. Arnol'd, V.~V. Kozlov, and A.~I. Neishtadt, \emph{Mathematical
2934:     Aspects of Classical and Celestial Mechanics}, Encyclopaedia of
2935:   Mathematical Sciences, vol.\ 3, Dynamical Systems III,
2936:   Springer-Verlag, Berlin, 1980, Translated by A.~Jacob.
2937: 
2938: \bibitem{APIDS}
2939: D.~K. Arrowsmith and C.~M. Place, \emph{An Introduction to Dynamical Systems},
2940:   Cambridge University Press, Cambridge, 1990.
2941: 
2942: \bibitem{BS}
2943: A.~Baker and W.~M. Schmidt, \emph{Diophantine approximation and {H}ausdorff
2944:   dimension}, Proc.\ Lond.\ Math.\ Soc. \textbf{21} (1970), 1--11.
2945: 
2946: \bibitem{BeardonGDG}
2947: A.~F. Beardon, \emph{The Geometry of Discrete Groups},
2948: Springer-Verlag, New York, 1983.
2949: 
2950: \bibitem{bbdd99}
2951: V.~V. Beresnevich, V.~I. Bernik, H.~Dickinson, and M.~M. Dodson, \emph{The
2952:   {K}hintchine-{G}roshev theorem for planar curves}, Proc.\ Roy.\ Soc.\ Lond.\
2953:   A \textbf{455} (1999), 3053--3063.
2954: 
2955: \bibitem{bbdbaker}
2956: V.~V. Beresnevich, V.~I. Bernik, and M.~M. Dodson, \emph{Regular systems,
2957:   ubiquity and {D}iophantine approximation}, in : A Panorama of Number
2958: Theory (G.~W{\"u}stholz, ed.), ETH, Z{\"u}rich, Cambridge University
2959: Press, Cambridge, 2002, pp.~260--279.
2960: 
2961: \bibitem{BDV03}
2962: V.~V. Beresnevich, H.~Dickinson, and S.~L. Velani, \emph{Measure theoretic laws
2963:   for limsup sets}, preprint, University of York, (2003).
2964: 
2965: \bibitem{bd99}
2966: V.~I. Bernik and M.~M. Dodson, \emph{Metric {D}iophantine approximation in the
2967:   field of complex numbers}, in: Number Theory and its Applications
2968: (K.~Gyory and S.~Kanemitsu, eds.), vol.~1, Kluwer Academic, Dordrecht,
2969: 1999, pp.~51--58. 
2970: 
2971: \bibitem{MDAM}
2972: \bysame, \emph{Metric {D}iophantine Approximation on Manifolds}, Cambridge
2973:   Tracts in Mathematics, No.~137, Cambridge University Press,
2974:   Cambridge, 1999.
2975: 
2976: \bibitem{bkm01}
2977: V.~I. Bernik, D.~Y. Kleinbock, and G.~A. Margulis, \emph{Khintchine-type
2978:   theorems on manifolds: the convergence case for standard and multiplicative
2979:   versions}, Inter.\ Math.\ Res.\ Notices \textbf{3} (2001), 453--485.
2980: 
2981: \bibitem{Bes28}
2982: A.~S. Besicovitch, \emph{On the fundamental geometrical properties of linearly
2983:   measurable plane sets of points}, Math.\ Ann. \textbf{98} (1928), 422--464.
2984: 
2985: \bibitem{Bes34}
2986: \bysame, \emph{Sets of fractional dimensions ({I}{V}): on rational
2987:   approximation to real numbers}, J.\ Lond.\ Math.\ Soc. \textbf{9} (1934),
2988:   126--131.
2989: 
2990: \bibitem{BJ97}
2991: C.~J. Bishop and P.~W. Jones, \emph{Hausdorff dimension and {K}leinian groups},
2992:   Acta Math. \textbf{111} (1997), 1--39.
2993: 
2994: \bibitem{Borel12}
2995: E.~Borel, \emph{Sur un probl{\`e}me de probabilit{\'e}s aux fractions
2996:   continues}, Math.\ Ann. \textbf{72} (1912), 578--584.
2997: 
2998: \bibitem{BD86}
2999: J.~D. Bovey and M.~M. Dodson, \emph{The {H}ausdorff dimension of systems of
3000:   linear forms}, Acta Arith. \textbf{45} (1986), 337--358.
3001: 
3002: \bibitem{MR1931199}
3003: Y.~Bugeaud, \emph{Approximation par des nombres alg\'ebriques de degr\'e
3004:   born\'e et dimension de {H}ausdorff}, J. Number Theory \textbf{96} (2002),
3005:   no.~1, 174--200.
3006: 
3007: \bibitem{Carat14}
3008: C.~Carath{\'e}odory, \emph{{\"U}ber das lineare {M}ass von {P}unktmengen, eine
3009:   {V}erallgemeinerung des {L}{\"a}ngenbegriffs}, G{\"o}tt.\ Nachr.\ (1914),
3010:   404--226.
3011: 
3012: \bibitem{CarlesonES}
3013: L.~Carleson, \emph{Selected Problems in Exceptional Sets}, van
3014: Nostrand, Princeton, N.J., 1967.
3015: 
3016: \bibitem{Casselshort}
3017: J.~W.~S. Cassels, \emph{An Introduction to {D}iophantine Approximation},
3018:   Cambridge University Press, New York, 1957.
3019: 
3020: \bibitem{CasselsLF}
3021: \bysame, \emph{Local Fields}, Cambridge University Press, Cambridge, 1986.
3022: 
3023: \bibitem{MR27:4799}
3024: Jing-run Chen, \emph{The lattice-points in a circle}, Sci. Sinica \textbf{12}
3025:   (1963), 633--649.
3026: 
3027: \bibitem{Cherbit91}
3028: G.~Cherbit~(ed.), \emph{Fractals: Nonintegral Dimensions and Applications},
3029:   John Wiley \& Sons, Chichester, 1991, translated by F.~Jellett from
3030:   {\em Fractals: Dimensions Non-enti{\`e}res et Applications} (Masson,
3031:   1987). 
3032: 
3033: \bibitem{chown02:_why}
3034: M.~Chown, \emph{Why should nature have a favourite number?}, New Scientist
3035:   \textbf{76} (2002), no.~2374, 54--56.
3036: 
3037: \bibitem{Chung}
3038: K.~L. Chung, \emph{A Course in Probability Theory}, 2nd ed., Academic
3039: Press, New York--London, 1974.
3040: 
3041: \bibitem{ChungErdos52}
3042: K.~L. Chung and P.~Erd{\H{o}}s, \emph{On the application of the
3043:   {B}orel-{C}antelli lemma}, Trans.\ Amer.\ Math.\ Soc. \textbf{72} (1952),
3044:   179--186.
3045: 
3046: \bibitem{Dani85}
3047: S.~G. Dani, \emph{Divergent trajectories of flows on homogeneous spaces and
3048:   homogeneous {D}iophantine approximation}, J.\ reine angew.\ Math.
3049:   \textbf{359} (1985), 55--89.
3050: 
3051: \bibitem{Dani86a}
3052: \bysame, \emph{Bounded orbits of flows on homogeneous spaces}, Comm.\ Math.\
3053:   Helv. \textbf{61} (1986), 636--660.
3054: 
3055: \bibitem{DaniNTDS}
3056: \bysame, \emph{On badly approximable numbers, {S}chmidt games and bounded
3057:   orbits of flows}, in: Number Theory and Dynamical Systems (M.~M. Dodson and
3058:   J.~A.~G. Vickers, eds.), LMS Lecture Note Series, vol. 134, Cambridge
3059:   University Press, Cambridge, 1987, pp.~69--86.
3060: 
3061: \bibitem{KAMtut}
3062: R.~de~la Llave, \emph{A tutorial on {KAM} theory}, in: Smooth Ergodic
3063: Theory and Applications, Proceedings of Symposia in Pure Mathematics,
3064: Summer Research Institute, Seattle WA, American Mathematical Society, Providence, RI, 2001.
3065: 
3066: \bibitem{deMa70}
3067: B.~de~Mathan, \emph{Approximations diophantiennes dans un corps local}, Bull.
3068:   Soc. Math. France Suppl. M\'em. \textbf{21} (1970), 1--93.
3069: 
3070: \bibitem{HD93}
3071: H.~Dickinson, \emph{The {H}ausdorff dimension of systems of simultaneously
3072:   small linear forms}, Mathematika \textbf{40} (1993), 367--374.
3073: 
3074: \bibitem{dd00}
3075: H.~Dickinson and M.~M. Dodson, \emph{Extremal manifolds and {H}ausdorff
3076:   dimension}, Duke Math.\ J. \textbf{101} (2000), 271--281.
3077: 
3078: \bibitem{ddj99}
3079: H.~Dickinson, M.~M. Dodson, and J.~Yuan, \emph{Hausdorff dimension and
3080:   {$p$}-adic {D}iophantine approximation}, Indag.\ Mathem., N.S. \textbf{10}
3081:   (199), 337--347.
3082: 
3083: \bibitem{HDSV97}
3084: H.~Dickinson and S.~L. Velani, \emph{Hausdorff measure and linear forms}, J.\
3085:   reine angew.\ Math. \textbf{490} (1997), 1--36.
3086: 
3087: \bibitem{mmd92}
3088: M.~M. Dodson, \emph{Hausdorff dimension, lower order and {K}hintchine's theorem
3089:   in metric {D}iophantine approximation}, J.\ reine angew. Math. \textbf{432}
3090:   (1992), 69--76.
3091: 
3092: \bibitem{mmdnewton}
3093: \bysame, \emph{Exceptional sets in dynamical systems and {D}iophantine
3094:   approximation}, in: Proceedings of Rigidity in Dynamics and Geometry
3095: Conference 2000 (M.~Burger and A.~Iozzi, eds.),
3096: Newton Institute, Cambridge, Springer-Verlag, Berlin, 2001, pp.~39--60.
3097: 
3098: \bibitem{DRV89a}
3099: M.~M. Dodson, B.~P. Rynne, and J.~A.~G. Vickers, \emph{Averaging in
3100:   multi-frequency systems}, Nonlinearity \textbf{2} (1989), 137--148.
3101: 
3102: \bibitem{DRV90a}
3103: \bysame, \emph{Diophantine approximation and a lower bound for {H}ausdorff
3104:   dimension}, Mathematika \textbf{37} (1990), 59--73.
3105: 
3106: \bibitem{DRV94}
3107: \bysame, \emph{The {H}ausdorff dimension of exceptional sets associated with
3108:   normal forms}, J.\ Lond.\ Math.\ Soc. \textbf{49} (1994), 614--624.
3109: 
3110: \bibitem{DV}
3111: M.~M. Dodson and J.~A.~G. Vickers, \emph{Exceptional sets in
3112:   {K}olmogorov-{A}rnol'd-{M}oser theory}, J.\ Phys.\ A \textbf{19} (1986),
3113:   349--374.
3114: 
3115: \bibitem{NTDS}
3116: M.~M. Dodson and J.~A.~G. Vickers~(eds.), \emph{Number Theory and Dynamical
3117:   Systems}, Lond.\ Mathematical Society Lecture Note Series, vol. 134,
3118:   Cambridge University Press, Cambridge, 1989.
3119: 
3120: \bibitem{FalcGFSshort}
3121: K.~Falconer, \emph{The Geometry of Fractal Sets}, Cambridge University
3122: Press, Cambridge, 1985.
3123: 
3124: \bibitem{FalcFG}
3125: \bysame, \emph{Fractal Geometry}, John Wiley, Chichester, 1989.
3126: 
3127: \bibitem{MR95c:35030}
3128: M.~Fe{\v{c}}kan, \emph{Periodic solutions of certain abstract wave equations},
3129:   Proc. Amer. Math. Soc. \textbf{123} (1995), no.~2, 465--470.
3130: 
3131: \bibitem{Fed}
3132: H.~Federer, \emph{Geometric Measure Theory}, Springer-Verlag, New
3133: York, 1969.
3134: 
3135: \bibitem{FM95}
3136: J.-L. Fernandez and M.~V. Meli{\'a}n, \emph{Bounded geodesics of {R}iemann
3137:   surfaces and hyperbolic manifolds}, Trans.\ Amer.\ Math.\ Soc. \textbf{9}
3138:   (1995), 3533--3549.
3139: 
3140: \bibitem{Ford25}
3141: L.~R. Ford, \emph{On the closeness of approach of complex rational fractions to
3142:   a complex irrational number}, Trans.\ Amer. \ Math.\ Soc. \textbf{27} (1925),
3143:   146--154.
3144: 
3145: \bibitem{Forder63}
3146: H.~G. Forder, \emph{A simple proof of a result on {D}iophantine approximation},
3147:   Math.\ Gaz. \textbf{47} (1963), 237--238.
3148: 
3149: \bibitem{Frostman}
3150: O.~Frostman, \emph{Potentiel d'\'equilibre et capacit{\'e} des ensembles avec
3151:   quelques applications {\`a} la th{\'e}orie des fonctions}, Meddel.\ Lunds
3152:   Univ.\ Math.\ Sem. \textbf{3} (1935), 1--118.
3153: 
3154: \bibitem{HW}
3155: G.~H. Hardy and E.~M. Wright, \emph{An Introduction to the Theory of Numbers},
3156:   4th ed., Clarendon Press, New York, 1960.
3157: 
3158: \bibitem{HarmanMNT}
3159: G.~Harman, \emph{Metric Number Theory}, LMS Monographs New Series, vol.~18,
3160:   Clarendon Press, New York, 1998.
3161: 
3162: \bibitem{Harman02}
3163: \bysame, \emph{Non-linear {D}iophantine approximation of complex numbers},
3164:   Math. \ Proc.\ Cambridge Philos. \ Soc. \textbf{133} (2002), 205--212.
3165: 
3166: \bibitem{Hausdorff19}
3167: F.~Hausdorff, \emph{Dimension und {\"a}usseres {M}ass}, Math.\ Ann. \textbf{79}
3168:   (1919), 157--179.
3169: 
3170: \bibitem{Herman87}
3171: M.~R. Herman, \emph{Recent results and some open questions on {S}iegel's
3172:   linearisation theorem on germs of complex analytic diffeomorphisms of
3173:   {$\mathbb{C}^n$} near a fixed point}, in: Proceedings of the VIIIth
3174: International Congress of Mathematical Physics, Marseilles 1986
3175: (M.~Mebkhout and R.~S{\/e}n{\/e}or, eds.), World Scientific,
3176: Singapore, 1987, pp.~138--184.
3177: 
3178: \bibitem{HV95}
3179: R.~Hill and S.~L. Velani, \emph{Ergodic theory of shrinking targets}, Invent.\
3180:   Math. \textbf{119} (1995), 175--198.
3181: 
3182: \bibitem{HVJBGF}
3183: \bysame, \emph{The {J}arn{\'\i}k-{B}esicovitch theorem for geometrically finite
3184:   {K}leinian groups}, Proc.\ Lond.\ Math.\ Soc. \textbf{77} (1998), 524--550.
3185: 
3186: \bibitem{Hurwitz1891}
3187: A.~Hurwitz, \emph{{\"U}ber die angen{\"a}herte {D}arstellung der
3188:   {I}rrationalzahlen durch rationale {B}r{\"u}che}, Math.\ Ann. \textbf{46}
3189:   (1891), 279--284.
3190: 
3191: \bibitem{Ja28}
3192: V.~Jarn{\'\i}k, \emph{Zur metrischen {T}heorie der diophantischen
3193:   {A}pproximationen}, Prace Mat.-Fiz. (1928-9), 91--106.
3194: 
3195: \bibitem{Ja29}
3196: \bysame, \emph{Diophantischen {A}pproximationen und {H}ausdorff\-sches {M}ass},
3197:   Mat.\ Sbornik \textbf{36} (1929), 371--382.
3198: 
3199: \bibitem{Ja30a}
3200: \bysame, \emph{N{\v e}kolik pozn{\'a}mek o {H}ausdorffov{\v e} m{\'\i}{\v r}e},
3201:   Rozpravy T{\'r}.\ {\v{C}}esk{\'e} Akad \textbf{40} (1930), c.~9, 1--8.
3202: 
3203: \bibitem{Ja30b}
3204: \bysame, \emph{Quelques remarques sur la mesure de {M}.\ {H}ausdorff}, Bull.\
3205:   Int.\ l'Acad. des Sci.\ Boh\^eme (1930), 1--6.
3206: 
3207: \bibitem{Ja31}
3208: \bysame, \emph{{\"U}ber die simultanen diophantischen {A}pproximat\-ionen},
3209:   Math.\ Z. \textbf{33} (1931), 503--543.
3210: 
3211: \bibitem{Ja45}
3212: \bysame, \emph{Sur les approximations diophantiques des nombres {$p$}-adiques},
3213:   Revista Ci., Lima \textbf{47} (1945), 489--505.
3214: 
3215: \bibitem{KatokHass}
3216: A~Katok and B.~Hasselblatt, \emph{Introduction to the Modern Theory of
3217:   Dynamical Systems}, Cambridge University Press, Cambridge, 1995.
3218: 
3219: \bibitem{Kh24}
3220: A.~I. Khintchine, \emph{Einige {S}{\"a}tze {\"u}ber {K}ettenbruche, mit
3221:   {A}nwendungen auf die {T}heorie der {D}iophantischen {A}pproximationen},
3222:   Math.\ Ann. \textbf{92} (1924), 115--125.
3223: 
3224: \bibitem{Kh25}
3225: \bysame, \emph{{\"U}ber die angen{\"a}herte {A}ufl{\"o}sung linearer
3226:   {G}leichungen in ganzen {Z}ahlen}, Rec.\ math.\ Soc.\ Moscou Bd. \textbf{32}
3227:   (1925), 203--218.
3228: 
3229: \bibitem{Kh25a}
3230: \bysame, \emph{Zwei {B}ermerkungen zu einer {A}rbeit des {H}errn {P}erron},
3231:   Math.\ Z. \textbf{22} (1925), 274--284.
3232: 
3233: \bibitem{Kh26}
3234: \bysame, \emph{Zur metrischen {T}heorie der diophantischen {A}pproximationen},
3235:   Math.\ Z. \textbf{24} (1926), 706--714.
3236: 
3237: \bibitem{KandT}
3238: J.~Kingman and S.~J. Taylor, \emph{An {I}ntroduction to {P}robability and
3239:   {M}easure}, Cambridge University Press, London, 1966.
3240: 
3241: \bibitem{Kleinbock99}
3242: D.~Y. Kleinbock, \emph{Badly approximable systems of affine forms}, J.~Number
3243:   Theory \textbf{79} (1999), 83--102.
3244: 
3245: \bibitem{KM98a}
3246: D.~Y. Kleinbock and G.~A. Margulis, \emph{Flows on homogeneous spaces and
3247:   {D}iophantine approximation on manifolds}, Ann.~of Math. (2)
3248: \textbf{148} (1998), no.~1, 339--360.
3249: 
3250: \bibitem{KS64}
3251: S.~Kochen and C.~Stone, \emph{A note on the {B}orel-{C}antelli lemma}, Ill.\
3252:   J.\ Math. \textbf{8} (1964), 248--251.
3253: 
3254: \bibitem{Koksma}
3255: J.~F. Koksma, \emph{Diophantische {A}pproximationen}, Ergebnisse d.\ {M}ath.\
3256:   u.\ ihrer {G}renzgebiete, vol.~4, Springer-Verlag, Berlin, 1936.
3257: 
3258: \bibitem{Kolmogorov54}
3259: A.~N. Kolmogorov, \emph{On the conservation of conditionally periodic motions
3260:   under small perturbations of the {H}amiltonian}, Dokl.\ Akad.\ Nauk SSSR
3261:   \textbf{98} (1954), 527--530, in Russian.
3262: 
3263: \bibitem{kristensen03}
3264: S.~Kristensen, \emph{Badly approximable linear forms over a field of formal
3265:   series}, preprint (2002).
3266: 
3267: \bibitem{kristensen:_dioph_schroed}
3268: \bysame, \emph{Diophantine approximation and the solubility of the
3269:   {S}chr{\"o}dinger equation}, preprint (2003).
3270: 
3271: \bibitem{kristensen02}
3272: \bysame, \emph{On well-approximable matrices over a field of formal series}, to
3273:   appear (2003).
3274: 
3275: \bibitem{leveque1952}
3276: W.~J. LeVeque, \emph{Continued fractions and approximations {I} and {II}},
3277:   Indag.\ Math. \textbf{14} (1952), 526--545.
3278: 
3279: \bibitem{JL98}
3280: J.~Levesley, \emph{A general inhomogeneous {J}arník--{B}esicovitch theorem},
3281:   J.~Number Theory \textbf{71} (1998), 65--80.
3282: 
3283: \bibitem{Lutz}
3284: E.~Lutz, \emph{Sur les Approximations Diophantiennes Lin\'eaires et
3285:   $p$-adiques}, Actualit\'es Sci.\ Indust., vol. 1224, Hermann, Paris,
3286: 1955. 
3287: 
3288: \bibitem{Mackay83}
3289: R.~S. Mackay, \emph{A renormalisation approach to invarant circles in
3290:   area-preserving maps}, Physica \textbf{7D} (1983), 283--300.
3291: 
3292: \bibitem{MattilaGS}
3293: P.~Mattila, \emph{Geometry of Sets and Measures in {E}uclidean Space},
3294:   Cambridge University Press, Cambridge, 1995.
3295: 
3296: \bibitem{MR99e:58145}
3297: C.~T. McMullen and D.~Sullivan, \emph{Quasiconformal homeomorphisms and
3298:   dynamics. {III}. {T}he {T}eichm\"uller space of a holomorphic dynamical
3299:   system}, Adv. Math. \textbf{135} (1998), no.~2, 351--395.
3300: 
3301: \bibitem{MP93}
3302: M.~V. Meli{\'a}n and D.~Pestana, \emph{Geodesic excursions into cusps in finite
3303:   volume hyperbolic manifolds}, Mich.\ Math.\ J. \textbf{40} (1993), 77--93.
3304: 
3305: \bibitem{Moser62}
3306: J.~Moser, \emph{On invariant curves of area-preserving maps of an annulus},
3307:   Nachr.\ Akad.\ Wiss.\ G{\"o}tt., Math.\ Phys.\ Kl. (1962), 1--20.
3308: 
3309: \bibitem{MoserSRM}
3310: \bysame, \emph{Stable and Random Motions in Dynamical Systems}, Princeton
3311:   University Press, Princeton, NJ, 1973.
3312: 
3313: \bibitem{Moser78}
3314: \bysame, \emph{Is the solar system stable?}, Math.\ Intelligencer \textbf{1}
3315:   (1978), 65--71.
3316: 
3317: \bibitem{NichollsETDG}
3318: P.~J. Nicholls, \emph{The Ergodic Theory of Discrete Groups}, LMS Lecture
3319:   Notes, vol. 143, Cambridge University Press, Cambridge, 1989.
3320: 
3321: \bibitem{NiteckiDD}
3322: Z.~Nitecki, \emph{Differentiable Dynamics}, MIT\ Press, Cambridge,
3323: Mass., 1971.
3324: 
3325: \bibitem{MR50:7828}
3326: B.~Nov{\'a}k, \emph{Remark on periodic solutions of a linear wave equation in
3327:   one dimension}, Comment. Math. Univ. Carolinae \textbf{15} (1974), 513--519.
3328: 
3329: \bibitem{NSJobit}
3330: B.~Nov\'ak and {\v{S}}t. Schwarz, \emph{Vojt{\v{e}}ch {J}arn{\'{\i}}k}, Acta
3331:   Arith. \textbf{20} (1972), 107--123.
3332: 
3333: \bibitem{Patterson76a}
3334: S.~J. Patterson, \emph{Diophantine approximation in {F}uchsian groups}, Phil.\
3335:   Trans.\ Roy.\ Soc.\ Lond.\ A \textbf{262} (1976), 527--563.
3336: 
3337: \bibitem{Patterson76b}
3338: \bysame, \emph{The limit set of a {F}uchsian group}, Acta Math. \textbf{136}
3339:   (1976), 241--273.
3340: 
3341: \bibitem{Patterson89}
3342: \bysame, \emph{Metric {D}iophantine approximation of quadratic forms},
3343: in: Number Theory and Dynamical Systems (M.~M Dodson and
3344: J.~A.~G. Vickers, eds.), LMS Lecture Notes, vol. 134, Cambridge
3345: University Press, Cambridge, 1989, pp.~37--48. 
3346: 
3347: \bibitem{MR88e:00019}
3348: H.-O. Peitgen and P.~H. Richter, \emph{The Beauty of Fractals},
3349:   Springer-Verlag, Berlin, 1986, Images of complex dynamical systems.
3350: 
3351: \bibitem{Poschel82}
3352: J.~P{\"o}schel, \emph{Integrability of {H}amiltonian systems on {C}antor sets},
3353:   Comm.\ Pure Appl. \ Math. \textbf{35} (1982), 653--696.
3354: 
3355: \bibitem{Ptashnik84}
3356: B.~I. Ptashnik, \emph{Improper Boundary Problems for Partial Differential
3357:   Equations}, Naukova Dumka, Kiev, 1984, in Russian.
3358: 
3359: \bibitem{RenyiPT}
3360: A.~R{\'e}nyi, \emph{Probability Theory}, North-Holland, Amsterdam, 1970.
3361: 
3362: \bibitem{Rogers}
3363: C.~A. Rogers, \emph{Hausdorff Measure}, Cambridge University Press,
3364: London, 1970.
3365: 
3366: \bibitem{Russmann83}
3367: H.~R. R{\"u}ssmann, \emph{On the existence of invariant curves of twist
3368:   mappings of the annulus}, in: Geometric {D}ynamics, Lecture Notes in
3369: Mathematics, vol. 1007, Springer-Verlag, Berlin, 1983, pp.~677--712.
3370: 
3371: \bibitem{Russmann91}
3372: \bysame, \emph{On the frequencies of quasi-periodic solutions of analytic
3373:   nearly integrable {H}amiltonian systems}, in: Progress in {N}onlinear
3374:   {D}ifferential {E}quations and their Applications, 12 (V.~Lazutkin S.~Kuksin
3375:   and J.~P{\"o}schel, eds.), vol.~12, Birkh{\"a}user Verlag, Basel, 1994,
3376:   pp.~51--58.
3377: 
3378: \bibitem{Russmann01}
3379: \bysame, \emph{Invariant tori in non-degenerate nearly integrable {H}amiltonian
3380:   systems}, Regul.\ Chaotic Dyn. \textbf{6} (2001), 119--204.
3381: 
3382: \bibitem{Rynne92b}
3383: B.~P. Rynne, \emph{The {H}ausdorff dimension of certain sets arising from
3384:   {D}iophantine approximation by restricted sequences of integer vectors}, Acta
3385:   Arith. \textbf{61} (1992), 69--81.
3386: 
3387: \bibitem{ASchmidt67}
3388: A.~L. Schmidt, \emph{Farey triangles and {F}arey quadrangles in the complex
3389:   plane}, Math.\ Scand. \textbf{21} (1967), 241--295.
3390: 
3391: \bibitem{ASchmidt69}
3392: \bysame, \emph{Farey simplices in the space of quaternions}, Math.\ Scand.
3393:   \textbf{24} (1969), 31--65.
3394: 
3395: \bibitem{ASchmidt74}
3396: \bysame, \emph{On the approximation of quaternions}, Math.\ Scand. \textbf{34}
3397:   (1974), 184--186.
3398: 
3399: \bibitem{ASchmidt75a}
3400: \bysame, \emph{{D}iophantine approximation of complex numbers}, Acta Math.
3401:   \textbf{134} (1975), 1--85.
3402: 
3403: \bibitem{ASchmidt84}
3404: \bysame, \emph{Classical quantum models and arithmetic problems}, Lecture Notes
3405:   in Pure and App.\ Math., ch.~{D}iophantine approximation of complex numbers,
3406:   pp.~353--377, Dekker, New York, 1984.
3407: 
3408: \bibitem{Schmidt60}
3409: W.~M. Schmidt, \emph{A metrical theorem in {D}iophantine approximation}, Can.\
3410:   J.\ Math. \textbf{12} (1960), 619--631.
3411: 
3412: \bibitem{Schmidt66}
3413: \bysame, \emph{On badly approximable numbers and certain games}, Trans.\ Amer.
3414:   \ Math.\ Soc. \textbf{123} (1966), 178--199.
3415: 
3416: \bibitem{Schmidt69}
3417: \bysame, \emph{Badly approximable systems of linear forms}, J.\ Number
3418: Theory \textbf{1} (1969), 139--154.
3419: 
3420: \bibitem{SchmidtDA}
3421: \bysame, \emph{Diophantine Approximation}, Lecture Notes in Mathematics, vol.
3422:   785, Springer-Verlag, Berlin, 1980.
3423: 
3424: \bibitem{Series85}
3425: C.~Series, \emph{The modular surface and continued fractions}, J.\ Lond.\
3426:   Math.\ Soc. \textbf{31} (1985), 69--80.
3427: 
3428: \bibitem{MR93g:11011}
3429: J.~Shallit, \emph{Real numbers with bounded partial quotients: a survey},
3430:   Enseign. Math. (2) \textbf{38} (1992), no.~1-2, 151--187.
3431: 
3432: \bibitem{shishikura1998}
3433: M.~Shishikura, \emph{The {H}ausdorff dimension of the boundary of the
3434:   {M}andelbrot set and {J}ulia sets}, Ann. \ of Math. \textbf{147} (1998),
3435:   225--267.
3436: 
3437: \bibitem{Siegel42}
3438: C.~L. Siegel, \emph{Iteration of analytic functions}, Ann.~of
3439: Math. (2) \textbf{43} (1942), 607--612.
3440: 
3441: \bibitem{Siegel52}
3442: \bysame, \emph{{\"U}ber die {N}ormalform analytischer {D}ifferentialgleichungen
3443:   in der {N}{\"a}he einer {G}leichgewichtsl{\"o}sung}, Nachr.\ Akad.\ Wiss.\
3444:   G{\"o}tt.\ Math-Phys.\ Kl (1952), 21--30.
3445: 
3446: \bibitem{SiegelMoserCM}
3447: C.~L. Siegel and J.~K. Moser, \emph{Lectures on Celestial Mechanics},
3448:   Springer-Verlag, Berlin, 1971.
3449: 
3450: \bibitem{Sprindzuk}
3451: V.~G. Sprind{\v z}uk, \emph{Metric Theory of {D}iophantine Approximations},
3452:   V. H. Winston \& Sons, Washington, D.C., 1979, Translated by
3453:   R.~A.~Silverman. 
3454: 
3455: \bibitem{Spr80}
3456: \bysame, \emph{Achievements and problems in {D}iophantine approximation
3457:   theory}, Usp.\ Mat.\ Nauk \textbf{35} (1980), 3--68, English transl. in
3458:   Russian Math.\ Surveys, {\bf{35}}\, (1980), 1--80.
3459: 
3460: \bibitem{Stratmann94}
3461: B.~Stratmann, \emph{Diophantine approximation in {K}leinian groups}, Math.\
3462:   Proc.\ Cambridge Philos.\ Soc. \textbf{116} (1994), 57--68.
3463: 
3464: \bibitem{SV95}
3465: B.~Stratmann and S.~L. Velani, \emph{The {P}atterson measure for geometrically
3466:   finite groups with parabolic elements, new and old}, Proc.\ Lond.\ Math. \
3467:   Soc. \textbf{71} (1995), 197--220.
3468: 
3469: \bibitem{Sullivan82}
3470: D.~Sullivan, \emph{Disjoint spheres, approximation by imaginary numbers, and
3471:   the logarithm law for geodesics}, Acta Math. \textbf{149} (1982), 215--237.
3472: 
3473: \bibitem{Sullivan84}
3474: \bysame, \emph{Entropy, {H}ausdorff measures old and new, and the limit set of
3475:   geometrically finite {K}leinian groups}, Acta Math. \textbf{153} (1984),
3476:   259--277.
3477: 
3478: \bibitem{MR87i:58103}
3479: \bysame, \emph{Quasiconformal homeomorphisms and dynamics. {I}. {S}olution of
3480:   the {F}atou--{J}ulia problem on wandering domains}, Ann. of Math. (2)
3481:   \textbf{122} (1985), no.~3, 401--418.
3482: 
3483: \bibitem{MR87i:58104}
3484: \bysame, \emph{Quasiconformal homeomorphisms and dynamics. {II}. {S}tructural
3485:   stability implies hyperbolicity for {K}leinian groups}, Acta Math.
3486:   \textbf{155} (1985), no.~3-4, 243--260.
3487: 
3488: \bibitem{Velani93}
3489: S.~L. Velani, \emph{Diophantine approximation and {H}ausdorff dimension in
3490:   {F}uchsian groups}, Math.\ Proc.\ Cambridge Philos.\
3491: Soc. \textbf{113} (1993), 343--354.
3492: 
3493: \bibitem{VelaniKTHD}
3494: \bysame, \emph{Geometrically finite groups, {K}hintchine-type theorems and
3495:   {H}ausdorff dimension}, Math.\ Proc.\ Cambridge Philos.\
3496: Soc. \textbf{120} (1996), 647--662.
3497: 
3498: \bibitem{Yoccoz84}
3499: J.-C. Yoccoz, \emph{Conjugaison diff{\'e}rentiable des diffeomorphismes du
3500:   cercle dont le nombre de rotation v{\'e}rifie une condition diophantienne},
3501:   Ann.\ Sci.\ Ec.\ Norm.\ Sup. \textbf{17} (1984), 333--359.
3502: 
3503: \bibitem{Yoccoz95a}
3504: \bysame, \emph{Petits diviseurs en dimension 1}, Ast{\'e}risque \textbf{231}
3505:   (1995).
3506: 
3507: \end{thebibliography}
3508: 
3509: 
3510: \end{document}
3511: 
3512: %%% Local Variables: 
3513: %%% mode: latex
3514: %%% TeX-master: t
3515: %%% End: 
3516: 
3517: 
3518: