0710.4109/a.tex
1: % Sept. 30, 2007: journal version (last edited by Adrian)
2: 
3: \documentclass[11pt]{article}
4: \usepackage{fullpage}
5: \usepackage{graphicx,amssymb,amsmath,theorem}
6: \usepackage{epsfig}
7: \usepackage{epsf}
8: \usepackage{times}
9: \usepackage{amsfonts}
10: \usepackage{latexsym}
11: \usepackage{bbm}     % blackboard characters (e.g. set of real numbers)
12: 
13: \newtheorem{theorem}{Theorem}
14: %[section]
15: \newtheorem{lemma}{Lemma}
16: \newtheorem{proposition}{Proposition}
17: \newtheorem{corollary}{Corollary}
18: \newtheorem{observation}{Observation}
19: \newtheorem{claim}{Claim}
20: \newtheorem{fact}{Fact}
21: \newtheorem{conjecture}{Conjecture}
22: \def \proof {\par \noindent {\bf Proof.}\hskip 5pt}
23: \def \endproof {\hfill $\Box$ \smallskip}%\linebreak \medskip}
24: 
25: \setlength{\oddsidemargin}{0in} \setlength{\evensidemargin}{0in}
26: \setlength{\textheight}{9in}
27: \setlength{\textwidth}{6.5in}
28: %\setlength{\topmargin}{-0.25in}
29: \setlength{\topmargin}{0in}
30: \setlength{\headheight}{0in}
31: \setlength{\headsep}{0in}
32: 
33: \long\def\ignore#1{}
34: \def\cut{\ignore}
35: \newcommand{\later}[1]{{}}
36: \newcommand{\old}[1]{{}}
37: 
38: \newenvironment{proofof}[1]
39:       {\medskip\noindent{\bf Proof of #1:}\hspace{1mm}}
40:       {\hfill$\Box$\medskip}
41: 
42: 
43: \newcommand{\NN}{\mathbb{N}} %  set of natural numbers
44: \newcommand{\ZZ}{\mathbb{Z}} %  set of integer numbers
45: \newcommand{\RR}{\mathbb{R}} %  set of real numbers
46: \newcommand{\QQ}{\mathbb{Q}} %  set of rational numbers
47: \newcommand{\SH}{\mathbb{S}} %  set of unit vectors
48: 
49: \def\reals{{\mathbb R}}
50: 
51: \def\A{{\cal A}}
52: \def\B{{\cal B}}
53: \def\C{{\cal C}}
54: \def\L{{L}}
55: \def\N{{\cal N}}
56: \def\T{{T}}
57: \def\U{{\cal U}}
58: \def\bd{{\partial}}
59: \def\polylog{{\rm polylog}}
60: \def\eps{\varepsilon}
61: \def\etal{{\it et~al.}\,}
62: \def\uu{{\bf u}}
63: \def\vv{{\bf v}}
64: \def\xx{{\bf x}}
65: \def\kk{{\bf k}}
66: 
67: 
68: \def\marrow{{\marginpar[\hfill$\longrightarrow$]{$\longleftarrow$}}}
69: \def\adrian#1{{\sc Adrian says: }{\marrow\sf #1}}
70: \def\micha#1{{\sc Micha says: }{\marrow\sf #1}}
71: \def\csaba#1{{\sc Csaba says: }{\marrow\sf #1}}
72: \newcommand\comm[1]{\typeout{Used \string\comm...}\marrow
73: \quad{\bf [[#1]]}\quad}
74: 
75: 
76: \title{Extremal problems on triangle areas in two and three dimensions}
77: 
78: \author{Adrian Dumitrescu\thanks{%
79: Department of Computer Science,
80: University of Wisconsin-Milwaukee, USA,
81: email: {\tt ad@cs.uwm.edu.}
82: Research partially supported by NSF CAREER grant CCF-0444188.}
83: \and
84: Micha Sharir\thanks{%
85: School of Computer Science, Tel Aviv University,
86: Tel Aviv, Israel, and Courant Institute, New York University, New
87: York, NY, USA,
88: email: {\tt michas@tau.ac.il.}
89: Research partially supported by NSF Grant CCR-05-14079,
90: by a grant from the U.S.-Israeli Binational Science Foundation,
91: by grant 155/05 from the Israel Science Fund, Israeli Academy of Sciences,
92: and by the Hermann Minkowski--MINERVA Center for Geometry at Tel Aviv
93: University.}
94: \and
95: Csaba D. T\'oth\thanks{%
96: Department of Mathematics,
97: Massachusetts Institute of Technology, Cambridge, MA, USA,
98: email: {\tt toth@math.mit.edu.}}}
99: 
100: \begin{document}
101: 
102: \maketitle
103: \thispagestyle{empty}
104: 
105: 
106: \begin{abstract}
107: 
108: The study of extremal problems on triangle areas was initiated in
109: a series of papers by Erd\H{o}s and Purdy in the early 1970s.
110: % Erd\H{o}s-type extremal discrete geometry problems have attracted
111: % continued attention in the combinatorics and computational geometry
112: % communities, due to their relevance for computational problems (in
113: % range counting, pattern matching, motion planning, and others) and the
114: % rich techniques developed for improving extremal bounds (such as the
115: % $\eps$-net theory, the Crossing Lemma, quasi-planar graphs,
116: % etc.), which proved to be instrumental in many areas of computational
117: % geometry.
118: In this paper we present new results on such problems, concerning the
119: number of triangles of the same area that are spanned by finite point
120: sets in the plane and in 3-space, and the number of distinct areas 
121: determined by the triangles. 
122: 
123: In the plane, our main result is an $O(n^{44/19}) =O(n^{2.3158})$
124: upper bound on the number of unit-area triangles spanned by $n$
125: points, which is the first breakthrough improving the classical bound of
126: $O(n^{7/3})$ from 1992. We also make progress in a number of important
127: special cases: We show that (i) For points in convex position, there
128: exist $n$-element point sets that span $\Omega(n\log n)$ triangles of
129: unit area. (ii) The number of triangles of minimum (nonzero) area
130: determined by $n$ points is at most $\frac{2}{3}(n^2-n)$; there exist
131: $n$-element point sets (for arbitrarily large $n$)
132: that span $(6/\pi^2-o(1))n^2$ minimum-area
133: triangles.  (iii) The number of acute triangles of minimum area
134: determined by $n$ points is $O(n)$; this is asymptotically tight. (iv)
135: For $n$ points in convex position, the number of triangles of minimum
136: area is $O(n)$; this is asymptotically tight. (v) If no three points
137: are allowed to be collinear, there are $n$-element point sets that
138: span $\Omega(n\log n)$ minimum-area triangles
139: (in contrast to (ii), where collinearities are allowed and
140: a quadratic lower bound holds).
141: 
142: In 3-space we prove an $O(n^{17/7}\beta(n))= O(n^{2.4286})$ upper
143: bound on the number of unit-area triangles spanned by $n$ points,
144: where $\beta(n)$ is an extremely slowly growing function related to the
145: inverse Ackermann function. The best previous bound, $O(n^{8/3})$, is
146: an old result of Erd\H{o}s and Purdy from 1971. We further show,
147: for point sets in 3-space:
148: (i) The number of minimum nonzero area triangles is at most $n^2+O(n)$,
149: and this is worst-case optimal, up to a constant factor.
150: (ii) There are $n$-element point sets that span
151: $\Omega(n^{4/3})$ triangles of maximum area, all incident to a common
152: point. In any $n$-element point set, the maximum number of
153: maximum-area triangles incident to a common point is
154: $O(n^{4/3+\eps})$, for any $\eps>0$.
155: (iii) Every set of $n$ points, not all on a line, determines at least
156: $\Omega(n^{2/3}/\beta(n))$ triangles of distinct areas, which share a
157: common side.
158: \end{abstract}
159: 
160: 
161: \newpage
162: \setcounter{page}{1}
163: 
164: \section{Introduction}
165: \label{sec:intro}
166: 
167: Given $n$ points in the plane, consider the following equivalence
168: relation defined on the set of (nondegenerate) triangles  spanned by
169: the points: two triangles are {\em equivalent} if they have the
170: same area.
171: Extremal problems typically ask for the maximum cardinality of an
172: equivalence class, and for the minimum number of distinct equivalence
173: classes, in a variety of cases. A classical example is when we call
174: two segments spanned by the given points equivalent if they
175: have the same length. Bounding the maximum size of an equivalence class
176: is the famous {\em repeated distances} problem
177: \cite{bmp-05,e-46,sst-84,s-97}, and bounding the minimum number of distinct
178: classes is the equally famous
179: {\em distinct distances} problem \cite{bmp-05,e-46,kt-04,st-01,s-97,t-03}.
180: In this paper, we make progress on several old extremal problems on
181: triangle areas in two and in three dimensions. We also study some
182: new and interesting variants never considered before. Our proof
183: techniques draw from a broad range of combinatorial tools such as
184: the Szemer\'edi-Trotter theorem on point-line incidences~\cite{st-83},
185: the Crossing Lemma~\cite{acns-82,l-84}, incidences between curves and
186: points and tangencies between curves and lines, extremal graph
187: theory~\cite{kst-54}, quasi-planar graphs~\cite{aapps-97},
188: Minkowski-type constructions, repeated distances on the
189: sphere~\cite{pa-95}, the partition technique of Clarkson~\etal~\cite{ceg-90},
190: various charging schemes, etc.
191: 
192: In 1967, A.~Oppenheim (see \cite{ep-95}) asked the following question:
193: Given $n$ points in the plane and $A>0$, how many triangles spanned
194: by the points can have area $A$? By applying an affine transformation,
195: one may assume $A=1$ and count the triangles of {\em unit} area.
196: Erd\H{o}s and Purdy~\cite{ep-71} showed that a $\sqrt{\log n}\times
197: (n/\sqrt{\log n})$ section of the integer lattice determines
198: $\Omega(n^2 \log\log{n})$ triangles of the same area. They also showed
199: that the maximum number of such triangles is at most $O(n^{5/2})$. In
200: 1992, Pach and Sharir~\cite{ps-92} improved the exponent and obtained
201: an $O(n^{7/3})$ upper bound using the Szemer\'edi-Trotter
202: theorem~\cite{st-83} on the number of point-line incidences. We
203: further improve the upper bound by estimating the number of incidences
204: between the points and a 4-parameter family of quadratic curves. We
205: show that $n$ points in the plane determine at most
206: $O(n^{44/19})=O(n^{2.3158})$ unit-area triangles. We also consider the
207: case of points in convex position, for which we construct $n$-element
208: point sets that span $\Omega(n\log n)$ triangles of unit area.
209: 
210: Bra\ss, Rote, and Swanepoel~\cite{brs-01} showed that $n$ points in
211: the plane determine at most $O(n^2)$ minimum-area triangles, and they
212: pointed out that this bound is asymptotically tight. We introduce a
213: simple charging scheme to first bring the upper bound down to
214: $n^2-n$ and then further to $\frac{2}{3}(n^2-n)$. Our charging scheme
215: is also instrumental in showing that a $\sqrt{n} \times \sqrt{n}$
216: section of the integer lattice spans $(6/\pi^2-o(1))n^2$ triangles of
217: minimum area.  In the lower bound constructions, there are many collinear
218: triples and most of the minimum-area triangles are obtuse. We show
219: that there are at most $O(n)$ {\em acute} triangles of minimum
220: (nonzero) area, for any $n$-element point set.
221: Also, we show that $n$ points in (strictly) 
222: convex position determine at most
223: $O(n)$ minimum-area triangles---these bounds are best possible apart from
224: the constant factors. If no three points are allowed to be collinear,
225: we construct $n$-element point sets that span $\Omega(n\log{n})$
226: triangles of minimum area.
227: 
228: Next we address analogous questions for triangles in 3-space.
229: The number of triangles with some extremal property
230: might go up (significantly) when one moves up one dimension.
231: For instance, Bra\ss, Rote, and Swanepoel~\cite{brs-01} have shown that
232: the number of maximum area triangles in the plane is at most $n$
233: (which is tight). In 3-space we show that this number is at
234: least $\Omega(n^{4/3})$ in the worst case.
235: In contrast, for minimum-area triangles, we
236: prove that the quadratic upper bound from the planar case remains in
237: effect for 3-space, with a different constant of proportionality.
238: 
239: As mentioned earlier, Erd\H{o}s and Purdy~\cite{ep-71} showed that a
240: suitable $n$-element section of the integer lattice determines
241: $\Omega(n^2 \log\log{n})$ triangles of the same area. Clearly, this
242: bound is also valid in 3-space. In the same paper, via a forbidden
243: graph argument applied to the incidence graph between points and
244: cylinders whose axes pass through the origin, Erd\H{o}s and Purdy
245: deduced an $O(n^{5/3})$ upper bound on the number of unit-area triangles
246: incident to a common point, and thereby an $O(n^{8/3})$ upper bound on
247: the number of unit-area triangles determined by $n$ points in 3-space.
248: Here, applying a careful (and somewhat involved) analysis of the
249: structure of point-cylinder incidences in $\RR^3$, we
250: prove a new upper bound of $O(n^{17/7}\beta(n))=O(n^{2.4286})$,
251: for $\beta(n) = \exp(\alpha(n)^{O(1)})$,
252: where $\alpha(n)$ is the
253: extremely slowly growing inverse Ackermann function.
254: 
255: It is conjectured~\cite{bmp-05,bp-79,eps-82} that $n$ points in
256: $\RR^3$, not all on a line, determine at least $\lfloor(n-1)/2\rfloor$
257: distinct triangle areas. This bound has recently been established in
258: the plane \cite{p-07}, but the question is still wide open in $\RR^3$. 
259: It is attained by $n$ equally spaced points distributed evenly on two
260: parallel lines (which is in fact a planar construction).
261: We obtain a first result on this question and show that
262: $n$ points in $\RR^3$, not all on a line, determine at least
263: $n^{2/3}\exp(-\alpha(n)^{O(1)})=\Omega(n^{.666})$ triangles of
264: distinct areas.  Moreover, all these triangles share a common side.
265: 
266: 
267: \section{Unit-area triangles in the plane}
268: \label{sec:unit2}
269: 
270: \noindent {\bf The general case.} We establish a new upper bound on the
271: maximum number of unit-area triangles determined by $n$ points the
272: plane. 
273: 
274: \begin{theorem}\label{thm:unit2}
275: The number of unit-area triangles spanned by $n$ points in the plane
276: is $O(n^{2+6/19})= O(n^{2.3158})$.
277: \end{theorem}
278: %
279: \begin{proof}
280: Let $S$ be a set of $n$ points in the plane. Consider a triangle
281: $\Delta{abc}$ spanned by $S$. We call the three lines containing the
282: three sides of $\Delta{abc}$, {\em base lines} of $\Delta$, and the
283: three lines parallel to the base lines and incident to the third
284: vertex, {\em top lines} of $\Delta$.
285: 
286: \cut{For any integer $k\in \NN$, let $L_k$ denote the set of $k$-rich
287:   lines (that is, the lines containing at least $k$ points of $S$). By
288:   the Szemer\'edi-Trotter theorem~\cite{st-83}, we have
289:   $|L_k|=O(n^2/k^3+n/k)$. In particular, $|L_k|=O(n^2/k^3)$ for $k\leq
290:   \sqrt{n}$ and $|L_k| = O(n/k)$ for $k\geq \sqrt{n}$.}
291: %
292: For a parameter $k$, $1\leq k\leq \sqrt{n}$, to be optimized
293: later, we partition the set of unit-area triangles as follows.
294: %\begin{itemize}
295: 
296: \smallskip
297: \noindent $\bullet$ $U_1$ denotes the set of unit-area triangles where
298: one of the top lines is incident to fewer than $k$ points of $S$.
299: 
300: \smallskip
301: \noindent $\bullet$ $U_2$ denotes the set of unit-area triangles where
302: all three top lines are {\em $k$-rich} (i.e., each contains at
303: least $k$\\
304: \indent points of $S$).
305: 
306: \smallskip
307: %\end{itemize}
308: \noindent We derive different upper bounds for each of these
309: types of unit-area triangles.
310: 
311: \paragraph{Bound for  $|U_1|$.}
312: For any two distinct points, $a,b\in \RR^2$, let $\ell_{ab}$ denote
313: the line through $a$ and $b$. The points $c$ for which the triangle
314: $\Delta{abc}$ has unit area lie on two lines
315: $\ell^-_{ab},\ell^+_{ab}$ parallel to $\ell_{ab}$ at distances
316: $2/|ab|$ on either side of $\ell_{ab}$. The ${n\choose 2}$ segments
317: determined by $S$ generate at most $2{n\choose 2}$ such lines
318: (counted with multiplicity).
319: If $\Delta abc \in U_1$ and its top line incident to the fewest
320: points of $S$ is $\ell_{ab}'\in \{\ell^-_{ab},\ell^+_{ab}\}$, then
321: $\ell_{ab}'$ is incident to at most $k$ points, so the segment $ab$ is
322: the base of at most $k$ triangles $\Delta{abc}\in U_1$ (with
323: $c\in\ell_{ab}'$).  This gives the upper bound
324: %
325: $$|U_1|\leq  2{n\choose 2} \cdot k = O(n^2 k).$$
326: 
327: \paragraph{Bound for $|U_2|$.}
328: Let $L$ be the set of $k$-rich lines, and let $m=|L|$. By the
329: Szemer\'edi-Trotter theorem~\cite{st-83}, we have $m=O(n^2/k^3)$ for
330: any $k\leq \sqrt{n}$. Furthermore, the cardinality of the set
331: $I(S,L)$ of point-line incidences between $S$ and $L$ is
332: $|I(S,L)|=O(n^2/k^2)$.
333: 
334: \old{
335: If $\Delta{abc}\in U_2$, then its top lines
336: $\ell_{ab}',\ell_{bc}',\ell_{ac}'\in L_k$ form a triangle of area 4,
337: in which the points $a,b,c$ are the midpoints of the three sides. We
338: give an upper bound on the number of 4-tuples
339: $(\ell_1,\ell_2,\ell_3,a)\in L_k\times S$ where the lines
340: $\ell_1,\ell_2,\ell_3$ form a triangle of area 4 and $a$ is the
341: midpoint of a side. Since every triangle in $U_2$ corresponds to
342: three 4-tuples of this form, we obtain an upper bound on $|U_2$.
343: } % old
344: 
345: For any pair of nonparallel lines $\ell_1,\ell_2\in L$, let
346: $\gamma(\ell_1,\ell_2)$ denote the locus of points $p\in \RR^2$,
347: $p\not\in\ell_1\cup \ell_2$, such that the parallelogram that has a
348: vertex at $p$ and two sides along $\ell_1$ and $\ell_2$,
349: respectively, has area 2. The set $\gamma(\ell_1,\ell_2)$ consists
350: of two hyperbolas with $\ell_1$ and $\ell_2$ as asymptotes.
351: See Figure~\ref{fig:hyper}. For instance, if
352: $\ell_1: y=0$ and $\ell_2: y= ax$, then $\gamma(\ell_1,\ell_2) =
353: \{(x,y)\in \RR^2:xy=y^2/a+2\}\cup \{(x,y)\in \RR^2:xy=y^2/a-2\}$.
354: Any two nonparallel lines uniquely determine two such hyperbolas.
355: Let $\Gamma$ denote the set of these hyperbolas. Note that
356: $|\Gamma|=O(m^2)$. The family of such hyperbolas for all pairs of
357: nonparallel lines form a 4-parameter family of quadratic curves
358: (where the parameters are the coefficients of the defining lines).
359: 
360: For any triangle $\Delta{abc}\in U_2$, any pair of its top lines,
361: say, $\ell_{ab}'$ and $\ell_{ac}'$, determine a hyperbola passing
362: through $a$, which is incident to the third top line $\ell_{bc}'$;
363: furthermore $\ell_{bc}'$ is tangent\footnote{%
364:   For a quick proof, let $\uu$ (resp., $\vv$) be a unit vector
365:   along $\ell'_{ac}$ (resp., $\ell'_{ab}$). The point $a$ can be
366:   parametrized as $\xx= t\uu + \frac{\kappa}{t}\vv$, where
367:   $\kappa=2/\sin\theta$, and $\theta$ is the angle between
368:   $\ell'_{ac}$ and $\ell'_{ab}$. Hence the tangent to the
369:   hyperbola at $a$ is given by
370:   $\dot{\xx} = \uu-\frac{\kappa}{t^2}\vv \,\|\, t\uu-\frac{\kappa}{t}\vv =
371:   \vec{cb}$.}
372: to the hyperbola at $a$. See Figure~\ref{fig:hyper}. Any hyperbola in
373: this 4-parameter family is uniquely determined by two incident
374: points and the two respective tangent lines at those points.
375: %\micha{Has anyone ever checked this carefully?}
376: %\csaba{There are two trivial quadratic curve incident to points $p$ and $q$,
377: %and tangent to lines $\ell_p$ and $\ell_q$ at those points:
378: %The union of lines $\ell_p\cup\ell_q$ (a degenerate hyperbola),
379: %and the line segment $pq$ (a degenerate ellipse). Any other quadratic curve
380: %is the linear combination of these two (since the incident points and
381: %the tangent lines impose linear constraints on the coefficients of
382: %the general quadratic curve $ax^2+by^2+cxy+dx+ey=1$. In this family
383: %of linear combinations, just one quadratic curve corresponds to
384: %unit triangles.
385: 
386: \begin{figure}[htbp]
387:   \begin{center}
388: \epsfig{file=hyper.eps,width=3.9cm,clip=}
389:  \caption{\small One of the hyperbolas defined by the triangle
390:  $\Delta abc$.
391: \label{fig:hyper}}
392: \end{center}
393:  \end{figure}
394: 
395: 
396: \old{If three lines $\ell_1,\ell_2,\ell_3\in L_k$ form a triangle
397: $\Delta$ of area 4, then $\ell_3$ is tangent to
398: $\gamma(\ell_1,\ell_3)$, and the intersection point
399: $\gamma(\ell_1,\ell_2)\cap \ell_3$ is the midpoint of the side of
400: $\Delta$ along $\ell_3$. Any hyperbola in the 4-parameter family
401: $\Gamma$ is uniquely determined by two incident points and two
402: tangent lines at those points.
403: } % old
404: 
405: We define a topological graph $G$ as follows. For each point
406: $p\in S$, which is incident to $d_p$ lines of $L$, we create $2d_p$
407: vertices in $G$, as follows (refer to Figure~\ref{fig:graph}).
408: Draw a circle
409: $C_\eps(p)$ centered at $p$ with a sufficiently small radius
410: $\eps>0$, and place a vertex at every intersection point of the
411: circle $C_\eps(p)$ with the $d_p$ lines incident to $p$. The number
412: of vertices is $v_G=2|I(S,L)|=O(n^2/k^2)$. Next, we define the edges
413: of $G$. For each connected branch $\gamma$ of every hyperbola in
414: $\Gamma$, consider the set $S(\gamma)$ of points $p\in S$ that are
415: (i) incident to $\gamma$ and (ii) some line of $L$ is tangent to
416: $\gamma$ at $p$. For any two consecutive points $p,q\in S(\gamma)$,
417: draw an edge along $\gamma$ between the two vertices of $G$ that
418: (i) correspond to the incidences $(p,\ell_p)$ and $(q,\ell_q)$, where
419: $\ell_p$ and $\ell_q$ are the tangents of $\gamma$ at $p$ and $q$,
420: respectively, and (ii) are closest to each other along $\gamma$.
421: Specifically, the edge follows $\gamma$ between the
422: circles $C_{2\eps}(p)$ and $C_{2\eps}(q)$ and follows straight line
423: segments in the interiors of those circles.
424: %\micha{I dont see why you need this double circle rule. Isnt it enough
425: %to use just the inner circles $C_\eps(p)$? What are you afraid of?}
426: %\csaba{With the inner circle we can give exactly the location
427: %of the new vertices. If we place these vertices on the outer circle,
428: %then the segments in the figure would be too close to each other.}
429: Choose $\eps>0$ sufficiently small so that the circles $C_{2\eps}(p)$
430: have disjoint interiors and
431: the portions of the hyperbolas in the interiors of the circles
432: $C_{2\eps}(p)$, for every $p\in S$, meet at $p$ only. This
433: guarantees that the edges of $G$ cross only at intersection points
434: of the hyperbolas. The graph $G$ is {\em simple} because two points and
435: two tangent lines uniquely determine a hyperbola in $\Gamma$. The
436: number of edges is at least $3|U_2|-2m^2$, since every triangle in
437: $U_2$ corresponds to three point-hyperbola incidences in $I(S,\Gamma)$
438: (satisfying the additional condition of tangency with the respective
439: top lines);
440: and along each of the $2m^2$ hyperbola branches, each of its
441: incidences with the points of $S$ (of the special kind under
442: consideration), except for one, contributes one edge to $G$.
443: %
444: \begin{figure}[htbp]
445:   \begin{center}
446: \epsfig{file=graph.eps,width=14cm,clip=}
447:  \caption{\small On the left: a point $p\in S$ incident to three lines
448: of $L$ (dashed) and 8 hyperbolas, each tangent to one of those
449: lines. On the right: the 6 vertices of $G$ corresponding to the 3
450: point-line incidences at $p$, and the drawings of the edges along the
451: hyperbolas.\label{fig:graph}}
452: \end{center}
453:  \end{figure}
454: %
455: Thus $G$ is a simple topological graph with $v_G=2I(S,L)=O(n^2/k^2)$
456: vertices and $e_G \geq 3|U_2|-2m^2$ edges. Since in this drawing of
457: $G$, every crossing is an intersection of two hyperbolas, the
458: crossing number of $G$ is upper bounded by ${\rm cr}(G)= O(|\Gamma|^2) =
459: O(m^4)$. We can also bound the crossing number of $G$ from below via
460: the Crossing Lemma of Ajtai~\etal~\cite{acns-82} and
461: Leighton~\cite{l-84}. It follows that
462: %
463: $$\Omega\left( \frac{e_G^3}{v_G^2}\right) -4v_G \leq {\rm cr}(G) \leq
464: O(m^4).$$
465: %
466: Rearranging this chain of inequalities, we obtain $e_G^3 = O(m^4v_G^2
467: + v_G^3)$, or $e_G= O(m^{4/3}v_G^{2/3} + v_G)$. Comparing this bound
468: with our lower bound $e_G\ge 3|U_2|-2m^2$, we have $|U_2|=
469: O(m^{4/3}v_G^{2/3} + v_G+ m^2)$. Hence, for $k\leq \sqrt{n}$, we
470: have
471: %
472: $$|U_2|= O\left( \left(\frac{n^2}{k^3}\right)^{4/3}
473:   \left(\frac{n^2}{k^2}\right)^{2/3} +\frac{n^2}{k^2} +
474:   \left(\frac{n^2}{k^3} \right)^2\right) =
475: O\left(\frac{n^4}{k^{16/3}}+\frac{n^2}{k^2}\right)=
476: O\left(\frac{n^4}{k^{16/3}}\right).$$
477: %
478: The total number of unit-area triangles is $|U_1|+|U_2| = O(n^2k+
479: n^4/k^{16/3})$. This expression is minimized for
480: $k=n^{6/19}$, and we get $|U_1|+|U_2| = O(n^{44/19})$.
481: \end{proof}
482: 
483: 
484: \subsection{Convex position}
485: 
486: The construction of Erd\H{o}s and Purdy~\cite{ep-71} with many triangles
487: of the same area, the $\sqrt{\log n}\times (n/\sqrt{\log n})$ section of
488: the integer lattice, also contains many collinear triples. Here we consider
489: the unit-area triangle problem in the special case of point sets in
490: strictly convex position, so no three points are collinear.
491: We show that $n$ points in convex position in the plane can determine a
492: superlinear number of unit-area triangles. On the other hand, we do not know
493: of any subquadratic upper bound.
494: 
495: \begin{theorem} \label{thm:convex-unit}
496: For all $n \geq 3$, there exist $n$-element point sets in convex
497: position in the plane that span $\Omega(n\log n)$ unit-area
498: triangles.
499: \end{theorem}
500: %
501: \begin{proof}
502: We recursively construct a set $S_i$ of $n_i=3^i$ points on the
503: unit circle that determine $t_i=i3^{i-1}$ unit-area triangles, for
504: $i=1,2,\ldots$.  Take a circle $C$ of unit radius centered
505: at the origin $o$. We start with a set $S_1$ of 3 points along the
506: circle forming a unit-area triangle, so we have $n_1=3$ points and
507: $t_1=1$ unit-area triangles. In each step, we triple the number of
508: points, i.e., $n_{i+1}=3n_i$, and create new unit-area triangles, so that
509: $t_{i+1}=3t_i+n_i$. This implies $n_i=3^i$, and $t_i= i 3^{i-1}$,
510: yielding the desired lower bound.
511: The $i$-th step, $i \geq 2$, goes as follows. Choose a
512: generic angle value $\alpha_i$, close to $\pi/2$, say,
513: and let $\beta_i$ be the angle such that the
514: three unit vectors at direction $0$, $\alpha_i$, and $\beta_i$ from
515: the origin determine a unit-area triangle, which we denote by $D_i$
516: (note that $\beta_i$ lies in the third quadrant).
517: Rotate $D_i$ around the origin to each position where its $0$ vertex
518: coincides with one of the $n_i$ points of $S_i$, and add the other
519: two vertices of $D_i$ in these positions
520: to the point set. (With appropriate choices of $S_1$ and the angles
521: $\alpha_i$, $\beta_i$, one can guarantee that no two points of any
522: $S_i$ coincide.) For each point of $S_i$, we
523: added two new points, so $n_{i+1}=3n_i$. Also, we have $n_i$ new unit-area
524: triangles from rotated copies of $D_i$; and each of the $t_i$
525: previous triangles have now two new copies rotated by $\alpha_i$ and
526: $\beta_i$. This gives $t_{i+1}=3t_i+n_i$.
527: \end{proof}
528: 
529: 
530: \section{Minimum-area triangles in the plane}
531: \label{sec:minimum2}
532: 
533: \paragraph{The general case.} We first present a simple but
534: effective charging scheme that gives an upper bound of $n^2-n$ on the
535: number of minimum (nonzero) area triangles spanned by $n$ points in
536: the plane (Lemma \ref{lem:min-area}).
537: This technique yields a very short proof of the minimum area result
538: from \cite{brs-01}, with a much better constant of proportionality.
539: Moreover, its higher-dimensional variants lead to asymptotically tight
540: bounds on the maximum number of minimum-volume $k$-dimensional
541: simplices in $\RR^d$, for any $1 \leq k \leq d$
542: (see Section \ref{sec:minimum3} for the case $k=2,d=3$, and
543: \cite{dt-07b} for the case $k=3,d=3$; the generalization to arbitrary
544: $1\leq k \leq d$ will be presented in the journal version of \cite{dt-07b}).
545: 
546: \begin{lemma} \label{lem:min-area}
547:  The number of triangles of minimum (nonzero)
548:  area spanned by $n$ points in the plane is at most $n^2-n$.
549: \end{lemma}
550: \begin{proof}
551: Consider a set $S$ of $n$ points in the plane. Assign every triangle
552: of minimum area to one of its longest sides. For a segment $ab$,
553: with $a,b\in S$, let $R_{ab}^+$ and $R_{ab}^-$ denote the two
554: rectangles of extents $|ab|$ and $2/|ab|$ with $ab$ as a common side.
555: If a minimum-area triangle $\Delta{abc}$ is assigned to $ab$, then
556: $c$ must lie in the relative interior of the side parallel to $ab$ in either
557: $R_{ab}^+$ or $R_{ab}^-$. If there were two points, $c_1$ and $c_2$,
558: on one of these sides, then the area of $\Delta{ac_1c_2}$ would be smaller
559: than that of $\Delta{abc}$, a contradiction. Therefore, at most two
560: triangles are assigned to each of the ${n \choose 2}$ segments
561: (at most one on each side of the segments), and so there are at
562: most $n^2-n$ minimum-area triangles.
563: \end{proof}
564: 
565: \smallskip
566: We now refine our analysis and establish a $\frac{2}{3}(n^2-n)$ upper bound, which
567: leaves only a small gap from our lower bound
568: $(\frac{6}{\pi^2}-o(1))n^2$; both bounds are presented in
569: Theorem~\ref{thm:min-area} below. Let us point out again
570: that here we allow collinear triples of points. The maximum number of
571: collinear triples is clearly ${n\choose 3}=\Theta(n^3)$.
572: The bounds below, however, consider only nondegenerate triangles of
573: {\em positive} areas.
574: 
575: % A segment $ab$ spanned by $S$ is a {\em visibility segment} if it
576: % does not contain any point of $S$ in its interior (see~\cite{kpw-05}
577: % for results on the graphs formed by visibility segments). Observe
578: % that every edge of a minimum area triangle spanned by $S$ must be a
579: % visibility segment. We deduce a better upper bound on the number of
580: % minimum area triangles by analyzing its connection to the number of
581: % visibility segments.
582: 
583: \begin{theorem} \label{thm:min-area}
584:  The number of triangles of minimum (nonzero)
585:  area spanned by $n$ points in the plane is at most
586:  $\frac{2}{3}(n^2-n)$.
587: The points in the $\lfloor\sqrt{n}\rfloor\times
588: \lfloor\sqrt{n}\rfloor$ integer grid span
589: $(\frac{6}{\pi^2}-o(1))n^2 \gtrapprox .6079 n^2$ minimum-area
590: triangles.
591: \end{theorem}
592: %
593: \begin{proof}
594: We start with the upper bound.
595: Consider a set $S$ of $n$ points in the plane, and let $L$ be the
596: set of connecting lines determined by $S$. Assume, without loss of
597: generality, that none of the lines in $L$ is vertical.
598: Let $T$ be the set of minimum (nonzero) area triangles spanned by $S$,
599: and put $t=|T|$.
600: There are $3t$ pairs $(ab,c)$ where $\Delta{abc}\in T$, and we may
601: assume, without loss of generality, that for at least half of these
602: pairs (i.e., for at least $\frac{3}{2}t$ pairs) $\Delta{abc}$ lies above
603: the line spanned by $a$ and $b$.
604: 
605: For each line $\ell\in L$, let $\ell'$ denote the line parallel to
606: $\ell$, lying above $\ell$, passing through some point(s) of $S$, and
607: closest to $\ell$ among these lines. Clearly, if $c\in S$
608: generates with $a,b\in\ell$ a minimum-area triangle which lies above
609: $ab$ then (i) $a$ and $b$ are a closest pair among the pairs of points
610: in $\ell\cap S$, and (ii) $c\in\ell_{ab}'$
611: (the converse does not necessarily hold).
612: 
613: Now fix a line $\ell\in L$; set $k_1=|\ell\cap S|\ge 2$, and
614: $k_2=|\ell'\cap S|\ge 1$, where $\ell'$ is as defined above.
615: The number of minimum-area triangles determined by a pair of points in
616: $\ell$ and lying above $\ell$ is at most $(k_1-1)k_2$. We have
617: %
618: \begin{equation} \label{eq:k12}
619: {k_1\choose 2} + {k_2\choose 2} \ge (k_1-1)k_2 .
620: \end{equation}
621: %
622: Indeed, multiplying by $2$ and subtracting the right-hand side from
623: the left-hand side gives
624: $$
625: k_1^2 -k_1 + k_2^2 -k_2 - 2k_1k_2 + 2k_2 =
626: (k_1-k_2)^2 - (k_1-k_2) \ge 0,
627: $$
628: which holds for any $k_1,k_2 \in \ZZ$.
629: 
630: We now sum (\ref{eq:k12}) over all lines $\ell\in L$. The sum of the
631: terms ${k_1\choose 2}$ is ${n\choose 2}$, and the sum of the terms
632: ${k_2\choose 2}$ is at most ${n\choose 2}$, because a line $\lambda\in L$
633: spanned by at least two points of $S$ can arise as the line $\ell'$ for at most
634: one line $\ell\in L$. Hence we obtain
635: $$
636: \frac{3}{2}t \le \sum_{\ell\in L} (k_1-1)k_2 \le 2{n\choose 2} =
637: n(n-1) ,
638: $$
639: thus $t\le \frac23(n^2-n)$, as asserted.
640: 
641: \smallskip
642: We now prove the lower bound. Consider the set $S$ of points in the
643: $\lfloor \sqrt{n}\rfloor \times \lfloor\sqrt{n}\rfloor$ section of the
644: integer lattice. Clearly $|S| \leq n$. The minimum nonzero area of
645: triangles in $S$ is $1/2$ (by Pick's theorem). Recall that the
646: charging scheme used in the proof of Lemma~\ref{lem:min-area}
647: assigns each triangle of minimum area to one of its
648: longest sides, which is necessarily a {\em visibility segment}
649: (a segment not containing any point of $S$ in its relative interior).
650: We show that every visibility segment $ab$ which is not
651: axis-parallel is assigned to exactly two triangles of minimum area.
652: 
653: Draw parallel lines to $ab$ through all points of the integer
654: lattice. Every line parallel to $ab$ and incident to a point of $S$
655: contains equally spaced points of the (infinite) integer lattice.
656: The distance between consecutive points along each line is exactly
657: $|ab|$.
658: % The gap between consecutive lines is at most $\sqrt{2}/2<1$.
659: This implies that each of the two lines parallel to $ab$ and closest
660: to it contains a lattice point on the side of the respective rectangle
661: $R_{ab}^-$ or $R_{ab}^+$,
662: opposite to $ab$, and this lattice point is in $S$.
663: Finally, observe that there are no empty acute triangles in the integer
664: lattice. It follows that our charging scheme uniquely assigns empty
665: triangles to visibility segments. An illustration is provided in
666: Figure~\ref{fig:grid}.
667: 
668: \begin{figure}[htbp]
669:   \begin{center}
670: \epsfig{file=grid.eps,width=2.5in,clip=}
671:  \caption{\small In an integer lattice section, every visibility segment
672:  which is not axis-parallel is the longest side of two triangles of
673:  minimum area.}
674: \end{center}
675: \label{fig:grid}
676: \end{figure}
677: 
678: A non-axis-parallel segment $ab$ is a visibility segment if
679: and only if the coordinates of the vector $\overrightarrow{ab}$ are
680: relatively prime. It is well known that $6/\pi^2$ is the limit of the
681: probability that a pair of integers $(i,j)$ with $ 1 \leq i,j \leq m$
682: are relatively prime, as $m$ tends to infinity \cite{v-54}.
683: Hence, a fraction of about
684: $6/\pi^2$ of the ${|S| \choose 2} \leq {n \choose 2}$ segments
685: spanned by $S$ are visibility segments which are not axis-parallel.
686: Each of these $(\frac{6}{\pi^2}-o(1)){n\choose 2}$ segments corresponds
687: to two unique triangles of minimum area, so $S$ determines at least
688: $(\frac{6}{\pi^2}-o(1))n^2$ minimum-area triangles.
689: \end{proof}
690: 
691: 
692: \subsection{Special cases}
693: 
694: In this subsection we consider some new variants of the minimum-area
695: triangle problem for the two special cases (i) where no three points are
696: collinear, and (ii) where the points are in convex position.
697: We also show that the maximum number of {\em acute} triangles
698: of minimum area, for any point set, is only linear.
699: 
700: \paragraph{Acute triangles.}
701: We have seen that $n$ points in an integer grid may span
702: $\Omega(n^2)$ triangles of minimum area. However, in that
703: construction, all these triangles are obtuse (or right-angled).
704: Here we prove that for any $n$-element point set in the plane, the
705: number of {\em acute} triangles of minimum area is only linear.
706:     This bound is attained in the following simple example.
707:     Take two groups of about $n/2$ equally spaced points on two parallel
708:     lines: the first group consist of the points
709:     $(i,0)$, for $i=0,\ldots,\lceil n/2 \rceil -1$, and the second
710:     group of the
711:     points $(i+1/2,\sqrt{3}/2)$, for $i=0,\ldots,\lfloor n/2 \rfloor -1$.
712:     This point set determines $n-2$ acute triangles of minimum area.
713: 
714: \begin{theorem} \label{thm:acute}
715: The maximum number of acute triangles of minimum area determined by $n$
716: points in the plane is $O(n)$. This bound is asymptotically tight.
717: \end{theorem}
718: %
719: \begin{proof}
720: Let $S$ be a set of $n$ points in the plane, and let $\T$ denote the
721: set of acute minimum-area triangles determined by $S$. Define a
722: geometric graph $G=(V,E)$ on $V=S$, where $uv \in E$ if and only if
723: $uv$ is a shortest side of a triangle in $\T$. We first argue that
724: every segment $uv$ is a shortest edge of at most two triangles in
725: $\T$, and then we complete the proof by showing that $G$ is planar
726: and so it has only $O(n)$ edges.
727: 
728: Let $\Delta{a_1 b_1 c_1} \in \T$ and assume that $b_1 c_1$ is a
729: shortest side of $\Delta{a_1 b_1 c_1}$.
730: % We show that $b_1 c_1$ is a shortest side of at most two acute
731: % triangles of minimum area.
732: Let $\Delta{a_2 b_2 c_2}$ be the triangle such that the midpoints of its
733: sides are $a_1,b_1,c_1$; and let $\Delta{a_3 b_3 c_3}$ be the
734: triangle such that the midpoints of its sides are $a_2,b_2,c_2$.
735: Refer to Figure~\ref{acute-obtuse}(a). Since $\Delta{a_1 b_1 c_1}$
736: has minimum area, then, in the notation of the figure,
737: each point of $S \setminus \{a_1,b_1,c_1\}$ lies in one of the
738: (closed) regions $R_1$ through $R_6$ or on one of the lines
739: $\ell_2$, $\ell_4$ or $\ell_5$; also, no point of $S \setminus
740: \{a_1,b_1,c_1\}$  lies in the interior of $\Delta{a_3 b_3 c_3}$.
741: Similarly, any point $a\in S$ of
742: a triangle $\Delta{ab_1c_1} \in \T$ must lie on $\ell_1$ or
743: $\ell_3$. Thus $a=a_1$ and $a=a_2$ are the only possible positions
744: of $a$. This follows from the fact that the triangles of $T$ are
745: acute: any point on, say, $\ell_1\cap\bd R_2$ or $\ell_1\cap\bd R_6$
746: forms an {\em obtuse} triangle with $b_1c_1$.
747: 
748: Consider two acute triangles $\Delta{a_1b_1c_1},\Delta{xyz} \in \T$
749: of minimum area with shortest sides $b_1 c_1\in E$ and $xy\in E$,
750: respectively. Assume that edges $b_1c_1$ and $xy$ cross each other.
751: %
752: We have the following four possibilities: (i) $x$ and $y$ lie in two
753: opposite regions $R_iR_{i+3}$, for some $i \in \{1,2,3\}$; (ii)
754: $x=a_1$ and $y \in R_4$; (iii) $x \in \ell_4$ and $y \in R_4$; (iv)
755: $x \in \ell_5$ and $y \in R_4$. Since $xy$ is a shortest side of
756: $\Delta{xyz}$, the distance from $z$ to the line through $x$ and $y$ is at
757: least $\sqrt{3}/2|xy|$. But then, in all four cases $\Delta{xyz}$
758: cannot be an acute triangle of minimum area, since it contains one
759: of the vertices of $\Delta{a_1 b_1 c_1}$ in its interior, a
760: contradiction. (For instance if $x \in R_1$ and $y \in R_4$,
761: $\Delta{xyc_1}$ would be obtuse and $\Delta{xyz}$ contains $c_1$ in
762: its interior, or if $x=a_1$ and $y \in R_4$, $\Delta{xyz}$ contains
763: either $b_1$ or $c_1$ in its interior.)
764: \end{proof}
765: 
766: %
767: \begin{figure}[htbp]
768: \centerline{\epsfxsize=3in \epsffile{acute.eps}
769: \hskip 0.3in
770: \epsfxsize=3in \epsffile{obtuse.eps}}
771: \centerline{\hfill (a)\hskip 2.4in (b)\hfill}
772: \caption{\small (a) Acute triangles: the graph $G$ is planar.
773: (b) Convex position: the graph $G$ is quasi-planar.}
774: \label{acute-obtuse}
775: \end{figure}
776: %
777: 
778: 
779: \paragraph{Convex position.}
780: For points in strictly convex position we prove a tight $\Theta(n)$
781: bound on the maximum possible number of minimum-area triangles.
782: Note that a regular $n$-gon has $n$ such triangles,
783: so it remains to show an $O(n)$ upper bound. Also,
784: $n$ points equally distributed on two parallel
785: lines (at equal distances) give a well-known quadratic lower bound, 
786: so the requirement that the points be in strictly convex position
787: is essential for the bound to hold.
788: 
789: \begin{theorem} \label{thm:convex-min}
790: The maximum number of minimum-area triangles determined by $n$ points in
791: (strictly) convex position in the plane is $O(n)$. 
792: This bound is asymptotically tight.
793: \end{theorem}
794: %
795: \begin{proof}
796: The argument below is similar to that in the proof of
797: Theorem~\ref{thm:acute}. Since there can be only $O(n)$
798: acute triangles of minimum area, it is sufficient to consider
799: right-angled and obtuse triangles (for simplicity,
800: we refer to both types as obtuse), even though
801: the argument also works for acute triangles. We use a
802: similar notation: now $\T$ denotes the set of obtuse triangles of
803: minimum area. We define a geometric graph $G=(V,E)$ on $V=S$, where
804: $uv \in E$ if and only if $uv$ is a shortest side of a triangle in
805: $\T$. See Figure \ref{acute-obtuse}(b).
806: 
807: Let $\Delta{a_1 b_1 c_1} \in \T$ with $b_1c_1$ a shortest side. By
808: convexity, at most four triangles in $\T$ can have a common shortest
809: side $b_1c_1$: at most two such triangles have a third vertex on
810: $\ell_1$ and at most another two of them have a third vertex on $\ell_3$.
811: A graph drawn in the plane is said to be {\em quasi-planar}
812: if it has no three edges which are pairwise crossing; it is
813: known~\cite{aapps-97} (see also \cite{at-07})
814: that any quasi-planar graph with $n$ vertices
815: has at most $O(n)$ edges.
816: We now show that $G$ is quasi-planar, which will complete the proof
817: of the theorem.
818: 
819: Consider the triangles
820: $\Delta{a_2 b_2 c_2}$ and $\Delta{a_3 b_3 c_3}$, defined as in the
821: proof of Theorem~\ref{thm:acute}.
822: Each point of $S \setminus \{a_1,b_1,c_1\}$ lies in one of the
823: (closed) regions $R_1$ through $R_6$;
824: in particular no such point
825: lies in the interior of $\Delta{a_3 b_3 c_3}$. 
826: (Here, unlike the previous analysis, strict convexity rules out points
827: on any of the three middle lines, such as $\ell_2$.)
828: In addition, by
829: convexity, the regions $R_1$, $R_3$ and $R_5$ are empty of points.
830: Assume now that $b_1 c_1$, $xy$, $uv$ form a triplet of pairwise
831: crossing edges, where $xy$ and $uv$ are distinct shortest sides of
832: two triangles $\Delta{xyz} \in \T$ and $\Delta{uvw} \in \T$. 
833: It follows that each of the two edges $xy$ and $uv$ must have one
834: endpoint at $a_1$ and the other in $R_4$ (since each crosses $b_1 c_1$). 
835: Thus two edges in this triplet have  a common
836: endpoint, and so they do not cross, which is a contradiction.
837: \end{proof}
838: 
839: \paragraph{No three collinear points.}
840: We conjecture that if no three points are collinear, then the maximum
841: number of triangles of minimum area is close to linear. It is not
842: linear, though: It has been proved recently~\cite{dpt-05} that there
843: exist $n$-element point sets
844: in the plane that span $\Omega(n \log{n})$ empty congruent triangles.
845: Here, we show that one can repeat this construction such that there is
846: no collinear triples of points and that the $\Omega(n \log{n})$ empty
847: congruent triangles have minimum (nonzero) area.
848: However, we do not know of any sub-quadratic upper bound.
849: 
850: 
851: \begin{theorem} \label{thm:general}
852:   For all $n \geq 3$, there exist $n$-element point sets in the plane that
853:   have no three collinear points and span $\Omega(n\log n)$ triangles of
854:   minimum (nonzero) area.
855: \end{theorem}
856: %
857: \begin{proof}
858: The construction is essentially the one given in \cite{dpt-05}, and
859: we provide here only a brief description. We then specify the
860: additional modifications needed for our purposes. First, a point set
861: $S$ is constructed with many, i.e., $\Omega(n \log{n})$, pairwise
862: congruent triples of collinear points, which can be also viewed as
863: degenerate empty congruent triangles. Then 
864: this construction is slightly perturbed to obtain a set of points $S$
865: with no collinear triples, so that these degenerate triangles become
866: non-degenerate empty congruent triangles of minimum (nonzero) area. 
867: The details are as follows (see~\cite{dpt-05}).
868: 
869: Let $n=3^k$ for some $k\in \NN$.
870: Consider $k$ unit vectors $b_1,\ldots,b_k$, and for
871: $1\le i\le k$, let $\beta_i$ be the counterclockwise angle from the
872: $x$-axis to $b_i$.
873: % We choose the angles $\beta_i$ independently and
874: % uniformly at random from the interval $(0, \pi/{2})$.
875: Let $\lambda \in (0,1)$ be fixed and let $a_i=\lambda b_i$. Consider
876: now all $3^k$ possible sums of these $2k$ vectors, $a_i$ and $b_i$,
877: $1\le i\le k$,
878: with coefficients $0$ or $1$, satisfying the condition that for each
879: $i$, at least one of $a_i$ or $b_i$ has coefficient $0$.
880: Let $S$ be the set of $3^k$ points determined by these vectors.
881: Clearly, each triple of the form
882: ($v$, $v+a_i$, $v+b_i$), where $v$ is a subset sum that does not
883: involve $a_i$ or $b_i$, consists of collinear points.
884: For such a triple, denote by $s_i(v)$ the segment whose endpoints are
885: $v$ and $v+b_i$.  We say that the collinear triple $(v, v+a_i, v+b_i)$
886: is of type $i$, $i=1,\ldots,k$.
887: For each $i$ there are exactly $3^{k-1}$
888: triples of type $i$, therefore a total of $ k 3^{k-1} = (n
889: \log{n})/(3 \log{3}) = \Omega(n \log{n}) $ triples of collinear
890: points. Clearly, all these triples form degenerate congruent
891: triangles in $S$.
892: Denote by $\ell_i(v)$ the line supporting the segment $s_i(v)$,
893: and by $\L$ the set of lines corresponding to these triples.
894: 
895: We need the following slightly stronger version of Lemma 1
896: in~\cite{dpt-05}. The proof is very similar to the proof of
897: Proposition 1 in~\cite{dpt-05}, and we omit the details.
898: 
899: \begin{lemma} \label{lem:exist}
900: There exist angles $\beta_1,\ldots,\beta_k$,
901: and $\lambda \in (0,1)$, such that
902: {\rm (i)} $S$ consists of $n$ distinct points;
903: {\rm (ii)} if $u,v,w \in S$ are collinear (in this order), then
904: $v=u+a_i$ and $w=u+b_i$.
905: \end{lemma}
906: 
907: Let $\eps$ be the minimum distance between points
908: $p \in S \setminus \{v, v+a_i, v+b_i\}$
909: and lines $\ell_i(v) \in \L$, over all pairs $(v,i)$.
910: By Lemma~\ref{lem:exist}, we have $\eps>0$. Now instead of choosing
911: $a_i$ to be collinear with $b_i$, slightly rotate $\lambda b_i$
912: counterclockwise from $b_i$ through a sufficiently small
913: angle $\delta$
914: about their common origin, so the collinearity disappears.
915: This modification is carried out at the
916: same time for all vectors $a_i$, $i=1,\ldots,k$,
917: that participate in the construction.
918: By continuity, there exists a sufficiently small $\delta=\delta(\eps)>0$,
919: so that (i) each of the triangles $\Delta(v, v+a_i, v+b_i)$
920: remains empty throughout this small perturbation, (ii) the point set
921: $S$ is in general position after the perturbation, and (iii) the
922: congruent triangles $\Delta(v, v+a_i, v+b_i)$ have minimum area.
923: This completes the proof.
924: \end{proof}
925: 
926: \vspace{-\baselineskip}
927: 
928: \section{Unit-area triangles in 3-space}
929: \label{sec:unit3}
930: 
931: Erd\H{o}s and Purdy~\cite{ep-71} showed that a $\sqrt{\log n}\times
932: (n/\sqrt{\log n})$ section of the integer lattice determines
933: \linebreak $\Omega(n^2 \log\log{n})$ triangles of the same area.
934: Clearly, this bound is also valid in 3-space. They have also
935: derived an upper bound of $O(n^{8/3})$ on the number of unit-area
936: triangles in $\RR^3$. Here we improve this bound to 
937: $O(n^{17/7}\beta(n)) = O(n^{2.4286})$.
938: We use $\beta(n)$ to denote any function of the form
939: $\exp(\alpha(n)^{O(1)})$, where $\alpha(n)$ is the extremely slowly
940: growing inverse Ackermann function. Any such function $\beta(n)$ is also
941: extremely slowly growing.
942: %In this section we obtain the following upper bound:
943: 
944: \begin{theorem}\label{thm:unit3}
945:   The number of unit-area triangles spanned by $n$ points in $\RR^3$
946:   is $O(n^{17/7}\beta(n)) = O(n^{2.4286})$.
947: \end{theorem}
948: 
949: The proof of the theorem is quite long, and involves several technical steps.
950: Let $S$ be a set of $n$ points in $\RR^3$. For each pair $a,b$ of
951: distinct points in $S$, let $\ell_{ab}$ denote the line passing
952: through $a$ and $b$, and let $C_{ab}$ denote the cylinder whose axis
953: is $\ell_{ab}$ and whose radius is $2/|ab|$. Clearly, any point $c\in
954: S$ that forms with $ab$ a unit-area triangle, must lie on $C_{ab}$.
955: The problem is thus to bound the number of incidences between
956: ${n\choose 2}$ cylinders and $n$ points, but it is complicated for two
957: reasons: (i) The cylinders need not be distinct.  (ii) Many distinct
958: cylinders can share a common generator line, which may contain many
959: points of $S$.
960: 
961: \vspace{-1mm}
962: \paragraph{Cylinders with large multiplicity.}
963: Let $\C$ denote the multiset of the ${n\choose 2}$ cylinders
964: $C_{ab}$, for $a,b\in S$. Since the cylinders in $\C$ may appear
965: with multiplicity, we fix a parameter $\mu=2^j$, $j=0,1,\ldots$,
966: and consider separately incidences with each of the sets $\C_\mu$,
967: of all the cylinders whose multiplicity
968: is between $\mu$ and $2\mu-1$. Write $c_\mu=|\C_\mu|$.
969: We regard $\C_\mu$ as a set (of distinct cylinders), and will multiply
970: the bound that we get for the cylinders in $\C_\mu$ by $2\mu$,
971: to get an upper bound on the number of
972: incidences that we seek to estimate. We will then sum up the resulting
973: bounds over $\mu$ to get an overall bound.
974: 
975: Let $C$ be a cylinder in $\C_\mu$. Then its axis $\ell$ must contain
976: $\mu$ pairs of points of $P$ at a fixed distance apart (equal to $2/r$,
977: where $r$ is the radius of $C$). That is, $\ell$ contains
978: $t>\mu$ points of $S$. Let us now fix $t$ to be a power of $2$, and
979: consider the subset $\C_{\mu,t} \subset \C_\mu$ of those cylinders in $\C_\mu$
980: that have at least $t$ and at most $2t-1$ points on their axis.
981: By the Szemer\'edi-Trotter Theorem~\cite{st-83} (or, rather, its
982: obvious extension to 3-space), the number of lines containing at
983: least $t$ %(and at most $2t-1$)
984: points of $S$ is $O(n^2/t^3 + n/t)$.
985: Any such line $\ell$ can be the axis of many cylinders in $\C_\mu$
986: (of different radii).  Any such cylinder ``charges'' $\Theta(\mu)$
987: pairs of points out of the $O(t^2)$ pairs along $\ell$,
988: and no pair is charged more than once.
989: Hence, for a given line $\ell$ incident to at least $t>\mu$ and at
990: most $2t-1$ points of $S$, the number of distinct cylinders in
991: $\C_\mu$ that have $\ell$ as axis is $O(t^2/\mu)$. Summing over
992: all axes incident to at least $t$ and at most $2t-1$ points
993: yields that the number of distinct cylinders in $\C_{\mu,t}$ is
994: %
995: \begin{equation}\label{nut}
996: c_{\mu,t}=
997: O\left(\left(\frac{n^2}{t^3} + \frac{n}{t}\right)\frac{t^2}{\mu}\right)=
998: O\left(\frac{n^2}{t\mu} + \frac{nt}{\mu}\right).
999: \end{equation}
1000: %
1001: We next sum this over $t$, a power of 2 between $\mu$ and $\nu$,
1002: and conclude that the number of distinct cylinders
1003: in $\C_\mu$ having at most $\nu$ points on their axis is
1004: %
1005: \begin{equation} \label{cmu}
1006: c_{\mu,\le\nu} = O\left( \frac{n^2}{\mu^2} + \frac{n \nu}{\mu} \right).
1007: \end{equation}
1008: %
1009: 
1010: \vspace{-1mm}
1011: \paragraph{Restricted incidences between points and cylinders.}
1012: We distinguish two {\em type}s of incidences, which we count
1013: separately. An incidence between a point $p$ and a cylinder $C$ is of
1014: {\em type 1} if the generator of $C$ passing through $p$ contains at
1015: least one additional point of $S$; otherwise it is of {\em type 2}. We
1016: begin with the following subproblem, in which we bound the number of
1017: incidences between the cylinders of $\C$, counted with multiplicity,
1018: and {\em multiple} points that lie on their generator lines, as well
1019: as incidences with cylinders with ``rich'' axes.
1020: Specifically, we have the following lemma.
1021: %
1022: \begin{lemma} \label{107}
1023:   Let $S$ be a set of $n$ points and $\C$ be the multiset of the
1024:   ${n\choose 2}$ cylinders $C_{ab}$, for $a,b\in S$ (counted with
1025:   multiplicity). The
1026:   total number of all incidences of type 1 and all incidences
1027:   involving cylinders having at least $n^{14/45}$ points on their axis
1028:   is bounded by $O(n^{107/45}\polylog(n))=O(n^{2.378})$.
1029: \end{lemma}
1030: %
1031: \begin{proof}
1032: Let $L$ denote the set of lines spanned by the points of $S$. Fix
1033: a parameter $k=2^i$, $i=1,\ldots$,
1034: and consider the set $L_k$ of all lines that contain at least $k$ and
1035: at most $2k-1$ points of $S$.
1036: We bound the number of incidences between
1037: cylinders in $\C$ that contain lines in $L_k$ as generators
1038: and points that lie on those lines. Formally, we bound the
1039: number of triples $(p,\ell,C)$, where $p\in S$, $\ell\in L_k$,
1040: and $C\in\C$, such that $p\in\ell$ and $\ell\subset C$.
1041: Summing these bounds over $k$ will give us a bound for the
1042: number of incidences of type 1. Along the way, we will also dispose of
1043: incidences with cylinders whose axes contain many points.
1044: 
1045: As already noted, the Szemer\'edi-Trotter Theorem~\cite{st-83} implies that
1046: ${\displaystyle
1047: \lambda_k:=|L_k|=O\left( \frac{n^2}{k^3} + \frac{n}{k} \right)}$.
1048: 
1049: \vspace{-1mm}
1050: \paragraph{Line-cylinder incidences.}
1051: Consider the subproblem of bounding the number of
1052: incidences between lines in $L_k$ and cylinders in $\C$, where
1053: a line $\ell$ is said to be incident to cylinder $C$ if $\ell$
1054: is a generator of $C$. We will then multiply the resulting
1055: bound by $2k$ to get an upper bound on the number of
1056: point-line-cylinder incidences involving $L_k$, and then
1057: sum the resulting bounds over $k$.
1058: 
1059: \vspace{-1mm}
1060: \paragraph{Generator lines with many points.}
1061: Let us first dispose of the case $k>n^{1/3}$. Any line
1062: $\ell\in L_k$ can be a generator of at most $n$ cylinders (counted
1063: with multiplicity), because, having fixed $a\in S$, the point
1064: $b\in S$ such that $C_{ab}$ contains $\ell$ is determined
1065: (up to multiplicity $2$). Hence the number of incidences between
1066: the points that lie on $\ell$ and the cylinders of $\C$ is
1067: $O(nk)$. Summing over $k=2^i>n^{1/3}$ yields the overall bound
1068: $$
1069: O\left( \sum_k nk\lambda_k \right) =
1070: O\left( \sum_k \left( \frac{n^3}{k^2} + n^2 \right) \right) = O(n^{7/3}) .
1071: $$
1072: Hence, in what follows, we may assume that $k\le n^{1/3}$.
1073: In this range of $k$ we have
1074: \begin{equation} \label{lambdak}
1075: \lambda_k=O\left( \frac{n^2}{k^3} \right) .
1076: \end{equation}
1077: 
1078: \vspace{-1mm}
1079: \paragraph{Axes with many points.}
1080: Let us also fix the multiplicity $\mu$ of the cylinders under
1081: consideration (up to a factor of $2$, as above).
1082: The number of distinct cylinders in $\C_\mu$
1083: having between $t>\mu$ and $2t-1$ points on their axes, is
1084: $O(n^2/(t\mu) + nt/\mu)$; see (\ref{nut}). While the first term
1085: is sufficiently small for our purpose, the second
1086: term may be too large when $t$ is large. To avoid this difficulty, we
1087: fix another threshold exponent $z < 1/2$ that we will optimize later,
1088: and handle separately the cases $t\ge n^{z}$ and $t< n^{z}$.
1089: That is, in the first case, for $t\ge n^{z}$ a power of 2, we seek an
1090: upper bound on the overall number of incidences between the points of
1091: $S$ and the cylinders in $\C$ whose axis contains between $t$ and $2t-1$
1092: points of $S$. (For this case, we combine all the multiplicities
1093: $\mu<t$ together.) By the Szemer\'edi-Trotter theorem~\cite{st-83},
1094: the number of such axes is $O(n^2/t^3+n/t)$.
1095: 
1096: Fix such an axis $\alpha$. It defines $\Theta(t^2)$ cylinders,
1097: and the multiplicity of any of these cylinders is at most $O(t)$. Since no
1098: two distinct cylinders in this collection can pass through the same
1099: point of $S$, it follows that the total number of incidences
1100: between the points of $S$ and these cylinders is $O(nt)$.
1101: Hence the overall number of incidences under consideration is
1102: $O(n^2/t^3+n/t) \cdot O(nt)= O(n^3/t^2+n^2)$.
1103: Summing over all $t\ge n^z$,
1104: a power of $2$, we get the overall bound $O(n^{3-2z})$.
1105: 
1106: Note that this bound takes care of {\em all} the incidences between
1107: the points of $S$ and the cylinders having at least $t\ge n^z$ points
1108: along their axes, not just those of type 1 (involving multiple points
1109: on generator lines).
1110: 
1111: %\paragraph{Incidences with low multiplicity.}
1112: \paragraph{Cylinders with low multiplicity.}
1113: We now confine the analysis to cylinders having fewer than $n^z$
1114: points on their axis, and go back to fixing the multiplicity $\mu$,
1115: which we may assume to be at most $n^z$.
1116: We thus want to bound the number of
1117: incidences between $\lambda_k$ distinct lines and $c_{\mu,\le n^z}$
1118: distinct cylinders in $\C_\mu$, for given $k\le n^{1/3}$, $\mu \leq n^z$.
1119: Note that a cylinder can contain a line if and only if it is parallel to the
1120: axis of the cylinder, so we can split the
1121: problem into subproblems, each associated with some direction
1122: $\theta$, so that in the $\theta$-subproblem we have a set of some
1123: $c_\mu^{(\theta)}$ cylinders and a set of some
1124: $\lambda_k^{(\theta)}$ lines, so that the lines and the cylinder axes
1125: are all parallel (and have direction $\theta$);
1126: we have $\sum_\theta c_\mu^{(\theta)} = c_{\mu,\le n^z}$,
1127: and $\sum_\theta \lambda_k^{(\theta)} = \lambda_k$.
1128: 
1129: For a fixed $\theta$, we project the cylinders and lines in the
1130: $\theta$-subproblem onto a plane with normal direction $\theta$,
1131: and obtain a set of $c_\mu^{(\theta)}$ circles and a set of
1132: $\lambda_k^{(\theta)}$ points, so that the number of line-cylinder
1133: incidences is equal to the number of point-circle incidences.
1134: By \cite{ANPPSS,ArS,MT},\footnote{%
1135:   The bound that we use, from \cite{MT}, is slightly better than
1136:   the previous ones.}
1137: the number of point-circle incidences between $N$ points and $M$ circles in the plane is
1138: $O(N^{2/3}M^{2/3} + N^{6/11}M^{9/11}\log^{2/11}(N^3/M)+N+M)$.
1139: It follows that the number of such line-cylinder incidences is
1140: \begin{equation} \label{eq:theta}
1141: O\left(
1142: (\lambda_k^{(\theta)})^{2/3}
1143: (c_\mu^{(\theta)})^{2/3} +
1144: (\lambda_k^{(\theta)})^{6/11}
1145: (c_\mu^{(\theta)})^{9/11}
1146:   \log^{2/11}((\lambda_k^{(\theta)})^3/c_\mu^{(\theta)})+
1147: \lambda_k^{(\theta)} +
1148: c_\mu^{(\theta)} \right) .
1149: \end{equation}
1150: Note that, for any fixed $\theta$, we have
1151: $\lambda_k^{(\theta)} \le n/k$ and
1152: $c_\mu^{(\theta)} \le n^{1+z}/\mu$.
1153: The former inequality is trivial.  To see the latter inequality,
1154: note that an axis with $t<n^z$ points
1155: defines ${t\choose 2}$ cylinders. Since we only consider cylinders
1156: with multiplicity $\Theta(\mu)$, the number of distinct such
1157: cylinders is $O(t^2/\mu)$, and the number of lines
1158: (of direction $\theta$) with about $t$
1159: points on them is at most $n/t$, for a total of at most $O(nt/\mu)$
1160: distinct cylinders. Partitioning the range $\mu < t\le n^{z}$ by
1161: powers of $2$, as above, and summing up the resulting bounds, the bound
1162: $c_\mu^{(\theta)} \le n^{1+z}/\mu$ follows.
1163: 
1164: Summing over $\theta$, and using H\"older's inequality, we have
1165: (here $x$ is a parameter between $2/11$ and $6/11$
1166: that we will fix shortly)
1167: $$
1168: \sum_\theta
1169: (\lambda_k^{(\theta)})^{6/11}
1170: (c_\mu^{(\theta)})^{9/11} \le
1171: \left(\frac{n}{k}\right)^{6/11-x}
1172: \left(\frac{n^{1+z}}{\mu}\right)^{x-2/11}
1173: \sum_\theta
1174: (\lambda_k^{(\theta)})^{x}
1175: (c_\mu^{(\theta)})^{1-x} \le
1176: $$
1177: $$
1178: \frac{n^{(4-2z)/11+xz}}{k^{6/11-x}\mu^{x-2/11}}
1179: \left( \sum_\theta
1180: \lambda_k^{(\theta)} \right)^{x}
1181: \left( \sum_\theta
1182: c_\mu^{(\theta)} \right)^{1-x} =
1183: \frac{n^{(4-2z)/11+xz}}{k^{6/11-x}\mu^{x-2/11}}
1184: \lambda_k^x c_{\mu,\le n^z}^{1-x} .
1185: $$
1186: We need to multiply this bound by $\Theta(k\mu)$.
1187: Substituting the bounds $\lambda_k=O(n^2/k^3)$ from (\ref{lambdak}),
1188: and $c_{\mu,\le n^z} = O(n^2/\mu^2 + n^{1+z}/\mu)$ from (\ref{cmu}),
1189: we get the bound
1190: \begin{eqnarray*}
1191: && O\left(
1192: n^{(4-2z)/11+xz} k^{5/11+x}\mu^{13/11-x}
1193: \left( \frac{n^2}{k^3} \right)^x
1194: \left( \frac{n^2}{\mu^2} + \frac{n^{1+z}}{\mu} \right)^{1-x}
1195: \log^{2/11} n\right)\\
1196: &=&O\left( k^{5/11-2x} \left(
1197: n^{2+(4-2z)/11+xz} \mu^{x-9/11} +
1198: n^{(15+9z)/11+x} \mu^{2/11} \right)
1199: \log^{2/11} n \right) .
1200: \end{eqnarray*}
1201: Choosing $x=5/22$ (the smallest value for which the exponent of $k$
1202: is non-positive), the first term becomes
1203: $O( n^{2+4/11+z/22} \log^{2/11} n)$, which we need to balance with
1204: $O(n^{3-2z})$; for this, we choose $z=14/45$ and obtain the bound
1205: $O(n^{107/45}\log^{2/11}n)=O(n^{2.378})$; for this choice of $z$,
1206: recalling that $\mu<n^z$, the second term is dominated by the first.
1207: Summing over $k,\mu$ only adds logarithmic factors, for a resulting
1208: overall bound $O(n^{2.378})$.
1209: 
1210: Similarly, we have (with a different choice of $x$, soon to be made)
1211: $$
1212: \sum_\theta
1213: (\lambda_k^{(\theta)})^{2/3}
1214: (c_\mu^{(\theta)})^{2/3} \le
1215: \left(\frac{n}{k}\right)^{2/3-x}
1216: \left(\frac{n^{1+z}}{\mu}\right)^{x-1/3}
1217: \sum_\theta
1218: (\lambda_k^{(\theta)})^{x}
1219: (c_\mu^{(\theta)})^{1-x} \le
1220: $$
1221: $$
1222: \frac{n^{(1-z)/3+xz}}{k^{2/3-x}\mu^{x-1/3}}
1223: \left( \sum_\theta
1224: \lambda_k^{(\theta)} \right)^{x}
1225: \left( \sum_\theta
1226: c_\mu^{(\theta)} \right)^{1-x} =
1227: \frac{n^{(1-z)/3+xz}}{k^{2/3-x}\mu^{x-1/3}}
1228: \lambda_k^x c_{\mu,\le n^z}^{1-x} .
1229: $$
1230: Multiplying by $k\mu$ and arguing as above, we get
1231: \begin{eqnarray*}
1232: &&O\left(
1233: n^{(1-z)/3+xz} k^{1/3+x}\mu^{4/3-x}
1234: \left( \frac{n^2}{k^3} \right)^x
1235: \left( \frac{n^2}{\mu^2} + \frac{n^{1+z}}{\mu} \right)^{1-x}
1236: \log^{2/11} n\right)\\
1237: &=&
1238: O\left( k^{1/3-2x} \left(
1239: n^{2+(1-z)/3+xz} \mu^{x-2/3} +
1240: n^{1+(1+2z)/3+x} \mu^{1/3} \right)
1241: \log^{2/11} n\right) .
1242: \end{eqnarray*}
1243: We choose here $x=1/6$ and note that, for $z=14/45$ and $\mu < n^z$,
1244: the bound is smaller than $O(n^{7/3})$, which is dominated by the
1245: preceding bound $O(n^{2.378})$.
1246: 
1247: Finally, the linear terms in (\ref{eq:theta}), multiplied by $k\mu$,
1248: add up to 
1249: $$
1250: k\mu \sum_{\theta} O\left(\lambda_k^{(\theta)}+c_\mu^{(\theta)}\right)=
1251: O\left( k\mu \left( \lambda_k + c_{\mu,\le n^z} \right)\right) =
1252: O\left( \frac{n^2\mu}{k^2} + \frac{n^2k}{\mu} + n^{1+z}k \right) ,
1253: $$
1254: which, by our assumptions on $k$, $\mu$, and $z$ is also dominated by
1255: $O(n^{2.378})$. Summing over $k,\mu$ only add logarithmic factors, for
1256: a resulting overall bound $O(n^{2.378})$.
1257: This completes the proof of Lemma~\ref{107}.
1258: \end{proof}
1259: 
1260: It therefore remains to count point-cylinder incidences of type 2,
1261: involving cylinders having at most $n^{14/45}$ points on their axes.
1262: 
1263: \vspace{-1mm}
1264: \paragraph{The intersection pattern of three cylinders.}
1265: We need the following technical lemma, whose proof is borrowed from a
1266: yet unpublished work \cite{EPS:cyl}, and is presented in the appendix.
1267: 
1268: \begin{lemma}\label{3int}
1269:   Let $C,C_1,C_2$ be three cylinders with no pair of parallel axes.
1270:   Then $C\cap C_1\cap C_2$ consists of at most 8 points.
1271: \end{lemma}
1272: 
1273: \paragraph{Point-cylinder incidences.}
1274: Using the partition technique~\cite{ch-05,ps-04}
1275: %\micha{Just making sure that this is the right REF?}
1276: %\csaba{This is a handbook summary of the partition technique.
1277: % For such a general method, it is better to cite a handbook, than research papers.}
1278: for disjoint cylinders in $\RR^3$, we show the following:
1279: 
1280: \begin{lemma}\label{lem:cut}
1281:   For any parameter $r$, $1\leq r\leq \min\{m,n^{1/3}\}$,
1282:   the maximum number of incidences
1283:   of type 2 between $n$ points and $m$ cylinders in 3-space satisfies
1284:   the following recurrence:
1285: \begin{equation}\label{eq:inci}
1286:   I(n,m)=O(n+mr^2\beta(r)) + O(r^3\beta(r))\cdot
1287:   I\left(\frac{n}{r^3},\frac{m}{r}\right),
1288: \end{equation}
1289: for some slowly growing function $\beta(n)$, as above.
1290: \end{lemma}
1291: %
1292: \begin{proof}
1293: Let $\C$ be a set of $m$ cylinders, and $S$ be a set of $n$ points.
1294: Construct a $(1/r)$-cutting of the arrangement $\A(\C)$.  The
1295: cutting has $O(r^3\beta(r))$ relatively open pairwise disjoint
1296: cells, each crossed by at most $m/r$ cylinders
1297: and containing at most $n/r^3$ points of $S$~\cite{ceg-89}
1298: (see also~\cite[p.~271]{as-95}); the first property is by definition
1299: of $(1/r)$-cuttings, and the second is enforced by subdividing cells
1300: with too many points.  The number of incidences between
1301: points and cylinders {\em crossing} their cells is thus
1302: $$
1303:   O(r^3\beta(r))\cdot I\left(\frac{n}{r^3},\frac{m}{r}\right) .
1304: $$
1305: (Note that any incidence of type 2 remains an incidence of type 2 in
1306: the subproblem it is passed to.)
1307: 
1308: It remains to bound the number of incidences between the points of
1309: $S$ and the cylinders that {\em contain} their cells. Let $\tau$ be
1310: a (relatively open) lower-dimensional cell of the cutting.
1311: If ${\rm dim}(\tau)=2$ then we can assign any point $p$ in $\tau$
1312: to one of the two neighboring full-dimensional cells, and count
1313: all but at most one of the incidences with $p$ within that cell.
1314: Hence, this increases the count by at most $n$.
1315: 
1316: If ${\rm dim}(\tau)=0$, i.e., $\tau$ is a vertex of the cutting,
1317: then any cylinder containing $\tau$ must cross or define one of the
1318: full-dimensional cells adjacent to $\tau$. Since each cell has at
1319: most $O(1)$ vertices, it follows that the total number of such
1320: incidences is $O(r^3\beta(r))\cdot (m/r) = O(mr^2\beta(r))$.
1321: 
1322: Suppose then that ${\rm dim}(\tau)=1$, i.e., $\tau$ is an edge of
1323: the cutting. An immediate implication of Lemma~\ref{3int} is that
1324: only $O(1)$ cylinders can contain $\tau$, unless $\tau$ is a line,
1325: which can then be a generator of arbitrarily many cylinders.
1326: 
1327: Since we are only counting incidences of type 2, this implies that
1328: any straight-edge 1-dimensional cell $\tau$ of the cutting generates
1329: at most one such incidence with any cylinder that fully contains
1330: $\tau$.  Non-straight edges of the cutting are contained in only
1331: $O(1)$ cylinders, as just argued, and thus the points on such edges
1332: generate a total of only $O(n)$ incidences with the cylinders. 
1333: Thus the overall number of incidences in
1334: this subcase is only $O(n+r^3\beta(r))$. Since $r\le m$,
1335: this completes the proof of the lemma.
1336: \end{proof}
1337: 
1338: \begin{lemma}\label{lem:distinct-cylinders}
1339:   The number of incidences of type 2 between $n$ points and $m$
1340:   cylinders in $\RR^3$ is
1341: \begin{equation} \label{177}
1342: O\left(\left( m^{6/7}n^{5/7}+m+n \right) \beta(n)\right).
1343: %O\left( m^{6/7}n^{5/7} \beta(n)+m+n \right).
1344: \end{equation}
1345: \end{lemma}
1346: %
1347: \begin{proof}
1348: Let $\C$ be a set of $m$ cylinders, and $S$ be a set of $n$ points.
1349: We first derive an upper bound of $O(n^5+m)$
1350: on the number of incidences of type 2 between $\C$ and $S$. We represent
1351: the cylinders as points in a dual 5-space, so that each cylinder
1352: $C$ is mapped to a point $C^*$, whose coordinates are the five
1353: degrees of freedom of $C$ (four specifying its axis and the fifth
1354: specifying its radius). A point $q\in \reals^3$ is mapped to a
1355: surface $q^*$ in $\reals^5$, which is the locus of all points dual
1356: to cylinders that are incident to $q$. With an appropriate choice of
1357: parameters, each surface $q^*$ is semi-algebraic of constant description
1358: complexity. By definition, this duality preserves incidences.
1359: 
1360: After dualization, we have an incidence problem involving $m$ points
1361: and $n$ surfaces in $\reals^5$.  We construct the arrangement
1362: $\A$ of the $n$ dual surfaces, and bound the number of their
1363: incidences with the $m$ dual points as follows.  The arrangement
1364: $\A$ consists of $O(n^5)$ relatively open cells of dimensions
1365: $0,1,\ldots,5$. Let $\tau$ be a cell of $\A$. We may assume that
1366: ${\rm dim}(\tau)\le 4$, because no point in a full-dimensional cell
1367: can be incident to any surface.
1368: 
1369: If $\tau$ is a vertex, consider any surface $\varphi$ that passes
1370: through $\tau$. Then $\tau$ is a vertex of the arrangement
1371: restricted to $\varphi$, which is a $4$-dimensional arrangement with
1372: $O(n^4)$ vertices. This implies that the number of incidences at
1373: vertices of $\A$ is at most $n\cdot O(n^4) = O(n^5)$.
1374: 
1375: Let then $\tau$ be a cell of $\A$ of dimension $\ge 1$, and let $u$
1376: denote the number of surfaces that contain $\tau$. If $u\le 8$ then each
1377: point in $\tau$ (dual to a cylinder) has at most $O(1)$ incidences
1378: of this kind, for a total of $O(m)$.
1379: 
1380: Otherwise, $u\ge 9$. Since ${\rm dim}(\tau)\ge 1$, it contains
1381: infinitely many points dual to cylinders (not necessarily in $\C$).
1382: By Lemma~\ref{3int}, back in the primal 3-space, if three cylinders
1383: contain the same nine points, then the axes of at least two of
1384: them are parallel. Hence all $u$ points lie on one line or on two
1385: parallel lines, which are common generators of these pair of
1386: cylinders. In this case, all cylinders whose dual points lie in $\tau$
1387: contain these generator(s). But then, by definition, the incidences
1388: between these points and the cylinders of $\C$ whose dual points lie
1389: on $\tau$ are of type 1, and are therefore not counted at all by
1390: the current analysis. Since $\tau$ is a face of $\A$, no other point
1391: lies on any of these cylinders, so we may ignore them completely.
1392: 
1393: Hence, the overall number of incidences under consideration is $O(n^5+m)$.
1394: 
1395: If $m>n^5$, this bound is $O(m)$. If $m<n^{1/3}$,
1396: we apply Lemma \ref{lem:cut} with $r=m$, which then yields that
1397: each recursive subproblem has at most one cylinder, so each point in a
1398: subproblem generates at most one incidence, for a total of $O(n)$
1399: incidences. Hence, in this case (\ref{eq:inci}) implies that the
1400: number of incidences between $\C$ and $S$ is
1401: $O(n+m^3\beta(m)) = O(n\beta(n))$.
1402: 
1403: Otherwise we have $n^{1/3} \leq m \leq n^5$, so we can apply
1404: Lemma~\ref{lem:cut} with parameter $r=(n^5/m)^{1/14}$; observe that
1405: $1 \leq r \leq \min\{m, n^{1/3}\}$ in this case. Using the above bound
1406: for each of the subproblems in the recurrence, we obtain
1407: $I(n/r^3,m/r)=O((n/r^3)^5+m/r)$, and thus the total number of
1408: incidences of type 2 in this case is
1409: $$
1410:   O(n+mr^2\beta(r)) + O(r^3\beta(r))\cdot O\left(
1411:     \left(\frac{n}{r^3}\right)^5 + \frac{m}{r}\right) = O\left(
1412:     \frac{n^5}{r^{12}} + mr^2 \right)\beta(r).
1413: $$
1414: The choice $r=(n^5/m)^{1/14}$ yields the bound (\ref{177}).
1415: Combining this with the other cases, the bound in the lemma follows.
1416: \end{proof}
1417: 
1418: We are now in position to complete the proof of
1419: Theorem~\ref{thm:unit3}.
1420: 
1421: \begin{proofof}{Theorem~\ref{thm:unit3}}
1422: We now return to our original setup, where the cylinders in $\C$ may
1423: have multiplicities. We fix some parameter $\mu$ and consider,
1424: as above, all cylinders in $C_\mu$, and recall our choice of $z=14/45$.
1425: The case $\mu\ge n^z$ is taken care of by Lemma~\ref{107}, accounting
1426: for at most  $O(n^{107/45}\polylog(n))$ incidences.
1427: In fact, Lemma~\ref{107} takes
1428: care of all cylinders that contain at least $n^z$ points on
1429: their axes. Assume then that $\mu < n^z$, and consider only those cylinders
1430: in $\C_\mu$ containing fewer than $n^z$ points on their axes.
1431: By (\ref{cmu}), we have $c_{\mu,\le n^z} = O(n^2/\mu^2)$.
1432: Consequently, the number of incidences with the remaining cylinders
1433: in $C_\mu$, counted with multiplicity, but excluding multiple points
1434: on the same generator line, is
1435: %
1436: $$
1437:   O\left(\mu\beta(n)\cdot \left( \left(\frac{n^2}{\mu^2}\right)^{6/7}
1438:    \cdot   n^{5/7} + \frac{n^2}{\mu^2} + n \right) \right)
1439:   = O\left(\left(\frac{n^{17/7}}{\mu^{5/7}}
1440:     + \frac{n^2}{\mu} + n\mu \right) \beta(n)\right) .
1441: $$
1442: Summing over all $\mu\leq n^z$ (powers of 2), and adding the
1443: bound $O(n^{107/45}\polylog(n))=O(n^{2.378})$ from Lemma~\ref{107} on
1444: the other kinds of incidences, we get the desired overall bound of
1445: $O(n^{17/7}\beta(n))=O(n^{2.4286})$.
1446: \end{proofof}
1447: 
1448: \noindent
1449: {\bf Remark.}
1450: In a nutshell, the ``bottleneck'' in the analysis is the case where
1451: $\mu$ is small (say, a constant) and we count incidences of type 2.
1452: The rest of the analysis, involved as it is, just shows that all the
1453: other cases contribute fewer (in fact, much fewer) incidences. One
1454: could probably simplify some parts of the analysis, at the cost of
1455: weakening the other bounds, but we leave these parts as they are, in
1456: the hope that the bottleneck case could be improved, in which case
1457: these bounds might become the dominant ones.
1458: 
1459: 
1460: \section{Minimum-area triangles in 3-space}
1461: \label{sec:minimum3}
1462: 
1463: Place $n$ equally spaced
1464: points on the three parallel edges of a right prism whose base is an
1465: equilateral triangle, such that inter-point distances are small along
1466: each edge. This construction yields $\frac{2}{3}n^2-O(n)$ minimum-area triangles,
1467: a slight improvement over the lower bound construction in the plane.
1468: Here is yet another construction with the same constant $2/3$ in the leading
1469: term: Form a rhombus in the $xy$-plane from two equilateral triangles
1470: with a common side, extend it to a prism in 3-space, and place $n/3$
1471: equally spaced points on each of the lines passing through the vertices
1472: of the shorter diagonal of the rhombus, and $n/6$ equally spaced
1473: points on each of the two other lines, where again the
1474: inter-point distances along these lines are all equal and small.
1475: The number of minimum-area triangles is
1476: $$ 2\left(\frac{1}{3 \cdot 3} + \frac{4}{3 \cdot 6}\right)n^2 -O(n)=
1477: \frac{2}{3}n^2-O(n). $$
1478: The following theorem shows that this bound is optimal up to a
1479: constant factor. No quadratic upper bound has previously been known for
1480: minimum-area triangles in $\RR^3$.
1481: 
1482: \begin{theorem} \label{thm:min3}
1483:   The number of triangles of minimum (nonzero) area spanned by $n$ points in
1484:   $\RR^3$ is at most $n^2+O(n)$.
1485: %This estimate is optimal up to a constant factor.
1486: \end{theorem}
1487: %
1488: \begin{proof}
1489: Consider a set $S$ of $n$ points in $\RR^3$, and let $T$ be the set of
1490: triangles of minimum (nonzero) area spanned by $S$.
1491: Without loss of generality, assume the minimum area to be $1$.
1492: Similarly to the planar case, we assign each
1493: triangle in $T$ to one of its longest sides, and argue that at most a
1494: constant number of triangles are assigned to each segment spanned by
1495: $S$. This immediately implies an upper bound of $O(n^2)$ on the
1496: cardinality of $T$. To improve the main coefficient in this bound, we
1497: distinguish between {\em fat} and {\em thin} triangles.
1498: A triangle is called fat (resp., thin) if the length of the height
1499: corresponding to its longest side is at least (resp., less than)
1500: half of the length of the longest side.
1501: We show that the number $N_1$ of thin triangles of minimum area is
1502: at most $2{n \choose 2}=n^2-n$, and that
1503: the number $N_2$ of fat triangles of minimum area is only $O(n)$.
1504: 
1505: Consider a segment $ab$, with $a,b\in S$, and let $h=|ab|$.  Every
1506: point $c\in S\setminus \{a,b\}$ for which the triangle
1507: $\Delta{abc}$ has minimum (unit) area must lie on a bounded cylinder $C$
1508: with axis $ab$, radius $r=2/h$,
1509: and bases that lie on the planes $\pi_a$ and $\pi_b$, incident to
1510: $a$ and $b$, respectively, and orthogonal to $ab$.  In fact, if
1511: $\Delta abc$ is assigned to $ab$ (that is, $ab$ is the longest side), 
1512: then $c$ must lie on a smaller
1513: portion $C'$ of $C$, bounded by bases that intersect $ab$ at points at
1514: distance $h-\sqrt{h^2-r^2}$ from $a$ and $b$, respectively.
1515: Assume for
1516: convenience that $ab$ is vertical, $a$ is the origin and
1517: $b=(0,0,h)$. Since $ab$ is the longest side of $\Delta abc$, the
1518: side of the isosceles triangle with base $ab$ and height $r$ must
1519: be no larger than $h$, i.e., $\frac14 h^2+r^2 \le h^2$, or
1520: $r^2\le \frac34 h^2$.
1521: Notice that the triangle formed by any two points
1522: of $S$ lying on $C'$ with either $a$ or $b$ is non-degenerate.
1523: 
1524: We first derive a simple formula that relates the area of any
1525: (slanted) triangle to the area of its $xy$-projection. Consider a
1526: triangle $\Delta$ that is spanned by two vectors $u,v$, and let
1527: $\Delta_0$, $u_0$, and $v_0$ denote the $xy$-projections of $\Delta$,
1528: $u$, and $v$, respectively. Write (where $\kk$ denotes, as usual, the
1529: vector $(0,0,1)$)
1530: $$
1531: u=u_0 + x\kk \quad\text{and}\quad v=v_0 + y\kk ,
1532: $$
1533: and put $A= {\rm area}(\Delta)$, $A_0= {\rm area}(\Delta_0)$.
1534: Then
1535: $$
1536: A^2 = \frac14 \|u\times v\|^2 =
1537: \frac14\| (u_0 + x\kk)\times (v_0 + y\kk) \| =
1538: \frac14\left( \| u_0\times v_0 \|^2 + \|yu_0 - xv_0\|^2 \right) ,
1539: $$
1540: or
1541: \begin{equation} \label{eq:pa}
1542: A^2 = A_0^2 + \frac14 \|yu_0 - xv_0\|^2 .
1543: \end{equation}
1544: 
1545: \vspace{-1mm}
1546: \paragraph{An initial weaker bound.}
1547: We claim that at most 10 triangles are assigned to $ab$. Assume, to
1548: the contrary, that this number is at least 11.  Divide $C$ into two
1549: equal slices by a horizontal plane orthogonal to $ab$ through its
1550: midpoint. Since more than 10 points of $S$ lie on $C$, at least
1551: 6 of them must lie on the same slice $C_0$, say the bottom slice.
1552: It follows that two points, $c$ and $d$,
1553: lie in some sector $\Upsilon$ of $C_0$ making a dihedral angle
1554: $\alpha$ at $ab$ of at most $360^\circ/6=60^\circ$.
1555: An illustration is provided in Figure~\ref{fig:min3}.
1556: 
1557: \begin{figure}[htbp]
1558:   \begin{center}
1559: \epsfig{file=min3.eps,width=3in,clip=}
1560:  \caption{\small Charging scheme for minimum-area triangles in
1561:    3-space; (a) the cylinder $C$; (b) the projection on $\pi_a$;
1562:    $c'$ and $d'$ are the respective projections of $c$ and $d$.}
1563: \end{center}
1564: \label{fig:min3}
1565: \end{figure}
1566: 
1567: We may assume, without loss of generality, that
1568: $$
1569: c=(r,0,x)=c_0+x\kk \quad\text{and}\quad
1570: d=(r\cos\alpha,r\sin\alpha,y)=d_0+y\kk ,
1571: $$
1572: where $0\le\alpha\le 60^\circ$ and $0\le x,y\le h/2$.
1573: Write $A={\rm area}(\Delta{acd})$. Using (\ref{eq:pa}), we have
1574: %The squared area of $\Delta acd$ is
1575: %$$
1576: %\frac14 \left|\overrightarrow{ac}\times\overrightarrow{ad}\right|^2 =
1577: %\frac14 \left| c_0\times d_0 + (yc_0-xd_0)\times\kk \right|^2 =
1578: %$$
1579: $$
1580: A^2=
1581: \frac14 \left\| c_0\times d_0\right\|^2 + \frac14\left\|yc_0-xd_0\right\|^2 =
1582: \frac{r^4\sin^2\alpha}{4} + \frac{r^2}{4}\left( x^2+y^2-2xy\cos\alpha\right).
1583: $$
1584: The expression $x^2+y^2-2xy\cos\alpha$ is the squared length of the
1585: third side of the triangle with sides $x$, $y$, with the angle
1586: $\alpha \leq 60^\circ$ between them. Since $x,y \leq h/2$, we clearly
1587: have $x^2+y^2-2xy\cos\alpha \leq h^2/4$.
1588: Thus, recalling that $r^2\le \frac34h^2$ and that $h^2r^2 = 4$, we have
1589: $$
1590: A^2 \le \frac{r^4\sin^2\alpha}{4} + \frac{r^2}{4} \cdot \frac{h^2}{4}=
1591: \frac{r^2}{4} \left( r^2\sin^2\alpha+ \frac{h^2}{4} \right)
1592: \leq \frac{r^2 h^2}{4} \left( \frac{9}{16} + \frac{1}{4} \right)=
1593: \frac{13}{16} <1,
1594: $$
1595: which contradicts the minimality of the area of $\Delta abc$.
1596: Hence, at most 10 triangles are assigned to each segment
1597: spanned by $S$.
1598: This already implies that there are at most $5(n^2-n)$
1599: minimum-area triangles.
1600: 
1601: \vspace{-1mm}
1602: \paragraph{A better bound.}
1603: We now improve the constant of proportionality, using a more careful
1604: analysis, which distinguishes between the cases in which the
1605: minimum-area triangles charged to the segment $ab$ are thin or fat.
1606: 
1607: \smallskip
1608: \noindent{\bf (a)} $r< \frac12h$ (thin triangles).
1609: We claim that in this case at most two triangles can be assigned
1610: to $ab$. Indeed, suppose to the contrary that at least three triangles
1611: are assigned to $ab$, so their third vertices, $c,d,e \in S$ lie on
1612: $C'\subset C$. 
1613: Write the $z$-coordinates of $c,d,e$ as $z_1h$, $z_2h$, $z_3h$,
1614: respectively, and assume, without loss of generality, that
1615: $0< z_1\le z_2\le z_3< 1$, and $z_2\le 1/2$.  Consider the
1616: triangle $\Delta{acd}$, and let $A$ denote its area.
1617: As before, write, without loss of generality,
1618: $$
1619: c=(r,0,z_1h) \quad\text{and}\quad
1620: d=(r\cos\alpha,r\sin\alpha,z_2h) ,
1621: $$
1622: for some $0\le\alpha\le 180^\circ$.  Using (\ref{eq:pa}), we get
1623: $$
1624: A^2=\frac14 r^4\sin^2\alpha + \frac14 r^2h^2 (z_1^2+z_2^2-2z_1z_2\cos\alpha).
1625: $$
1626: Thus, recalling that $r< \frac12h$ and that $h^2r^2 = 4$, we get
1627: \begin{equation} \label{eq:A2}
1628: A^2 <
1629: \frac14 r^2h^2 \left( \frac14\sin^2\alpha +
1630: z_1^2+z_2^2-2z_1z_2\cos\alpha \right) =
1631: \frac14\sin^2\alpha + z_1^2+z_2^2-2z_1z_2\cos\alpha .
1632: \end{equation}
1633: Let us fix $z_1,z_2$ and vary only $\alpha$. Write
1634: $$
1635: f(\alpha) =
1636: \frac14\sin^2\alpha + z_1^2+z_2^2 - 2z_1z_2\cos\alpha ,
1637: \quad\text{and}\quad
1638: f'(\alpha) =
1639: \frac12\sin\alpha\cos\alpha + 2z_1z_2\sin\alpha .
1640: $$
1641: $f$ attains its maximum at the zero of its derivative, namely at
1642: $\alpha_0$ that satisfies
1643: $$
1644: \cos\alpha_0 = - 4z_1z_2 .
1645: $$
1646: (Note that since $z_1\le z_2\le\frac12$, we always have $4z_1z_2\le 1$.
1647: Also, at the other zero $\alpha=0$, $f$ attains its minimum
1648: $(z_1-z_2)^2$.)
1649: 
1650: Substituting $\alpha_0$ into (\ref{eq:A2}),
1651: and using $z_1\le z_2\le\frac12$, we get
1652: $$
1653: A^2 < \frac{1-16z_1^2z_2^2}{4} + z_1^2 + z_2^2 + 8z_1^2z_2^2 =
1654: \frac14 + z_1^2 + z_2^2 + 4z_1^2z_2^2 =
1655: \left(\frac12 + 2z_1^2\right) \left(\frac12 + 2z_2^2\right) \le 1 ,
1656: $$
1657: which contradicts the minimality of the area of $\Delta abc$
1658: (recall that $\Delta acd$ is non-degenerate).
1659: 
1660: We have thus shown that at most two thin triangles of minimum area can be
1661: assigned to any segment $ab$, so $N_1 \leq 2{n \choose 2}=n^2-n$.
1662: 
1663: 
1664: \smallskip
1665: \noindent{\bf (b)} $r\ge \frac12h$ (fat triangles).
1666: Recall that we always have $r\le \frac{\sqrt{3}}{2}h$. Multiplying
1667: these two inequalities by $h/2$, we get
1668: $$
1669: \frac{h^2}{4} \le 1 \le \frac{h^2\sqrt{3}}{4} ,
1670: \quad\text{or}\quad
1671: \frac{2}{3^{1/4}} \le h \le 2 .
1672: $$
1673: % Note that $r > 3/4$.
1674: Let $E$ denote the set of all segments $ab$ such that the minimum-area
1675: triangles charged to $ab$ are fat. Note that the length of each edge
1676: in $E$ is in the interval $[2/3^{1/4}, 2]$.
1677: 
1678: We next claim that, for any pair of points $p,q\in S$ with
1679: $|pq| < 1$, neither $p$ nor $q$ can be an endpoint of an edge in $E$.
1680: Indeed, suppose to the contrary that $p,q$ is such a pair and that $pa$
1681: is an edge of $E$, for some $a\in S$; by construction, $a\ne q$.
1682: Let $\Delta pab$ be a fat minimum-area triangle charged to $pa$.
1683: If $q$ is collinear with $pa$, then $\Delta pqb$ is a nondegenerate
1684: triangle of area strictly smaller than that of $\Delta pab$ (recall that
1685: $|pq|< 1 < |pa|$), a contradiction.
1686: If $q$ is not collinear with $pa$, $\Delta paq$ is a nondegenerate
1687: triangle of area $ \leq \frac{|pa| \cdot |pq|}{2} < \frac{2 \cdot 1}{2} =1$,
1688: again a contradiction.
1689: 
1690: Let $S' \subseteq S$ be the set obtained by repeatedly removing
1691: the points of $S$ whose nearest neighbor in $S$ is at distance
1692: smaller than $1$.  Clearly, the minimum inter-point distance in $S'$ is at
1693: least $1$, and the endpoints of each edge in $E$ lie in $S'$. This implies,
1694: via an easy packing argument, that the number of edges of $E$ incident
1695: to any fixed point in $S'$ (all of length at most $2$) is only $O(1)$.
1696: Hence $|E|=O(n)$. Since each edge in $E$ determines at most $10$
1697: minimum-area triangles, as shown in the first part of our proof,
1698: we conclude that $N_2=O(n)$, as claimed.
1699: 
1700: Hence there are at most $2{n\choose 2} + O(n) = n^2+O(n)$
1701: minimum-area triangles in total.
1702: \end{proof}
1703: 
1704: 
1705: \section{Maximum-area triangles in 3-space}
1706: \label{sec:maximum3}
1707: 
1708: \'Abrego and Fern\'andez-Merchant~\cite{af-02} showed that one can
1709: place $n$ points on the unit sphere in $\RR^3$ so that they determine
1710: $\Omega(n^{4/3})$ pairwise distances of $\sqrt{2}$
1711: (see also~\cite[p.~191]{pa-95} and~\cite[p.~261]{bmp-05}).
1712: % They also showed that one can obtain $\Omega(n^{4/3})$ congruent
1713: % copies of any given triangle $T$. By selecting $T$ as a right
1714: % angled isosceles triangle, we can also ensure that $T$ has maximum
1715: % area among all triangles, and thus obtain
1716: This implies the following result:
1717: %
1718: \begin{theorem}\label{thm:maximum3}
1719: For any integer $n$, there exists an $n$-element point set in $\RR^3$
1720: that spans $\Omega(n^{4/3})$ triangles of maximum area, all incident to
1721: a common point.
1722: \end{theorem}
1723: %
1724: \begin{proof}
1725: Denote the origin by $o$, and consider a unit sphere centered at
1726: $o$. The construction in~\cite{af-02} consists of a set
1727: $S=\{o\}\cup S_1\cup S_2$ of $n$
1728: points, where $S_1\cup S_2$ lies on the unit sphere,
1729: $|S_1|=\lfloor (n-1)/2\rfloor$, $|S_2|=\lceil (n-1)/2 \rceil$,
1730: and there are $\Omega(n^{4/3})$ pairs of orthogonal segments of the form
1731: $(os_i,os_j)$ with  $s_i \in S_1$ and $s_j \in S_2$.
1732: 
1733: Moreover, this construction can be realized in such a way that
1734: $S_1$ lies in a small neighborhood of $(1,0,0)$, and $S_2$ lies in a
1735: small neighborhood of $(0,1,0)$, say. The area of every right-angled
1736: isosceles triangle $\Delta os_is_j$ with $s_i\in S_1$ and $s_j\in
1737: S_2$ is $1/2$. All other triangles have
1738: smaller area: this is clear if at least two vertices of a triangle
1739: are from $S_1$ or from $S_2$; otherwise the area is given by
1740: $\frac{1}{2}\sin \alpha$, where $\alpha$ is the angle of the two
1741: sides incident to the origin, so the area is less than $1/2$ if
1742: these sides are not orthogonal.
1743: \end{proof}
1744: 
1745: \smallskip
1746: We next show that the construction in Theorem~\ref{thm:maximum3} is
1747: almost tight, in the sense that at most $O(n^{4/3+\eps})$ maximum-area
1748: triangles can be incident to any point of an $n$-element point
1749: set in $\RR^3$, for any $\eps>0$.
1750: 
1751: \begin{theorem}  \label{thm:max-3d}
1752:   The number of triangles of maximum area spanned by a set $S$ of $n$
1753:   points in $\RR^3$ and incident to a fixed point $a\in S$ is
1754:   $O(n^{4/3+\eps})$, for any $\eps>0$.
1755: \end{theorem}
1756: 
1757: Assume, without loss of generality, that the maximum area is $1$.
1758: Similarly to the proof of Theorem~\ref{thm:unit3}, we map
1759: maximum-area triangles to point-cylinder incidences. Specifically,
1760: if $\Delta{abc}$ is a maximum-area triangle spanned by a point set $S$,
1761: then every point of $S$ lies on, or in the interior of, the cylinder
1762: with axis $ab$ and radius $2/|ab|$ ($c$ itself lies on the cylinder).
1763: The following two lemmas give upper bounds on the number of
1764: point-cylinder incidences in this setting. First we prove a
1765: weaker bound (Lemma~\ref{lem:kst-cylinder}) which, combined with the
1766: partition technique, gives an almost tight bound
1767: (Lemma~\ref{lem:st-cylinder}).
1768: %\micha{Why is the bound tight?}\csaba{construction is given after the lemma.}
1769: %
1770: Our proof is somewhat reminiscent of an argument of Edelsbrunner and
1771: Sharir~\cite{es-91}, where it is shown that the number of point-sphere
1772: incidences between $n$ points and $m$ spheres in $\RR^3$ is
1773: $O(n^{2/3}m^{2/3}+n+m)$, provided that no point lies in the exterior
1774: of any sphere.
1775: 
1776: \begin{lemma}\label{lem:kst-cylinder}
1777:   Let $S$ be a set of $n$ points, and $\C$ a set of $m$ cylinders in
1778:   $\RR^3$, such that the axis of each cylinder passes through the
1779:   origin, and no point lies in the exterior of any cylinder. Then
1780:   the number of point-cylinder incidences is
1781:   $O(nm^{\frac{1+\eps}{2}}+m)$, for any $\eps>0$.
1782: \end{lemma}
1783: %
1784: \begin{proof}
1785:   Assume, without loss of generality, that the horizontal plane $h$
1786:   incident to the origin does not contain any point of $S$, and that
1787:   the points above $h$ participate in at least half of the
1788:   point-cylinder incidences. For simplicity, continue to denote by
1789:   $S$ the subset of the at most $n$ points lying above $h$.
1790:   Consider the 3-dimensional dual arrangement $(S^*,\C^*)$, where
1791:   the dual of a point $p\in \RR^3\setminus \{o\}$ is the cylinder
1792:   $p^*$ with axis $op$ and radius $2/|op|$; and the dual of a cylinder
1793:   $\gamma$ whose axis passes through the origin is a point $\gamma^*$
1794:   above $h$ that lies on the axis of $\gamma$ at distance
1795:   $2/{\rm radius}(\gamma)$ from the origin. Note that incidences
1796:   between points and cylinders are preserved, and that no point of
1797:   $\C^*$ lies in the exterior of any cylinder of $S^*$. It therefore
1798:   suffices to prove that the number of incidences between $S^*$ and 
1799:   $\C^*$ is
1800:   $I(\C^*,S^*)=O(nm^{\frac{1+\eps}{2}}+m)$.
1801: 
1802:   Consider the intersection $B$ of the interiors of all cylinders in
1803:   $S^*$. Since the interior of each cylinder is convex, $B$ is a convex
1804:   body homeomorphic to a ball, whose boundary is composed of patches of
1805:   cylinders. Faces, edges, and vertices of $B$ can be defined as
1806:   connected components of the intersections of one, two, and three
1807:   cylinders, respectively.
1808: %
1809: Each of the points of $\C^*$ that lie on faces of $\bd B$ contributes
1810: one incidence. Since all the cylinder axes pass through the origin, no
1811: edge of $\bd B$ can be straight, so it cannot be contained in any
1812: cylinder of $S^*$ other than the two defining it (recall
1813: Lemma~\ref{3int}). Hence the points of $\C^*$ that lie on faces
1814: or edges of $\bd B$ contribute at most $2m$ incidences.
1815: 
1816: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%5
1817: We are left with the task of bounding the
1818: number of point-cylinder incidences involving points at vertices 
1819: of $B$. Note that there may exist cylinders incident to a vertex $p$ 
1820: of $B$ and not containing any other points of $\bd B$ in the 
1821: vicinity of $p$. To account for such cylinders too, perturb
1822: the radii of each cylinder in $S^*$, so that each radius $r$ is 
1823: decreased to the radius $(1-\delta r)r$,
1824: for a sufficiently small $\delta>0$ (that is, the radii of larger 
1825: cylinders decrease by a larger factor).
1826: As a result, every cylinder incident to a vertex $p\in \bd B$ is
1827: replaced by a cylinder that defines a face in a sufficiently
1828: small neighborhood of $p$ (even though it is not incident to
1829: $p$ after this perturbation). The number of point-cylinder incidences 
1830: between $\C^*$ and the vertices of $\bd B$ is proportional to
1831: the number of vertices of the resulting $\bd B'$ after the perturbation. 
1832: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%5
1833:   By a result of Halperin and Sharir~\cite{hs-95}, the
1834:   complexity of a single cell in the arrangement of $n$ constant
1835:   degree algebraic surfaces in $\RR^3$ is $O(n^{2+\eps})$,
1836:   for any $\eps>0$.
1837:   % (see~\cite{b-03} for extensions and generalizations
1838:   % to higher dimensions).
1839:   Hence, we obtain an upper bound of
1840:   $I(S,\C)=O(m+n^{2+\eps})$, for any $\eps>0$.
1841: 
1842:   Partition $S$ into $\lceil n/\sqrt{m}\rceil$ subsets, each
1843:   containing at most $\sqrt{m}$ points. The preceding argument
1844:   implies that each subset $S'\subset S$ has at most
1845:   $I(S',\C)=O(m+(\sqrt{m})^{2+\eps}) = O(m^{1+\eps/2})$ incidences with
1846:   the cylinders of $\C$. Therefore, altogether there are at most
1847:   $\lceil n/\sqrt{m}\rceil \cdot O(m^{1+\eps/2})
1848:   =O(nm^{\frac{1+\eps}{2}}+m)$ incidences.
1849: \end{proof}
1850: 
1851: \begin{lemma}\label{lem:st-cylinder}
1852:   Let $S$ and $\C$ be as in the preceding lemma.
1853:   Then the number of point-cylinder
1854:   incidences is $O((n^{2/3}m^{2/3}+n+m)^{1+\eps})$, for any $\eps>0$.
1855: \end{lemma}
1856: %
1857: \begin{proof}
1858: If $m>n^2$, then Lemma~\ref{lem:kst-cylinder} gives an upper bound
1859: of $O(nm^{\frac{1+\eps}{2}}+m) = O(m^{1+\eps})$.
1860: We may therefore assume henceforth that $m\leq n^2$.
1861: 
1862:   For an integer $r\in \NN$, to be specified later, choose a random
1863:   sample $R\subset \C$ of $r$ cylinders, and let $B$ denote the
1864:   intersection of the interiors of the cylinders in $R$.
1865:   By~\cite{hs-95}, the combinatorial complexity of $B$ is
1866:   $O(r^{2+\eps})$, for any $\eps>0$. Hence, the convex body $B$ can
1867:   be partitioned into $O(r^{2+\eps})$ cells, each bounded by a
1868:   constant number of constant-degree algebraic surfaces.
1869:   (This can be done, e.g., by first partitioning $\bd B$ into
1870:   pseudo-trapezoidal cells, and then by
1871:   taking the convex hull of each cell on $\bd B$ with the origin.)
1872:   By the $\eps$-net theory (see, e.g.,~\cite[Chap.~10.3]{m-02}), with
1873:   constant probability, the interior of each cell intersects at most
1874:   $O(m \frac{\log r}{r}) = O(m/r^{1-\eps})$ cylinders of $\C$.
1875:   We may assume then that our sample $R$ has this property.
1876:   Similarly to the proof of
1877:   Lemma~\ref{lem:cut}, assign each point to a unique cell. Assign
1878:   every point in the interior of a cell $\sigma_i$ to $\sigma_i$;
1879:   assign every point on the boundary of several cells to the cell with
1880:   minimum index. Let $n_i$ denote the number of points assigned to
1881:   cell $\sigma_i$.
1882: 
1883:   Applying Lemma~\ref{lem:kst-cylinder} in each cell $\sigma_i$, we
1884:   get the upper bound
1885:   ${\displaystyle O\left(n_i
1886:     \left(\frac{m}{r^{1-\eps}}\right)^{\frac{1+\eps}{2}} +
1887:     \left(\frac{m}{r^{1-\eps}}\right)\right) }$
1888: on the number of incidences between points assigned
1889:   to $\sigma_i$ and cylinders intersecting the interior of $\sigma_i$.
1890:   Summing over all $O(r^{2+\eps})$ cells, we have
1891: %
1892:   $$
1893:   \sum_{i} O\left(n_i
1894:     \left(\frac{m}{r^{1-\eps}}\right)^{\frac{1+\eps}{2}} +
1895:     \left(\frac{m}{r^{1-\eps}}\right)\right) = O\left(n\left(
1896:       \frac{m}{r^{1-\eps}}\right)^{\frac{1+\eps}{2}} +
1897:     mr^{1+2\eps}\right) =
1898:   O\left(\frac{nm^{\frac{1+\eps}{2}}}{r^{\frac{1-\eps}{2}}} +
1899:     mr^{1+2\eps}\right)
1900:   $$
1901:   incidences of this kind. By choosing
1902:   $r=\min\left\{\lfloor n^{2/3}/m^{1/3}\rfloor,\,m\right\}$, 
1903:   this is at most
1904:   $O(n^{2/3+\eps'}m^{2/3+\eps'}+n^{1+\eps'})$, for another, still
1905: arbitrarily small, $\eps'>0$.  Finally, the number of incidences
1906:   between points assigned to one cell and cylinders that do not
1907:   intersect the interior of that cell can be bounded similarly to the
1908:   proof of Lemma~\ref{lem:cut}: This number is proportional to the
1909:   number of cells plus the number of points, which is
1910:   $O(n+r^{2+\eps}) = O(n^{1+\eps})$, as is easily checked.
1911: (In this final argument, we use the fact all axes pass through the
1912: origin, so no 1-dimensional edge of $\bd B$ can be contained in more
1913: than two cylinders; see also the proof of Lemma~\ref{lem:kst-cylinder}.)
1914: \end{proof}
1915: 
1916: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%5
1917: The upper bound of Lemma~\ref{lem:st-cylinder} is almost tight: For any 
1918: $n$ and $m$, there are $n$ points and $m$ cylinders with axes through 
1919: the origin and containing no points in their exterior, which determine 
1920: $\Omega(n^{2/3}m^{2/3}+n+m)$ point-cylinder incidences. To construct
1921: such a configuration, take $n$ points and $m$ lines on the plane 
1922: $\pi:z=1$ in $\RR^3$ with $\Omega(n^{2/3}m^{2/3}+n+m)$ 
1923: point-line incidences~\cite{st-83}. Project these points and lines centrally
1924: from the origin onto the unit sphere, to obtain a system of $n$ points 
1925: and $m$ great circles with the same number of incidences.
1926: Each great circle of the unit sphere lies in a unique cylinder of unit 
1927: radius whose axis passes through the origin, and every such cylinder 
1928: contains all the other points of the unit sphere in its interior. 
1929: This gives $n$ points on the unit sphere and $m$ cylinders of 
1930: unit radius whose axes pass through the origin
1931: (so that no point lies in the exterior of any cylinder),
1932: with $\Omega(n^{2/3}m^{2/3}+n+m)$ point-cylinder incidences.
1933: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%5
1934: 
1935: \begin{proofof}{Theorem~\ref{thm:max-3d}}
1936: Let $A$ denote the maximum triangle area determined by a set $S$ of
1937: $n$ points in $\RR^3$. For every point $a\in S$, consider the system
1938: of $n-1$ points in $S\setminus \{a\}$ and $n-1$ cylinders, each defined
1939: by a point $b\in S\setminus \{a\}$, and has axis $ab$ and radius
1940: $2A/|ab|$.
1941: Every point-cylinder incidence corresponds to a triangle of area $A$
1942: spanned by $S$ and incident to $a$. Since $A$ is the maximum area, no
1943: point of $S$ may lie in the exterior of any cylinder.  By
1944: Lemma~\ref{lem:st-cylinder}, the number of such triangles is
1945: $O(n^{4/3+\eps})$, for any $\eps>0$.
1946: \end{proofof}
1947: 
1948: Theorems \ref{thm:maximum3} and \ref{thm:max-3d} imply the
1949: following bounds on the number of maximum-area triangles in $\RR^3$:
1950: 
1951: \begin{theorem}\label{thm:allmax-3d}
1952: The number of triangles of maximum area spanned by $n$ points in
1953: $\RR^3$ is $O(n^{7/3+\eps})$, for any $\eps>0$.
1954: For all $n \geq 3$, there exist $n$-element point sets in $\RR^3$
1955: that span $\Omega(n^{4/3})$ triangles of maximum area.
1956: \end{theorem}
1957: 
1958: 
1959: \section{Distinct triangle areas in 3-space}\label{sec:dist3d}
1960: 
1961: Following earlier work by Erd\H{o}s and Purdy~\cite{ep-76}, Burton and
1962: Purdy~\cite{bp-79}, and Dumitrescu and T\'oth~\cite{dt-07a},
1963: Pinchasi~\cite{p-07} has recently proved that $n$ noncollinear points in
1964: the plane always determine at least $\left\lfloor
1965:   \frac{n-1}{2}\right\rfloor$ distinct triangle areas, which is
1966: attained by $n$ equally spaced points distributed evenly on two
1967: parallel lines. No linear lower bound is known in 3-space, and
1968: the best we can show is the following:
1969: 
1970: \begin{theorem}\label{thm:dist3d}
1971:   Any set $S$ of $n$ points in $\RR^3$, not all on a line, determines at
1972:   least $\Omega(n^{2/3}/\beta(n))$ triangles of distinct areas, for
1973:   some extremely slowly growing function $\beta(n)$.
1974:   Moreover, all these triangles share a common side.
1975: \end{theorem}
1976: 
1977: For the proof, we first derive a new upper bound
1978: (Lemma~\ref{lem:clarkson}) on the number of point-cylinder
1979: incidences in $\RR^3$, for the special case where the axes of the
1980: cylinders pass through the origin (but without the additional requirement
1981: that no point lies outside any cylinder).
1982: %
1983: Consider a set $\C$ of $m$ such cylinders.
1984: These cylinders have only three degrees of
1985: freedom, and we can dualize them to points in 3-space. We use a
1986: duality similar to that used in the proof of Lemma~\ref{lem:kst-cylinder}.
1987: Specifically, we fix some generic halfspace $H$ whose bounding plane
1988: passes through the origin, say, the halfspace $z>0$. We then map
1989: each cylinder with axis $\ell$ and radius $\varrho$ to the point on
1990: $\ell\cap H$ at distance $1/\varrho$ from the origin;
1991: and we map each point $p\in H$ to the cylinder whose axis is the line
1992: spanned by $op$ and whose radius is $1/|op|$.
1993: As argued above, this duality preserves point-cylinder incidences.
1994: 
1995: By (a dual version of) Lemma~\ref{3int}, any three points
1996: can be mutually incident to at most eight cylinders whose axes
1997: pass through the origin. That is, the
1998: bipartite incidence graph (whose two classes of vertices correspond to
1999: the points of $S$ and the cylinders of $\C$, respectively, and an edge
2000: represents a point-cylinder incidence) is $K_{3,9}$-free. It follows
2001: from the theorem of K\H{o}v\'ari, S\'os and Tur\'an~\cite{kst-54} (see
2002: also~\cite[p.~121]{pa-95})
2003: that the number of point-cylinder incidences is $O(nm^{2/3}+m)$.
2004: We then combine this bound with the partition technique of
2005: Clarkson~\etal~\cite{ceg-90}, to prove a sharper upper bound on
2006: the number of point-cylinder incidences of this kind. Specifically, we
2007: have:
2008: 
2009: \begin{lemma}\label{lem:clarkson}
2010:   Given $n$ points and $m$ cylinders, whose axes pass through the
2011:   origin, in 3-space, the number of point-cylinder incidences is
2012:   $O(n^{3/4}m^{3/4}\beta(n)+n+m).$
2013: \end{lemma}
2014: %
2015: \begin{proof}
2016: Let $\C$ be the set of the $m$ given cylinders, and $S$ be the set of
2017: the $n$ given points.
2018: Let $h$ be a plane containing the origin, but no point of $S$, and
2019: assume, without loss of generality, that the subset $S'$ of points
2020: lying in the positive hafspace $h^+$ contributes at least half of
2021: the incidences with $\C$.  If $m>n^3$, then the
2022: K\H{o}v\'ari-S\'os-Tur\'an Theorem yields an upper bound of
2023: $I(S',\C)=O(nm^{2/3}+m)=O(m)$. Similarly, if $m<n^{1/3}$,
2024: the duality mentioned above leads to the bound
2025: $I(S',\C)=O(mn^{2/3}+n)=O(n)$. For these two cases we have
2026: then $I(S,\C) \leq 2I(S',\C)=O(m+n)$.
2027: Assume henceforth that $n^{1/3}\leq m\leq n^3$.
2028: 
2029: We apply Lemma~\ref{lem:cut} with parameter 
2030: $r= \lfloor n^{3/8}/m^{1/8} \rfloor$, and
2031: use the K\H{o}v\'ari-S\'os-Tur\'an Theorem to bound the number of
2032: incidences between the at most $n/r^3$ points and $m/r$ cylinders
2033: in each subproblem.
2034: Note that $1 \leq r \leq m$ in the above range of $m$.
2035: The total number of incidences is thus
2036: %
2037: \begin{eqnarray*}
2038: I(S,\C) &=& O(n+mr^2\beta(r)) + O(r^3\beta(r))\cdot
2039: O \left( \frac{n}{r^3} \cdot \left(\frac{m}{r}\right)^{2/3}
2040: +\frac{m}{r} \right)\\
2041: &=&
2042: O \left( n+ \frac{m^{2/3}n}{r^{2/3}}\beta(n) + m r^2\beta(r)\right) =
2043: O\left(n+ n^{3/4}m^{3/4}\beta(n)\right).
2044: \end{eqnarray*}
2045: Putting all three cases together gives the bound in the theorem.
2046: \end{proof}
2047: 
2048: 
2049: \begin{proofof}{Theorem~\ref{thm:dist3d}}
2050:   If there are $n/100$ points in a plane but not all on a line, then
2051:   the points in this plane already determine $\Omega(n)$ triangles of
2052:   distinct areas~\cite{bp-79}. We thus assume, in the remainder of the
2053:   proof, that there are at most $n/100$ points on any plane.
2054: 
2055:   According to a result of Beck~\cite{b-83}, there is an absolute
2056:   constant $k\in \NN$ such that if no line is incident to $n/100$ points
2057:   of $S$, then $S$ spans $\Theta(n^2)$ distinct lines, each of which
2058:   is incident to at most $k$ points of $S$. Since each point of $S$ is
2059:   incident to at most $n-1$ of these lines, there is a point $a\in S$
2060:   incident to $\Theta(n)$ such lines. Select a point of $S\setminus \{a\}$
2061:   on each of these lines, to obtain a set $P$ of $\Theta(n)$ points.
2062: 
2063:   Let $t$ denote the number of distinct triangle areas determined by
2064:   $S$, and let $\alpha_1,\alpha_2,\ldots ,\alpha_t$ denote these
2065:   areas.  For each point $b\in P$ and $i=1,2,\ldots ,t$, we define a
2066:   cylinder $C(ab,\alpha_i)$ with axis (the line spanned by) $ab$ and radius
2067:   $2\alpha_i/|ab|$.
2068:   Every point $c\in S$ for which the area of the
2069:   triangle $\Delta{abc}$ is $\alpha_i$ must lie on the cylinder
2070:   $C(ab,\alpha_i)$. Let $\C$ denote the set of the $O(nt)$ cylinders
2071:   $C(ab,\alpha_i)$, for $b\in P$ and $i=1,2,\ldots,t$.
2072:   For each point $b\in P$, there are
2073:   $n-k=\Theta(n)$ points off the line through $ab$, each of which must
2074:   lie on a cylinder $C(ab,\alpha_i)$ for some $i=1,2,\ldots , t$.
2075:   Therefore, the number $I(S,\C)$ of point-cylinder incidences between
2076:   $S$ and $\C$ is $\Omega(n^2)$. On the other hand, by
2077:   Lemma~\ref{lem:clarkson}, we have
2078: %
2079: $$\Omega(n^2)\leq I(S,\C)\leq O(n^{3/4}(nt)^{3/4}\beta(n)+n+nt)
2080: = O(n^{3/2}t^{3/4}\beta(n)),$$
2081: %
2082: which gives $t=\Omega (n^{2/3}/\beta^{4/3}(n))=\Omega
2083: (n^{2/3}/\beta'(n))$, for another function $\beta'(n)$ of the same
2084: slowly growing type, as required.
2085: \end{proofof}
2086: 
2087: \section{Conclusion} \label{sec:conclusion}
2088: 
2089: We have presented many results on the number of triangles of specific areas
2090: determined by $n$ points in the plane or in three dimensions. Our results improve
2091: upon the previous bounds, but, most likely, many of them are not
2092: asymptotically tight. This leaves many open problems of closing the
2093: respective gaps. 
2094: Even in cases where the bounds are asymptotically tight, such as
2095: those involving minimum-area triangles in two and three dimensions,
2096: determining the correct constants of proportionality still offers
2097: challenges.
2098: 
2099: Here is yet another problem on triangle areas, of a slightly
2100: different kind, with triangles determined by lines, not points
2101: (motivated in fact by the question of bounding $|U_2|$ in the
2102: proof of Theorem \ref{thm:unit2}). Any three nonconcurrent, and
2103: pairwise non-parallel lines in the plane determine a triangle of
2104: positive area. What is the maximum number of unit area
2105: triangles determined by $n$ lines in the plane?
2106: 
2107: 
2108: \begin{theorem}\label{thm:lines}
2109: The maximum number of unit-area triangles determined by $n$ lines in 
2110: the plane is $O(n^{7/3})$, and for any $n \geq 3$, there are $n$ lines that
2111: determine $\Omega(n^2)$ unit-area triangles. 
2112: \end{theorem}
2113: %
2114: \begin{proof}
2115: {\em Lower bound}: Place $n/3$ equidistant parallel lines at angles 
2116: $0$, $\pi/3$, and $2\pi/3$, through the points of an appropriate section of 
2117: the triangular lattice, and observe that
2118: there are $\Omega(n^2)$ equilateral triangles of unit side (i.e., of
2119: the same area) in this construction.
2120: 
2121: {\em Upper bound}: Let $L$ be a set of $n$ lines in the plane. 
2122: We define a variant
2123: of the hyperbolas used in the proof of Theorem \ref{thm:unit2}:
2124: For any pair of non-parallel lines $\ell_1,\ell_2\in L$, let
2125: $\gamma(\ell_1,\ell_2)$ denote the locus of points $p\in \RR^2$,
2126: $p\not\in\ell_1\cup \ell_2$, such that the parallelogram that has a
2127: vertex at $p$ and two sides along $\ell_1$ and $\ell_2$,
2128: respectively, has area $1/2$. The set $\gamma(\ell_1,\ell_2)$ is the
2129: union of two hyperbolas with $\ell_1$ and $\ell_2$ as asymptotes 
2130: (four connected branches in total).
2131: Any two non-parallel lines uniquely determine two such hyperbolas.
2132: Let $\Gamma$ denote the set of the branches of these hyperbolas, and
2133: note that $|\Gamma|=O(n^2)$.
2134: Observe now that, if $\ell_1$, $\ell_2$, and $\ell_3$ determine a
2135: unit area triangle, then $\ell_3$ is tangent to one of the two
2136: hyperbolas in $\gamma(\ell_1,\ell_2)$.
2137: 
2138: We first derive a weaker bound. Construct two bipartite graphs 
2139: $G_1,G_2\subseteq L\times\Gamma$. We put an edge $(\ell,\gamma)$ 
2140: in $G_1$ (resp., $G_2$) if $\ell$ is tangent to $\gamma$ and 
2141: $\ell$ lies below (resp., above) $\gamma$. The edges of $G_1$ 
2142: and $G_2$ account for all line-curve tangencies. 
2143: Observe that neither graph contains a $K_{5,2}$, 
2144: that is, there cannot be five distinct lines in $L$ tangent 
2145: to two branches of hyperbolas from above (or from below). 
2146: Indeed, this would force the two branches to intersect at five
2147: points, which is impossible for a pair of distinct quadrics.
2148: It thus follows from the theorem of 
2149: K\H{o}v\'ari, S\'os and Tur\'an~\cite{kst-54} (see 
2150: also~\cite[p.~121]{pa-95}) that the number of line-hyperbola tangencies 
2151: between any $n_0$ lines in $L$ and any $m_0$ hyperbolas in $\Gamma$ 
2152: is $O(n_0 m_0^{4/5}+m_0)$. With $n_0=n$ and $m_0=O(n^2)$, this already
2153: gives a bound of $O(n \cdot n^{8/5}+n^2)=O(n^{13/5})$ on the number of
2154: unit-area triangles determined by $n$ lines in the plane. We next
2155: derive an improved bound.  
2156: 
2157: Let $L$ be the given set of $n$ lines, and let $\Gamma$ be the
2158: corresponding set of $m=O(n^2)$ hyperbola branches. 
2159: We can assume that no line in $L$ is vertical, and apply a standard
2160: duality which maps each line $\ell\in L$ to a point $\ell^*$. A
2161: hyperbolic branch $\gamma$ is then mapped to a curve $\gamma^*$,
2162: which is the locus of all points dual to lines tangent to $\gamma$;
2163: it is easily checked that each $\gamma^*$ is a quadric.
2164: Let $L^*$ denote the set of the $n$ dual points, and let
2165: $\Gamma^*$ denote the set of $m=O(n^2)$ dual curves. 
2166: A line-hyperbola tangency in the primal plane is then
2167: mapped to a point-curve incidence in the dual plane. 
2168: 
2169: We next construct a $(1/r)$-cutting for $\Gamma^*$, partitioning
2170: the plane into $O(r^2)$ relatively open cells of bounded description
2171: complexity, each of which contains at most $n/r^2$ points and is
2172: crossed by at most $m/r$ curves. By using the previous bound for each
2173: cell, the total number of incidences involving points in the interior
2174: of these cells is
2175: % 
2176: $$ O\left( 
2177: r^2 \left(\frac{n}{r^2} \left(\frac{m}{r}\right)^{4/5} + \frac{m}{r}\right) 
2178: \right)= O\left(n \left(\frac{m}{r}\right)^{4/5} + mr\right).
2179: $$ 
2180: We balance the two terms by setting $r=n^{5/9}/m^{1/9}$, and observe
2181: that $1\le r\le m$ if $m\le n^5$ and $n\le m^2$; since
2182: $m=\Theta(n^2)$, both inequalities do hold in our case. Hence, the
2183: total number of incidences under consideration is
2184: $O(m^{8/9}n^{5/9})=O(n^{7/3})$.
2185: 
2186: It remains to bound the overall number of incidences involving points 
2187: lying on the boundaries of at least two cells. A standard
2188: argument, which we omit, shows that the number of these incidences
2189: is also $O(n^{7/3})$, and thereby completes the proof of the theorem.
2190: \end{proof}
2191: 
2192: Some remarks are in order: The line variant of unit-area triangle 
2193: problems is {\em not} equivalent to the point variant, under the 
2194: standard point-line duality.  Specifically: Let $S$ be a set of $n$ 
2195: points in the plane having distinct $x$-coordinates. Consider the 
2196: duality transform that maps a point
2197: $p=(a,b)$ to the line $p^*:\, y=ax-b$, and vice versa.
2198: %
2199: It is easy to see that
2200: there is no absolute constant $A>0$ such that, for $p,q,r \in S$,
2201: triangle $\Delta{pqr}$ has unit area if and only if the triangle
2202: $\Delta {p^*q^*r^*}$ formed by the three dual lines has area $A$.
2203: %
2204: 
2205: Yet, there is a connection between the point- and the line-variants of
2206: the unit-area problem in the plane. 
2207: Go back to the notation in the proof of Theorem \ref{thm:unit2},
2208: where, for a parameter $k \leq n^{1/3}$, we had $|U_1|= O(n^2 k)$.
2209: Recall that $U_2$ denotes the set of unit-area triangles where
2210: all three top lines are {\em $k$-rich}, and that there are 
2211: $|L_k|=O(n^2/k^3)$ such lines. Observe that the three top lines 
2212: of each triangle in $U_2$ determine a triangle of area $4$. We thus 
2213: face the question of bounding the number of triangles of area 4 
2214: determined by the $k$-rich lines in $L_k$. 
2215: By Theorem~\ref{thm:lines}, there are most $O((n^{2}/k^{3})^{7/3})$
2216: such triangles. Balancing $|U_1|$ with $|U_2|$ yields
2217: $k=n^{1/3}$, thereby implying that $|U_1|+|U_2| = O(n^{7/3})$.
2218: 
2219: We note that the bound $O(n^{44/19})$ of Theorem \ref{thm:unit2}
2220: could be re-derived with this new approach, if the bound of
2221: Theorem~\ref{thm:lines} could be improved to $O(n^{11/5})$.
2222: Moreover, an $o(n^{11/5})$ bound for the line-variant would in turn
2223: lead to an improvement in our current bound for the classical
2224: point-variant of the unit area problem in the plane.
2225: 
2226: 
2227: \begin{thebibliography}{10}\topsep=0pt\itemsep=-1pt\parsep=1pt
2228: 
2229: \bibitem{af-02}
2230: B.~M.~\'Abrego and S.~Fern\'andez-Merchant, Convex polyhedra in
2231: $\RR^3$ spanning $\Omega(n^{4/3})$ congruent triangles, {\em J.
2232: Combinat. Theory, Ser.~A} {\bf 98} (2002), 406--409.
2233: 
2234: \bibitem{at-07}
2235: E. Ackerman and G. Tardos,
2236: On the maximum number of edges in quasi-planar graphs,
2237: {\em J. Combinat. Theory, Ser. A} {\bf 114} (2007), 563--571.
2238: 
2239: \bibitem{aapps-97}
2240: P.~K. Agarwal, B. Aronov, J. Pach, R. Pollack, and M. Sharir,
2241: Quasi-planar graphs have a linear number of edges,
2242: {\em Combinatorica} {\bf 17} (1997), 1--9.
2243: 
2244: \bibitem{ANPPSS}
2245: P.~K.~Agarwal, E.~Nevo, J.~Pach, R.~Pinchasi, M.~Sharir, and S.~Smorodinsky,
2246: Lenses in arrangements of pseudocircles and their applications,
2247: {\it J. ACM} {\bf 51} (2004), 139--186.
2248: 
2249: \bibitem {acns-82}
2250: M.~Ajtai, V.~Chv\'atal, M.~Newborn, and E.~Szemer\'edi,
2251: Crossing-free subgraphs,
2252: {\em Annals Discrete Math.} {\bf 12} (1982), 9--12.
2253: 
2254: %\bibitem{as-05}
2255: %R.~Apfelbaum and M.~Sharir, Repeated angles in three and four
2256: %dimensions, {\em SIAM J. Discrete Math.} {\bf 19} (2005), 294--300.
2257: 
2258: %\bibitem{aks-05}
2259: %B.~Aronov, V.~Koltun, and M.~Sharir, Incidences between points and
2260: %circles in three and higher dimensions, {\em Discrete Comput. Geom.}
2261: %{\bf 33} (2005), 185--206.
2262: 
2263: \bibitem{ArS}
2264: B.~Aronov and M.~Sharir,
2265: Cutting circles into pseudo-segments and improved bounds for incidences,
2266: \textit{Discrete Comput. Geom.} {\bf 28} (2002), 475--490.
2267: 
2268: \bibitem{b-03}
2269: S.~Basu,
2270: The combinatorial and topological complexity of a single cell,
2271: {\em Discrete Comput. Geom.} {\bf 29} (2003), 41--59.
2272: 
2273: \bibitem{Baumann}
2274: E.~Baumann,
2275: Intersection of cylinders,
2276: {\tt http://private.mcnet.ch/baumann/all\_e.htm}
2277: 
2278: \bibitem{b-83}
2279: J.~Beck,
2280: On the lattice property of the plane and some problems of
2281: Dirac, Motzkin and Erd\H{o}s in combinatorial geometry,
2282: {\em Combinatorica} {\bf 3} (1983), 281--297.
2283: 
2284: %\bibitem{bk-03}
2285: %P.~Bra\ss\ and C.~Knauer, On counting point-hyperplane incidences,
2286: %{\em Comput. Geom.} {\bf 25} (2003), 13--20.
2287: 
2288: \bibitem{bmp-05}
2289: P.~Bra\ss , W.~Moser, and J.~Pach,
2290: {\em Research Problems in Discrete Geometry},
2291: Springer, New York, 2005.
2292: 
2293: \bibitem{brs-01}
2294: P.~Bra\ss, G.~Rote, and K.~J.~Swanepoel,
2295: Triangles of extremal area or perimeter in a finite planar point set,
2296: {\em Discrete Comput.  Geom.} ~{\bf 26} (2001), 51--58.
2297: 
2298: \bibitem{bp-79}
2299: G.~R.~Burton and G.~Purdy,
2300: The directions determined by $n$ points in the plane,
2301: {\em J.~London Math. Soc.} {\bf 20} (1979), 109--114.
2302: 
2303: \bibitem{ch-05}
2304: B.~Chazelle,
2305: Cuttings,
2306: In {\em Handbook of Data Structures and Applications}
2307: (D. Mehta and S. Sahni, editors), chap.~25, Chapman and Hall/CRC Press, 2005.
2308: 
2309: \bibitem{ceg-89}
2310: B.~Chazelle, H.~Edelsbrunner, L.~J.~Guibas, and M.~Sharir,
2311: A singly-expenential stratification scheme for real semi-algebraic
2312: varieties and its applications,
2313: in {\em Proc. 16th ICALP}, vol.~372 of
2314: LNCS, Springer, Berlin, 1989, pp.~179--193.
2315: 
2316: \bibitem{ceg-90}
2317: K.~L.~Clarkson, H.~Edelsbrunner, L.~G.~Guibas, M.~Sharir, and
2318: E.~Welzl, Combinatorial complexity bounds for arrangements of curves
2319: and spheres, {\em Discrete Comput. Geom.} {\bf 5} (1990), 99--160.
2320: 
2321: \bibitem{dpt-05}
2322: A.~Dumitrescu, J.~Pach, and G.~T\'oth,
2323: The maximum number of empty congruent triangles
2324: determined by a point set,
2325: {\em Revue Roumaine de Math. Pures et
2326: %{\em Revue Roumaine de Math\'ematiques Pures et
2327: Appliqu\'ees} {\bf 50} (2005), 613--618.
2328: 
2329: \bibitem {dt-07a}
2330: A. Dumitrescu and Cs.~D.~T\'oth,
2331: Distinct triangle areas in a planar point set,
2332: in {\em Proc. 12th Conf. on Integer Programming and
2333: Combinat. Opt.}, vol. 4513 of LNCS, Springer, 2007,
2334: pp.~119--129.
2335: 
2336: \bibitem{dt-07b}
2337: A. Dumitrescu and Cs. D. T\'oth,
2338: On the number of tetrahedra with minimum, uniform, and distinct volumes
2339: in 3-space,
2340: in {\em Proc. 18th ACM-SIAM Symp. Discrete Alg.}, 2007,
2341: ACM Press, pp.~1114--1123.
2342: 
2343: \bibitem{es-91}
2344: H.~Edelsbrunner ans M.~Sharir,
2345: A hyperplane incidence problem with applications to counting distances,
2346: in {\em Applied Geometry and Discrete Mathematics},
2347: vol.~4 of DIMACS Ser. Discrete Math. \& Theoret. Comput. Sci., AMS,
2348: Providence, 1991, pp.~253--263.
2349: 
2350: \bibitem {e-46}
2351: P. Erd\H{o}s,
2352: On sets of distances of $n$ points,
2353: {\em American Mathematical Monthly} {\bf 53} (1946), 248--250.
2354: 
2355: \bibitem{ep-71}
2356: P.~Erd\H{o}s and G.~Purdy,
2357: Some extremal problems in geometry,
2358: {\em J. Combinat. Theory} {\bf 10} (1971), 246--252.
2359: 
2360: %\bibitem{ep-75}
2361: %P.~Erd\H os and G.~Purdy, Some extremal problems in geometry III,
2362: %{\em Proc. 6th South-Eastern Conf. on Combinatorics, Graph Theory, and
2363: %Computing}, 1975, 291--308.
2364: 
2365: \bibitem{ep-76}
2366: P.~Erd\H os and G.~Purdy,
2367: Some extremal problems in geometry IV,
2368: {\em Congressus Numerantium} {\bf 17} (Proc. 7th South-Eastern Conf.
2369: on Combinatorics, Graph Theory, and Computing), 1976, 307--322.
2370: 
2371: %\bibitem{ep-77}
2372: %P.~Erd\H os and G.~Purdy, Some extremal problems in geometry V,
2373: %{\em Proc. 8th South-Eastern Conf. on Combinatorics, Graph Theory, and
2374: %Computing}, 1977, 569--578.
2375: 
2376: \bibitem{ep-95}
2377: P.~Erd\H os and G.~Purdy,
2378: Extremal problems in combinatorial geometry.
2379: in {\em Handbook of Combinatorics}
2380: (R. Graham, M. Gr\"otschel and L. Lov\'asz, editors),
2381: Vol. 1, 809--874, Elsevier, Amsterdam, 1995.
2382: 
2383: \bibitem{eps-82}
2384: P.~Erd\H os, G.~Purdy, and E.~G.~Straus, On a problem in
2385: combinatorial geometry, {\em Discrete Mathematics} {\bf 40} (1982),
2386: 45--52.
2387: 
2388: \bibitem{EPS:cyl}
2389: E.~Ezra, J.~Pach, and M.~Sharir,
2390: Regular vertices on the union of cylinders,
2391: manuscript, 2006.
2392: 
2393: \bibitem{hs-95}
2394: D.~Halperin and M.~Sharir,
2395: Almost tight upper bounds for the single cell and zone problems
2396: in three dimensions,
2397: {\em Discrete Comput. Geom.} {\bf 14} (1995), 385--410.
2398: 
2399: \bibitem{kpw-05}
2400: J.~K\'ara, A.~P\'or, and D.~R.~Wood,
2401: On the chromatic number of the visibility graph of a set of
2402: points in the plane,
2403: {\em Discrete Comput. Geom.} {\bf 34} (2005), 497--506.
2404: 
2405: \bibitem{kt-04}
2406: N. H. Katz and G. Tardos,
2407: A new entropy inequality for the Erd\H os distance problem,
2408: in {\em Towards a Theory of Geometric Graphs}
2409: (J. Pach, ed.), vol. 342 of Contemporary Mathematics, AMS, Providence,
2410: RI, 2004, 119--126.
2411: 
2412: %\bibitem{Ko}
2413: %V.~Koltun,
2414: %Almost tight upper bounds for vertical decomposition in four dimensions,
2415: %{\it J.~ACM} {\bf 51} (2004), 699--730.
2416: 
2417: \bibitem{kst-54}
2418: T.~K\H{o}vari, V.~T.~S\'os, and P.~Tur\'an,
2419: On a problem of K.~Zarankiewicz,
2420: {\em Colloquium Math.} {\bf 3} (1954), 50--57.
2421: 
2422: \bibitem{l-84}
2423: T. Leighton,
2424: New lower bound techniques for VLSI,
2425: {\em Math. Systems Theory} {\bf 17} (1984), 47--70.
2426: 
2427: \bibitem{MT}
2428: A.~Marcus and G.~Tardos,
2429: Intersection reverse sequences and geometric applications,
2430: {\em J. Combinat. Theory Ser. A} {\bf  113} (2006), 675--691.
2431: 
2432: \bibitem{m-02}
2433: J.~Matou\v{s}ek,
2434: {\em Lectures on Discrete Geometry},
2435: Springer-Verlag, Berlin, 2002.
2436: 
2437: \bibitem{pa-95}
2438: J.~Pach and P.~K.~Agarwal,
2439: {\em Combinatorial Geometry},
2440: Wiley-Interscience, New York, 1995.
2441: 
2442: \bibitem{ps-92}
2443: J.~Pach and M.~Sharir,
2444: Repeated angles in the plane and related problems,
2445: {\em J. Combinat. Theory Ser. A} {\bf 59} (1992), 12--22.
2446: 
2447: %\bibitem{ps-98} J.~Pach and M.~Sharir, On the number of
2448: %  incidences between points and curves, {\em Combin.  Probab.
2449: %    Comput.\/} {\bf 7} (1998), 121--127.
2450: 
2451: \bibitem{ps-04} J.~Pach and M.~Sharir, Geometric incidences, in
2452:   {\em Towards a Theory of Geometric Graphs\/}, AMS, vol. 342 of {\em
2453:     Contemp. Math.\/}, 2004, pp. 185--223.
2454: 
2455: \bibitem{p-07}
2456: R.~Pinchasi,
2457: The minimum number of distinct areas of
2458: triangles determined by a set of $n$ points in the plane,
2459: manuscript, 2007.
2460: 
2461: \bibitem{as-95}
2462: M.~Sharir and P.~K.~Agarwal,
2463: {\em Davenport-Schinzel sequences and their geometric applications},
2464: Cambridge University Press, New York, 1995.
2465: 
2466: \bibitem {st-01}
2467: J. Solymosi and Cs. D. T\'{o}th,
2468: Distinct distances in the plane,
2469: {\em Discrete Comput. Geom.} {\bf 25} (2001), 629--634.
2470: 
2471: \bibitem{sst-84}
2472: J. Spencer, E.~Szemer\'edi and W.~T.~Trotter,
2473: Unit distances in the Euclidean plane,
2474: in {\em Graph Theory and Combinatorics}, (B. Bollob\'as, editor),
2475: Academic Press, London, 1984, pp.~293--303.
2476: 
2477: \bibitem{s-97}
2478: L.~Sz\'ekely,
2479: Crossing numbers and hard {Erd\H os} problems in discrete geometry,
2480: {\em Combinat. Probab. Comput.} {\bf 6} (1997), 353--358.
2481: 
2482: \bibitem{st-83}
2483: E.~Szemer\'edi and W.~T.~Trotter,
2484: Extremal problems in discrete geometry,
2485: {\em Combinatorica} {\bf 3} (1983), 381--392.
2486: 
2487: \bibitem {t-03}
2488: G. Tardos,
2489: On distinct sums and distinct distances,
2490: {\it Advances in Math.} {\bf 180} (2003), 275--289.
2491: 
2492: \bibitem {v-54} I. M. Vinogradov,
2493: {\it Elements of Number Theory}, Dover Publications, New York, 1954.
2494: 
2495: 
2496: \end{thebibliography}
2497: 
2498: 
2499: \newpage
2500: \section*{Appendix}
2501: 
2502: \begin{proofof}{Lemma~\ref{3int}}
2503: Let us recall from~\cite{EPS:cyl} the structure of the
2504: intersection curve between two cylinders.  Let $C$ and $C'$ be two
2505: cylinders with nonparallel axes, so each pair of axes are either skew
2506: to each other or concurrent.  Let $\gamma$ denote the curve of their
2507: intersection.
2508: 
2509: To simplify the analysis, we assume, without loss of generality, that
2510: the axis $\alpha$ of $C$ is the $z$-axis and that its radius is $1$.
2511: Let $\alpha'$ and $\rho'$ denote respectively the axis and radius of
2512: $C'$. Let $\pi$ be the plane passing through $\alpha'$ and through the
2513: shortest segment $e$ connecting the axes $\alpha,\alpha'$. If
2514: $\alpha,\alpha'$ are skew lines, $e$ and $\pi$ are well defined.  If
2515: $\alpha$ and $\alpha'$ are concurrent, we take $\pi$ to be the plane
2516: passing through $\alpha'$ and orthogonal to the plane spanned by
2517: $\alpha$ and $\alpha'$.
2518: 
2519: Let $\sigma$ denote the ellipse $C\cap\pi$. We use a cylindrical
2520: coordinate system $\theta,z$ on $C$, and write the equation of
2521: $\sigma$ as $z=a\cos\theta+b\sin\theta+c$, where $z=ax+by+c$ is
2522: the quation of $\pi$.
2523: 
2524: As shown in \cite{EPS:cyl}, the equation of $\gamma$ is
2525: $$
2526: z=\sigma(\theta)\pm
2527:   \frac{1}{\sin\beta}\sqrt{(\rho')^2-d^2(\sigma(\theta),\alpha')} ,
2528: $$
2529: where $\beta$ is the angle between the axes. Moreover,
2530: $d(\sigma(\theta),\alpha')$, being the distance, within $\pi$,
2531: of a point on the ellipse $\sigma$ from the line $\alpha'$,
2532: can also be expressed as $|p\cos\theta+q\sin\theta+r|$, for
2533: appropriate parameters $p,q,r$.
2534: 
2535: Let now $C,C_1,C_2$ be three cylinders with no pair of parallel axes.
2536: Suppose to the contrary that $|C\cap C_1\cap C_2|\ge 9$.
2537: Let $\gamma_i$ denote the intersection curve $C\cap C_i$, for
2538: $i=1,2$. Write the equations of $\gamma_1,\gamma_2$ as
2539: $$
2540: z=a_i\cos\theta+b_i\sin\theta+c_i \pm
2541:   \frac{1}{\sin\beta_i}\sqrt{(\rho_i)^2-(p_i\cos\theta+q_i\sin\theta+r_i)^2},
2542: $$
2543: for $i=1,2$, with the appropriate parameters as above.
2544: We can re-parameterize these curves by putting $t=\tan(\theta/2)$
2545: and $w=z(1+t^2)$, to obtain two equations of the form
2546: \begin{eqnarray*}
2547: w & = & Q_1(t)\pm\sqrt{K_1(t)} \\
2548: w & = & Q_2(t)\pm\sqrt{K_2(t)} ,
2549: \end{eqnarray*}
2550: where $Q_1,Q_2$ are quadratic polynomials and $K_1,K_2$ are
2551: quartic polynomials. We are given that these two equations have
2552: at least $9$ common roots (it is easy to check that distinct roots of
2553: the original system are mapped to distinct roots of the new system).
2554: 
2555: If $Q_1(t)\equiv Q_2(t)$ then the common roots must satisfy
2556: $K_1(t)=K_2(t)$. Since there are at least $9$ such roots and this
2557: is a quartic equation, we must also have $K_1(t)\equiv K_2(t)$.
2558: 
2559: We will get to this case soon, but let us first consider the case
2560: $Q_1(t)\not\equiv Q_2(t)$. After squaring, the
2561: equations become
2562: \begin{eqnarray*}
2563: (w-Q_1(t))^2 & = & K_1(t) \\
2564: (w-Q_2(t))^2 & = & K_2(t) .
2565: \end{eqnarray*}
2566: Hence
2567: $$
2568: w = -\frac{K_2(t)-K_1(t)}{2(Q_2(t)-Q_1(t))} + \frac{Q_1(t)+Q_2(t)}{2} ,
2569: $$
2570: so $t$ must satisfy the equation
2571: \begin{equation} \label{deg8eq}
2572: \left(
2573:  -\frac{K_2(t)-K_1(t)}{2(Q_2(t)-Q_1(t))} + \frac{Q_2(t)-Q_1(t)}{2}
2574: \right)^2 = K_1(t) ,
2575: \end{equation}
2576: which is a polynomial equation of degree at most $8$. Since it has $9$
2577: roots, it must vanish identically.
2578: 
2579: Since the left-hand side of (\ref{deg8eq}) is a square, $K_1$ must
2580: also be a square.  However, $K_1(t)$ is proportional to
2581: $$
2582: \biggl(\rho_1(1+t^2)\biggr)^2 -
2583: \biggl(p_1(1-t^2)+2q_1t+r_1(1+t^2)\biggr)^2 =
2584: $$
2585: $$
2586: \biggl(\rho_1(1+t^2)-(p_1(1-t^2)+2q_1t+r_1(1+t^2))\biggr) \cdot
2587: \biggl(\rho_1(1+t^2)+(p_1(1-t^2)+2q_1t+r_1(1+t^2))\biggr) .
2588: $$
2589: It follows that either each of these factors is a square, or they
2590: are multiples of each other. In the former case, we must have
2591: \begin{eqnarray*}
2592: q_1^2 & = & (\rho_1+p_1-r_1)(\rho_1-p_1-r_1) = (\rho_1-r_1)^2-p_1^2 \\
2593: q_1^2 & = & (\rho_1-p_1+r_1)(\rho_1+p_1+r_1) = (\rho_1+r_1)^2-p_1^2 ,
2594: \end{eqnarray*}
2595: implying that $\rho_1-r_1 = \pm (\rho_1+r_1)$, so either $\rho_1=0$
2596: or $r_1=0$. The first equality is impossible---our cylinders have
2597: positive radii. The second equality implies that
2598: $\rho_1^2=p_1^2+q_1^2$. However, as argued in \cite{EPS:cyl}, by
2599: shifting $\theta$, we may assume that $q_1=0$ and $p_1$ is half the
2600: major axis of $\sigma_1$. This implies that $\sigma_1$ is a circle
2601: (since its minor axis is always equal to $2\rho_1$),
2602: which can happen only when $\alpha_1$ is orthogonal to $\alpha$.
2603: Moreover, $r_1=0$ implies that $\alpha$ and $\alpha'$ are concurrent.
2604: 
2605: In the latter case, since $\rho_1\ne 0$, the two factors are
2606: proportional to each other only when $p_1(1-t^2)+2q_1t$ is a
2607: multiple of $1+t^2$, which can only happen when $p_1=q_1=0$,
2608: which again is impossible.
2609: 
2610: Since the only remaining case is that of orthogonal concurrent axes,
2611: it follows, using a
2612: symmetric argument, that in the only remaining case, the three axes
2613: $\alpha,\alpha_1,\alpha_2$ are concurrent, at a common point, and
2614: mutually orthogonal. It is easily checked that in this case the cylinders
2615: can intersect in at most $8$ points, contrary to assumption.
2616: (This special case of three intersecting cylinders has been studied
2617: a lot; see, e.g.,~\cite{Baumann}.)
2618: 
2619: Hence, $Q_1(t)\equiv Q_2(t)$ and $K_1(t)\equiv K_2(t)$. However, the
2620: first identity implies that $\sigma_1=\sigma_2$,
2621: so the plane containing the axis of $C_1$
2622: also contains the axis of $C_2$. Since these axes are nonparallel,
2623: they must be concurrent. Since the analysis is fully symmetric with
2624: respect to the three cylinders, it follows that all three axes are
2625: either coplanar or concurrent. If they are coplanar but not concurrent,
2626: then it is easy to check that the planes $\pi_1$ and $\pi_2$ (with
2627: respect to $C$ as the ``base'' cylinder) cannot be equal. If the
2628: three axes are concurrent then again the identity of the planes
2629: $\pi_1,\pi_2$ implies that both $\alpha_1$ and $\alpha_2$ must be
2630: orthogonal to $\alpha$, and the fact that the argument is fully symmetric
2631: implies that all three axes must be concurrent and mutually orthogonal,
2632: a case that we have already ruled out.
2633: %
2634: This completes the proof of Lemma~\ref{3int}.
2635: \end{proofof}
2636: 
2637: 
2638: \end{document}
2639: 
2640: