1: \documentclass[article,oneside,12pt]{article}
2:
3: \input{epsf}
4: \usepackage{amsmath}
5: \usepackage{amsthm}
6: \usepackage{amsfonts}
7: \usepackage{amssymb}
8: \usepackage{mathrsfs}
9: \usepackage[dvips]{graphics}
10: \usepackage{graphicx}
11: \usepackage{float}
12: \usepackage{anysize}
13: \setlength{\topmargin}{-0.5in}
14:
15: \setlength{\textheight}{9.8in}
16:
17: \newtheorem{prop}{Proposition}
18: \newtheorem{lem}{Lemma}
19: \newtheorem{df}{Definition}
20: \newtheorem{theorem}{Theorem}
21:
22:
23:
24: %\newcommand\vex#1{\textbf{\textit{#1}}}
25: \newcommand\tdots{,\dots ,}
26: \newcommand\rstr{R^{*}}
27: \newcommand\muN{\mu^{\otimes N}}
28: \newcommand{\xb}{\textbf{\textit{x}}}
29: \newcommand{\R}{\mathbb{R}}
30: %\newcommand{\xb}{\mbox{{\boldmath$x$}}}
31:
32:
33: \begin{document}
34:
35: \title {Simulation Studies of Some Voronoi Point Processes%
36: \footnote{This research was supported by the ARC Centre of Excellence for Mathematics
37: and Statistics of Complex Systems.}}
38:
39: \author{K.A. Borovkov\footnote{University of Melbourne, email: k.borovkov@ms.unimelb.edu.au}
40: and D.A. Odell\footnote{MASCOS, Melbourne, email: d.odell@ms.unimelb.edu.au}}
41:
42:
43: \maketitle
44:
45: %\parindent = 1em
46: %\parskip = 12pt
47:
48: \begin{abstract}
49: We introduce a new class of dynamic point process models with simple and intuitive
50: dynamics that are based on the Voronoi tessellations generated by the processes. Under
51: broad conditions, these processes prove to be ergodic and produce, on stabilisation, a
52: wide range of clustering patterns. In the paper, we present results of simulation
53: studies of three statistical measures (Thiel's redundancy, van Lieshout and Baddeley's
54: $J$-function and the empirical distribution of the Voronoi nearest neighbours' numbers)
55: for inference on these models from the clustering behaviour in the stationary regime.
56: In particular, we make comparisons with the area-interaction processes of Baddeley and
57: van Lieshout.
58: \end{abstract}
59:
60: \section{Introduction}
61:
62:
63: In many diverse fields such as biology, geography, business and telecommunications,
64: spatial point configurations evolve according to criteria dependent on a {\em zone of
65: influence\/} in some general sense. A natural way of formalising this concept is based
66: on using Voronoi tessellations~(see e.g.~\cite{OBS}).
67:
68: Let $(M,d)$ be a metric space and $N>0$ the initial number of points in our process.
69: The associated configuration space $\mathscr{X}$ consists of all $N$-point subsets of
70: $M$: $\mathscr{X}:=\{\xb\subset M :\, \text{card} (\xb)=N \}$ (`multiple points' will
71: be a.s. impossible in our models). For an $\xb\in \mathscr{X}$, the {\em Voronoi
72: cell\/} of $x_i$ relative to $\xb$ is defined as
73: $$
74: C^{\xb}_{x_i}:= \Bigl\{y \in M: \, d(y,x_i) = \min_{j\le N} d(y,x_j) \Bigr\}.
75: $$
76: The set $\mathscr{T}_{\xb}= \{C^{\xb}_{x_i}: \, x_i\in \xb \}$ is called the {\em
77: Voronoi tessellation\/} generated by $\xb$.
78:
79:
80:
81:
82: We define a Voronoi point process as a discrete-time Markov processes $\{\xb_n
83: \}_{n\geq 0}$ with values in $\mathscr{X}$ which evolves as follows: at each step, a
84: point is chosen from the current configuration $\xb_n$ according to a probability rule
85: determined by $\mathscr{T}_{\xb_n}$ and removed from the configuration, and at the same
86: time a new point is added at a random location, according to a fixed probability
87: measure $\mu$ on $M$. Our initial interest interest in such dynamics was prompted by
88: its relevance to some of the real-life processes in the above-mentioned application
89: areas. In addition, simulations showed that processes constructed according to this
90: model display very interesting forms of clustering behaviour, and constructing models
91: of point processes producing desired clustering patterns is of substantial interest both
92: theoretically and practically.~(see e.g.~\cite{BV95, OBS}).
93:
94: Despite the fact that it is notoriously difficult to obtain analytic results for
95: processes of such a nature, we were able to prove ergodicity in a number of interesting
96: cases. To formulate the results, we first need to further specify the `culling rule' in
97: the model.
98:
99: We assume that a non-negative `selection function' $S$ is given on the set of all
100: possible Voronoi cells in $M$. Then, given that $\xb=\{x_{1},\dots ,x_{N}\}$ is the
101: current configuration of the point process, in the next step a random point $x_{J}\in
102: \xb$ is chosen to be removed, with the distribution of the index $J$ given by
103: \begin{equation}
104: \label{eq2}
105: \mathbf{P} (J=j|\xb)
106: :=\frac{S(C_{x_{j}}^{\xb})}{\sum_{i=1}^{N}S(C_{x_{i}}^{\xb})},
107: \qquad 1\leq J \leq N.
108: \end{equation}
109: The function $S$ can be based on different properties of a Voronoi cell. In this note,
110: we consider only two of them: the `volume' of the cell and the number of its edges.
111: More precisely, we introduce the following two classes of Voronoi point processes:
112:
113: \medskip
114:
115: (A) The volume-based Voronoi point process, or $v$-process, on $M=\mathbb{S}^1$ or $M=
116: [0,1]^2$. Let $\lambda$ be the `volume measure' on $M$ (length on $\mathbb{S}^1$, area
117: on $[0,1]^2$). We assume that the value of the selection function $S(C_{x_{j}}^{\xb})$
118: is determined by the volume of the cell $C_{x_{j}}^{\xb}$: for a function
119: $S_v:\R^{+}\mapsto\R^{+}$, one puts $S(C_{x_{j}}^{\xb}):=S_{v}(\lambda
120: (C_{x_{j}}^{\xb}))$. If $S_v$ is increasing, then points with Voronoi cells of large
121: volume are more likely to be culled, and so the selection favours points with small
122: cells, \text{i.e.} points restricted by `close neighbours'. A decreasing $S_v$ favours
123: points with large cells. Functions of the form $S_v(u)=u^{\alpha}$, $\alpha \in \R$,
124: produce scale-independent models (in this case the dynamics of the process will clearly
125: be invariant under scale transformations of~$M$).
126:
127: \medskip
128:
129: (B) The neighbour-based Voronoi point process, or $n$-process, on $M= [0,1]^2$. The set
130: $\xb[x_{j}]$ of Voronoi-nearest neighbours of a point $x_j\in \xb$ is defined as the
131: collection of those generators $x_k\in \xb$ whose Voronoi cells share an edge with
132: $C_{x_{j}}^{\xb}$:
133: $$
134: \xb[x_{j}] :=\Bigl\{x_{k}\in \xb : \, k\neq j,\, {\rm card}(C_{x_{k}}^{\xb} \cap
135: C_{x_{j}}^{\xb})>1 \Bigr\}.
136: $$
137: We assume that, for a function $S_n:\{1\tdots N\}\mapsto\R^{+}$, we have
138: $S(C_{x_{j}}^{\xb})=S_{n}({\rm card}( \xb[x_{j}] ))$.
139:
140: \medskip
141: In both cases (A) and (B), we assume the placement probability $\mu=\lambda$.
142:
143:
144: \begin{theorem}
145: {\em (i)} The neighbour-based Voronoi point process with a positive selection function
146: $S_n: \{1,\dots , N \}\mapsto \R^+$ is Harris ergodic.
147:
148: {\em (ii)} The volume-based Voronoi point process with a selection function
149: $S_v:(0,\lambda(M)]\mapsto\R^+$, such that both $S_v$ and $1/S_v$ are bounded on closed
150: subsets, is Harris ergodic.
151: \end{theorem}
152:
153: Recall that Harris ergodicity entails convergence to the stationary distribution in
154: total variation (see e.g. p.560 in~\cite{MeTw92} or p.154 in~\cite{As87}). For the
155: proof of the theorem, see~\cite{borode}.
156:
157:
158:
159: This result justifies the approach taken in the present note, which is devoted to
160: simulation studies of point patterns emerging in stationary regime in the evolution of
161: Voronoi point processes. More specifically, we are interested in the performance of
162: three relatively simple statistics one could employ to characterise the resulting
163: stationary distributions. These are Thiel's redundancy measure (which is essentially
164: the relative entropy for the set of cell volumes) and, in the two-dimensional case, the
165: distribution of the number of Voronoi nearest neighbours for a random cell and the
166: $J$-function of Baddeley and van Lieshout (which provides a comparison of the
167: environment of a `random point' of the configuration with that of a `random point' of
168: the underlying space over a range of scales). We demonstrate that the $n$- and
169: $v$-processes produce (for different choices of parameters) quite different clustering
170: patterns, and that the combination of the above three measures work reasonably well in
171: distinguishing between different Voronoi processes, and also between Voronoi processes
172: and the area-interaction processes of Baddeley and Van Lieshout~\cite{BV95}.
173:
174:
175:
176:
177:
178:
179:
180:
181:
182: \section{Simulation Studies}
183: \label{s4}
184: \subsection{One-dimensional $v$-processes}
185:
186: The simple model of scale-free process on $M=\mathbb{S}^1$ with $S_v(u)=u^{\alpha}$
187: demonstrates that very interesting dynamics and point patterns arise even in one-dimension.
188: This includes a `phase change' observed when varying the value of
189: $\alpha$. Figure~\ref{3oneD} depicts side-by-side realisations of three different
190: $v$-processes of this type, each having the same number of points $N=128$, but
191: different $\alpha$ values. The base of each rectangle represents the circle
192: $\mathbb{S}^1$ opened out into a line segment by a cut.
193:
194: \begin{figure}[htbp]
195: \centering
196: \includegraphics{3oneD.eps}
197: \caption{\footnotesize Evolution of $v$-processes on circle. Each of the three processes
198: (with the selection functions $S_v(u)=u^{\alpha}$ with (a) $\alpha = -1.0$, (b) $\alpha = 0.5$, (c) $\alpha=1.5,$ resp.)
199: was run with $N = 128$ points for $T = 4096$ steps (the vertical axis represents time). }
200: \label{3oneD}
201: \end{figure}
202:
203:
204:
205: We see a dramatic phase shift in behaviour between (b) and (c) that can be located more
206: precisely at $\alpha=1$; its sharpness was found to increase with~$N$. When $\alpha
207: >1$, a cluster forms whose stability also increases with $N$. When $\alpha \leq
208: 1$, we observe a degree of clustering in $M$ that varies with $\alpha$.
209:
210: When $\alpha=0$, the points are uniformly distributed, so the associated Voronoi cell
211: volumes have pretty nearly a Gamma distribution with shape parameter 2 (as cell-width
212: is half the sum of two consecutive uniform spacings). For $\alpha \leq 1$, the
213: histogram of cell volumes stabilises after a large number of steps to a result
214: well-fitted by a Gamma distribution with a shape parameter varying inversely with
215: $\alpha$. This suggests that the maximum likelihood estimator for the Gamma shape
216: parameter for the cell-volume data could be a suitable statistic for inference on
217: $\alpha$. Furthermore, under the assumption of Gamma distribution, the shape parameter
218: can equivalently be estimated from the entropy---and the latter can also be used
219: without the above assumption, namely in the form of {\em Thiel's redundancy measure}
220: which is introduced as follows.
221:
222:
223:
224: For a fixed configuration $\xb,$ let $p_j := \lambda (C^{\xb}_{x_j})/\lambda(M),$ $
225: 1\leq n\leq N$, be the probability that a random uniformly distributed point in $M$
226: lies in~$C_{x_{i}}^{\xb}$. Then Thiel's redundancy measure $R^*(\xb)$ of $\xb$ is
227: defined~\cite{LE79, OBS} as the entropy of the distribution $\{p_j\}_{j=1}^N$ relative
228: to the uniform one on $\{1,\dots, N\}$:
229: $$
230: R^*(\xb):= \text{ln}N+\sum_{j=1}^N p_j \ln p_j.
231: $$
232: Simulations show that this statistic works quite well over a reasonably large range of
233: $\alpha$ values: there is a very small variance in its values when $N\ge 10^3$, and a
234: strong enough dependence on~$\alpha$ to use $R^*$ for reliable estimation of the
235: parameter. Figure\,\ref{Rstar1D} displays (connected) average values of 25 independent
236: realizations for each of the values $\alpha = -2 + 0.3k,$ $0\le k\le 10,$ of
237: $R^*(\xb_T)$ after a large number of steps~$T$ (cf.~Fig.\,\ref{Rstarevo}).
238: \begin{figure}[h]
239: \centering
240: \includegraphics{Rstar1D.eps}
241: \caption{\footnotesize Empirical Thiel's redundancy measure for $v$-processes on
242: $\mathbb{S}^1$ with
243: $S_v(u)=u^{\alpha}$, $N=10^3$.
244: The dotted lines show the 95\%\ probability intervals.}
245: \label{Rstar1D}
246: \end{figure}
247:
248: The behaviour of the statistic $R^* (\xb_T)$ as $T$ increases could also be used to
249: estimate the rate of convergence to stationarity. Figure\,\ref{Rstarevo} (left) gives
250: an indication of that rate for a range of $\alpha$ values. It shows that the rate is
251: quite high for $\alpha <0.5$, whereas when $\alpha$ approaches one, it takes the
252: process much longer to settle in a stable regime, and also that the oscillations of
253: $\{R^* (\xb_T)\}_{T\ge 0}$ are higher for those values of~$\alpha$. At the threshold
254: value $\alpha=1$, the sequence oscillates in quite diverse ranges of values
255: demonstrating metastable beahviour, see Fig.\,\ref{Rstarevo} (right).
256:
257:
258: \begin{figure}[!ht]
259: \centering
260: \includegraphics{rst1Comb.eps}
261: \caption{\footnotesize Evolution of $\{R^* (\xb_T)\}_{T\ge 0}$ for $v$-processes on
262: $\mathbb{S}^1$:
263: $N=10^3$, $S_v(u)=u^{\alpha}$. Left: $\alpha =-1;$ $-0.53;$ $-0.05;$ $0.19;$
264: $0.45;$ $0.66;$ $0.9,$ on the time interval $T\le 10 N.$ Right: $\alpha=1$, on the time interval $T\le 10^3 N.$}
265: \label{Rstarevo}
266: \end{figure}
267:
268:
269:
270:
271:
272:
273:
274: \subsection{Two dimensions: edge effects}
275:
276: The problem of `edge effects' (where the region boundary proximity can affect the
277: shape/size of Voronoi cells) in simulations of Voronoi tessellations in (a rectangle)
278: $M\subset\R^2$ is commonly dealt with by treating the underlying space as a torus or by
279: considering only those parts of $M$ that lie within a window significantly distant from
280: the edge $\partial M$ of~$M$. This approaches is satisfactory for `static' point
281: processes~\cite{OBS}. In the case of the Voronoi processes, edge effects may propagate
282: in the course of the evolution, which presents an interesting problem by itself.
283:
284: Let $\xb\subset M= [0,1]^2$ be finite, $A\subset \R^2$. The Voronoi tessellation
285: $\mathscr{T}_{\xb}$ induces a nearest-neighbour (NN) distance $d^{\xb}_{N\! N}(x_j,A)$
286: between $x_j\in \xb$ and $A$, defined as the length of the shortest `path' $x_j=x(1),
287: \dots, x(i)\in\xb$ with the properties that the Voronoi cells of $x(m)$ and $x(m+1)$
288: have a common edge, $1 \le m <i,$ and $C^{\xb}_{x(i)}\cap A\neq \varnothing$. We can
289: gauge the significance of edge effects by observing the changes in $\rstr$ when we
290: restrict attention to $x_j\in\xb$ with $d^{\xb}_{N\!N}(x_j, \partial M)\ge m$ for a
291: fixed $m>0$. Edge effects were assessed by a two-way ANOVA using NN distance as one
292: factor, and the selection function as the other. In the case of the $v$-process with
293: $S_v(u)=u^{\alpha}$, we found that edge effects were only significant in the layer of
294: cells adjacent to the boundary, without any significant effect from~$\alpha$. For the
295: $n$-processes, significant edge effects were detected only for the `anti-few' and
296: `anti-many' selection functions (see subsection~\ref{s5} below), with no significant
297: propagation beyond the NN-depth of two.
298:
299: To minimise edge effects in our results, the statistics were generally computed from the
300: set of cells with $d^{\xb}_{N\!N}(x_j,\partial M)\geq 3$.
301:
302:
303:
304:
305:
306:
307:
308: \subsection{Two-dimensional $n$-processes}
309: \label{s5}
310:
311:
312: In this family of Voronoi processes, the evolution is driven by one of the most natural
313: local characteristics: the number of neighbours of the process points, \text{i.e.} the
314: number of the edges of their Voronoi cells. For non-degenerate random configurations in
315: $\mathbb R^2$, the vertices of the Voronoi tessellation \text{a.s.} terminate three
316: edges, so by Euler's theorem the mean number of edges for a cell is six. There can be
317: cells with any number of edges $m\ge 3$, although in our simulations cells with $m> 13$
318: were rare, with relative frequency $<10^{-3}$.
319:
320: We studied the $n$-processes on $M=[0,1]^2$ with the following selection functions:
321: (i)~$S_n (n)=n$ (`vanilla', in which cells with a large number of neighbours are more
322: likely to be culled); (ii)~$S_n (n)=n^2$ (`anti-many', which more severely penalises
323: cells with large numbers of neighbours); (iii)~$S_n (n)=(n-2)^{-2}$ (`anti-few', which
324: does just the opposite); (iv)~$S_n (n)=(0.1+|n-6|)^{-2}$ (`anti-6', which penalises
325: cells with six or close to six neighbours); (v)~$S_n (n)=|n-6|^2$ (`pro-6',which does
326: the opposite); and `sharp filters' focussing on cells with a given number neighbours:
327: (vi)~$S_n (5)=5000$ and $ S(n)=1$ if $n\neq 5$ (`anti-5'); and (vii)~$S_n (5)=1$ and
328: $S_n (n)=5000,$ if $n \neq 5$ (`pro-5'), and in addition the `anti-' and `pro-'
329: selectors for four and seven.
330:
331: The statistics used in the study included:
332:
333: \smallskip
334:
335: (a)~the empirical probability mass function (EPMF) for the number of Voronoi NN's,
336:
337: (b)~Thiel's redundancy measure, and also
338:
339: (c)~the $J$-function of van Lieshout and Baddeley~\cite{BV95}.
340:
341: \smallskip
342:
343: The $J$-function $J(r)$ compares the `environment' of a `typical generator' of the
344: process with that of a `random point' in the underlying space, and is defined as the
345: ratio of the probabilities that a disk of radius $r$ centred at the given point is empty
346: of (other) generators of the process.
347:
348: Formally, let $\xb$ be a stationary isotropic point process in $\R^2$. Set
349: $$
350: B(r):=\{x\in \mathbb R^2:\, |x|\le r\},\quad
351: F(r):=\mathbf{P}\bigl(\xb\, \cap B(r) \neq \varnothing \bigr),\quad
352: G(r):= \mathbf{P}^{!0}\bigl({\xb} \cap B(r) \neq \varnothing \bigr),
353: $$
354: where $\mathbf{P}^{!0}$ is the reduced Palm distribution of $\xb$ (the conditional
355: distribution of $\xb \backslash \{0 \}$ given that there is a point at the origin).
356: Then the $J$-function is defined as
357: \begin{equation}
358: \label{jfunc}
359: J(r):=\frac{1-G(r)}{1-F(r)} \quad \text{ for all } r\geq 0\text{ \ such that } F(r)<1.
360: \end{equation}
361: In the case of a Poisson point process of constant intensity, or complete spatial
362: randomness (CSR), clearly $J(r)\equiv 1$, see e.g.~\cite{Amb02}. If there is a
363: tendency towards clustering obesrvable at scale $r$, then $J(r)<1$, while $J(r)>1$ if
364: there is a tendency toward regular spacing and $r$ is of the order of the average
365: distance between neighbours.
366:
367:
368: To estimate $J(r)$, one uses empirical distribution functions $\hat{F}(r)$ and
369: $\hat{G}(r)$:
370: $$
371: \hat{J} (r) :=\frac{1-\hat{G}(r)}{1-\hat{F}(r)}.
372: $$
373: This is only reliably computable when $F(r) \leq 0.85$. Moreover, simulations showed
374: that the values of~$\hat{J}$ have relatively large variances when $N\asymp 10^3$, so
375: we averaged the $\hat{J}$ values over 25--75 independent draws of the process, and
376: performed a weighted cubic regression on the resulting points (with weights equal to
377: the reciprocals of the standard deviations; higher order regressions add little to the
378: goodness of fit).
379:
380: The results of our study for a selection of the above-listed $n$-processes are
381: summarised in Fig.\,\ref{nprocsComb} (statistics (c), (a)) and Table\,\ref{tabrst}
382: (statistic~(b)).
383:
384:
385:
386: \begin{figure}[!h]
387: \centering
388: \includegraphics{nprocsComb.eps}
389: \caption{\footnotesize
390: Comparison of $n$-processes ($N= 2000$, after $T=12N$ steps). Left: smoothed plots of $\ln \hat{J} (r)$.
391: Right: Proportions of cells with given numbers of neighbours (given
392: at the right-hand end).
393: Standard error is \text{approx.} 1\% of the given value in each case.}
394: \label{nprocsComb}
395: \end{figure}
396:
397:
398:
399:
400:
401: \begin{table}[!h]
402: \begin{center}
403: \begin{tabular}{c|c|c|c}
404: \hline
405: $\vphantom{\frac{1}{1}}$ $n$-process & $\overline R^*$ & s.d.($\rstr$) &s.e.($\overline R^*$)\\
406: \hline
407: vanilla& 0.1450& 0.0068& 0.0014\\
408: anti-many& 0.1548& 0.0073& 0.0015\\
409: anti-few& 0.1074& 0.0047& 0.0009\\
410: pro-6& 0.1615& 0.0083& 0.0012\\
411: anti-6& 0.1524& 0.0059& 0.0008\\
412: pro-5& 0.1479& 0.0067& 0.0013\\
413: anti-5& 0.1255& 0.0063& 0.0013
414: \end{tabular}
415: \end{center}
416: \caption{\footnotesize $\rstr$ values for various $n$-processes, with standard
417: deviations and standard errors. The values $\overline R^*$ are averages of 25
418: independent draws of $R^*$ ($N=2000$, after $T=12N$ steps). For CSR, $R^*=0.135.$}
419: \label{tabrst}
420: \end{table}
421:
422: The $J$-function curves provide the most nuanced `picture' of the distribution
423: patterns, but these curves were obtained by averaging over a number of observations of
424: highly variable statistics and are not easily applied in practice. The statistic
425: $\rstr$ provides a crude but fairly sensitive measure of deviation from CSR. The NN
426: EPMF is probably the most effective statistic for the $n$-processes, but it doesn't
427: distinguish well between CSR and `anti-many' models (and also between `anti-few' and
428: `pro-6'), whereas the $J$-curves for these models are very different. The $J$-curves
429: for these and for the `sharp filters' on four and seven neighboured cells reveal the
430: interesting fact that selection functions favouring cells with $m>6$ ($m<6$) Voronoi
431: neighbours produced more (less, resp.) uniform [than CSR] configurations. The
432: `anti-few' and `anti-5' processes produce very similar $J$-curves, but are
433: distinguished by their effects on $\rstr$: `anti-few' gives $\rstr= 0.107$, while
434: `anti-5' has $\rstr=0.126$, which are significantly different from each other as well
435: as being significantly below the CSR value of 0.135. The low value for $\rstr$
436: corresponds to less variablity in cell areas, so this is consistent with the upward
437: $J$-curves. Notice also the unusual curvature of the `anti-6' curve, a result which was
438: confirmed by increasing the number of independent draws in this case to 75. In all
439: these cases, the $\rstr$ and $\hat J(r)$ statistics gave consistent results,
440: \text{i.e.} upward curves corresponded to $\rstr<0.135$.
441:
442:
443:
444:
445:
446: \subsection{Two-dimensional $v$-processes}
447:
448:
449:
450: We study here the $v$-processes on $M=[0,1]^2$ with the scale-free selection functions
451: $S_v(u)=u^{\alpha}$. Figure\,\ref{vprocmatrix} shows a few examples of point patterns
452: after a considerable period of evolution. There is a continuous gradation of patterns,
453: from almost regularly spaced ones for $\alpha<0$ through increasing levels of clustering
454: until we reach the threshold level at $\alpha=1$. For $\alpha>1$, the patterns become
455: increasingly concentrated in a small region, usually in a corner of the square, as the
456: number of steps increases.
457:
458: \begin{figure}[!h]
459: \centering
460: \includegraphics{vprocmatrix.eps}
461: \caption{\footnotesize Realisations of the $v$-process with
462: $S_v(u)=u^{\alpha}$ for some $\alpha$ values. Left-right, top-bottom:
463: $\alpha=-3;$ 0.2; 1; 1.4. In each case, $N=2000$, $T=12N$.}
464: \label{vprocmatrix}
465: \end{figure}
466:
467: For $\alpha \leq 1$ we can employ the same statistics as for the $n$-process. As in the
468: one-dimensional case, Thiel's redundancy measure $\rstr$ serves well for inferring
469: the value of $\alpha$, and also provides a good indicator of the stabilisation of the
470: process, see~Fig.\,\ref{vprocComb1}. From Figs.\,\ref{vprocComb1},\,\ref{Rstarevo} it
471: appears that the two-dimensional processes settle into equilibrium at roughly the same
472: rate as the one-dimensional processes. The $\ln \hat J(r)$ curves for $\alpha\in
473: (-3,1)$ form a fan as expected, see Fig.\,\ref{vprocComb2}\,(left). More detailed
474: analysis reveals surprising (statistically significant) changes in curvature that occur
475: for $\alpha$ between $-0.3$ and $-0.2$ and between 0.5 and 0.6: it appears that a
476: subtle phase change does indeed occur around these values. In
477: Fig.\,\ref{vprocComb2}\,(right) we have plotted the NN EPMF as a function of $\alpha$.
478: Note that, as $\alpha$ increases (and so the amount of clustering), the proportions of
479: cells with different numbers of edges tend to come together; in other words, the cell
480: geometries become more variegated.
481:
482:
483:
484: \begin{figure}[!h]
485: \centering
486: \includegraphics{vprocComb1.eps}
487: \caption{\footnotesize
488: Thiel's redundancy measure for $v$-processes on $[0,1]^2$ ($N=2000$,
489: $S_v(u)=u^{\alpha}$). Left: Evolution of $\{\rstr (x_T)\}_{T\le 12N}$ for
490: $\alpha=-3;$ $-2.2$; $-1$;$-0.2$; $0.2$; $0.6$; 1.
491: Right: $\rstr$ vs $\alpha$ (after $T/N=12$ steps).}
492: \label{vprocComb1}
493: \end{figure}
494:
495:
496: \begin{figure}[!h]
497: \centering
498: \includegraphics{vprocComb2.eps}
499: \caption{\footnotesize
500: Graphical summary of results for $v$-processes on $[0,1]^2$ with
501: $S_v(u)=u^{\alpha}$ for different $\alpha$ values ($N=2000$, $T=8N$).
502: Left: Smoothed curves for $\ln \hat J(r)$
503: (numbers of independent draws range from 9 ($\alpha \leq -1.0$) to 59
504: ($\alpha = 0.5, 0.6$)). Note the rapid change in the curve as $\alpha$ approaches 1.
505: Right: Proportions of cells with given numbers of neighbours (given
506: at the right-hand end) as
507: functions of $\alpha$.}
508: \label{vprocComb2}
509: \end{figure}
510:
511:
512:
513:
514:
515: \section{Comparison with the area-interaction process}
516:
517:
518:
519: The area-interaction point process (AIPP) of Baddeley and van Lieshout~\cite{BV95} is a
520: model which successfully produces a range of different clusterings. Kendall~\cite{K98}
521: presented a practical method for `perfect simulation' of this process. This is
522: summarised in Ambler~\cite{Amb02} from whose detailed account we derived our computer
523: program for simulations.
524:
525: Let $M=[0,1]^2$ and $\mathscr{M}$ be the space of all finite configurations of points
526: in~$M$. The AIPP is specified by the density of its distribution with respect to that
527: of the unit rate Poisson process on $M$. Let $\lambda$ be Lebesgue measure on $M$,
528: $G:=B(\rho)$ for a fixed $\rho>0$, and $\xb\oplus G := \left\{z:\, z=x+g, x\in \xb,
529: g\in G\right\}$ for $\xb\in \mathscr{M}$. The above density is given by
530: $$
531: p(\xb)=C \beta^{\,\text{card} (\xb)}\gamma^{-\lambda(\xb\oplus G)},
532: $$
533: where $\beta, \gamma>0$ are parameters, $C$ a normalising constant. Note that the
534: Poisson process on $M$ with constant rate $\beta$ has density $\beta^{\,
535: \text{card}(\xb)}$. The parameter $\gamma$ defines the interactive component of the
536: process: when $0<\gamma <1$, configurations with high values of $\lambda(\xb\oplus G)$
537: for a given $\text{card}(\xb)$ are favoured, and so the interaction is described as {\em
538: repulsive\/}, while $\gamma >1$ favours $\xb$ with low $\lambda(\xb\oplus G)$ and so
539: constitutes the {\em attractive\/} case.
540:
541: The expected number of points in $M$ is a complicated function of $\gamma$, $\beta$
542: and $\rho$. We chose $\rho=0.01$, $\gamma=\gamma_{1}^{10^4}$ with $\gamma_{1}\in [0.3,
543: 1.5]$, and adjusted $\beta$ so that the number of points produced in the simulations
544: of the AIPP was close to 2000. This is because the AIPP doesn't re-scale in a simple
545: way, so comparison with the Voronoi $v$-process, in particular in regard to the
546: $J$-function, requires the samples to have roughly the same number of points.
547:
548: \begin{figure}[h]
549: \centering
550: \includegraphics{AIprocGammR.eps}
551: \caption{\footnotesize
552: Plot of $\rstr$ vs $\gamma_{1}$ for the AIPP
553: as described above, with cubic regression line.
554: Standard deviations of $\rstr$ are close to $0.005$. }
555: \label{AIGR}
556: \end{figure}
557: \begin{figure}[h]
558: \centering
559: \includegraphics{AIJfns.eps}
560: \caption{\footnotesize
561: Left: Plots of smoothed $\ln \hat J(r) $ curves for the AIPP's
562: with various values of $\gamma_{1}$, showing a range of attractive/repulsive effects.
563: Right: The graph of $\ln \hat J(r)$ for the AIPP with $\gamma_1=1.5$ is compared to
564: that for the
565: scale-free Voronoi $v$-process with $\alpha=0.5$ (these two processes produce almost
566: identical $\rstr$ values and NN EPMF's).}
567: \label{AIJfns}
568: \end{figure}
569:
570: Figures \ref{AIGR}--\ref{AInn} show the values of our three statistics for the AIPP. We
571: observe a continuous range of degrees of clustering on either side of CSR (when $\ln
572: J(r)\equiv 1$). It is interesting to compare these with those derived from the Voronoi
573: $v$-process. In particular, we can match results which produce close values for
574: $\rstr$. We have done this for the cases of the AIPP with $\gamma_{1} = 1.5$ and
575: Voronoi $v$-process with $\alpha = 0.5$ (for both processes, $\rstr\approx 0.2$). The
576: right-hand plots in Figs.\,\ref{AIJfns}, \ref{AInn} show that there is a significant
577: difference in the $J$-function curves, but that the Voronoi NN EPMF's do not yield
578: significant difference.
579:
580: In summary, the $v$-processes and the AIPP's produce a range of clusterings depending on
581: a continuous parameter. The difference in the resulting point patterns can be detected
582: by the combination of our three statistics (a)--(c).
583:
584: %The authors are grateful to the referee whose valuable comments helped to improve the
585: %exposition of the paper.
586:
587:
588: \begin{figure}[!h]
589: \centering
590: \includegraphics{AInn.eps}
591: \caption{\footnotesize Left: Proportions of cells with given numbers of neighbours (given
592: at the right-hand end) for the AIPP's, plotted vs $\gamma_{1}$.
593: Right: Comparison of the NN EPMF's for the AIPP with $\gamma_{1} = 1.5$, the
594: Voronoi $v$-process with $\alpha = 0.5$, and CSR.
595: Standard errors are close to 0.005. }
596: \label{AInn}
597: \end{figure}
598:
599:
600:
601:
602:
603:
604: \begin{thebibliography}{1}
605: \bibitem{Amb02}
606: {\sc Ambler,\.G.\,K.} (2002) {\em Dominated Coupling From The Past and Some Extensions
607: of the Area-Interaction Process.} PhD thesis, Department of Mathematics,
608: University of Bristol, U.K.
609:
610: \bibitem{As87}
611: {\sc Asmussen,~S.} (1987) {\em Applied probability and queues.} John Wiley, Chichester.
612:
613: \bibitem{BV95}
614: {\sc Baddeley,\,A. and van Lieshout,\,M.} (1995) {\rm Area-interaction point
615: processes.} \em Annals of the Institute of Statistical Mathematics, \bf 47, \rm
616: 601--619.
617:
618: \bibitem{BV96}
619: {\sc Baddeley,\,A. and van Lieshout,\,M.} (1996) {\rm A nonparametric measure of
620: spatial interaction in point patterns.} \em Statistica Neerlandica, \bf 50, \rm
621: 344--361.
622:
623:
624: \bibitem{borode}
625: {\sc Borovkov,\,K.A. and Odell,\,D.} (2006) {\rm On spatial culling-immigration point
626: processes based on Voronoi cells.} (Submitted; avaliable as {\em arXiv.org\/} paper
627: math.PR/0610606.)
628:
629:
630: \bibitem{K98}
631: {\sc Kendall,\,W.S.} (1998) {\rm Perfect simulation for the area-interaction point
632: process.} (1979) In: L.\,Accardi and C.C.\,Heyde (Eds.), \em Probability Towards 2000,
633: \rm 218--234, Springer, NY.
634:
635: \bibitem{LE79}
636: {\sc Lenz,~R. } {\rm Redundancy as an Index of Change in Point Pattern Analysis.} \em
637: Geographical Analysis, \bf 11, \rm 374--388.
638:
639:
640: \bibitem{MeTw92}
641: {\sc Meyn,\,S.P. and Tweedie,\,R.L.} (1992) Stability of Markovian processes I:
642: Criteria for discrete-time chains.{\em Adv. Appl. Prob.} {\bf 24,} 542--574.
643:
644: \bibitem{OBS}
645: {\sc Okabe,~A., Boots,\,B., Sugihara,\,K., and Chiu,\,S. } (2000) {\em Spatial
646: Tessellations: Concepts and Applications of Voronoi Diagrams.} John Wiley, Chichester.
647:
648:
649: \end{thebibliography}
650:
651:
652: \end{document}
653:
654:
655: Harris ergodicity: see e.g. Definition on
656:
657: 1) p. 560 of
658:
659: Meyn, S.P. and Tweedie, R.L. (1992) Stability of Markovian processes I: Criteria for
660: discrete-time chains. Adv. Appl. Prob. 24, 542-574.
661:
662: 2) p.869 of
663:
664: Exponential Ergodicity in Markovian Queueing and Dam Models Pekka Tuominen, Richard L.
665: Tweedie Journal of Applied Probability, Vol. 16, No. 4 (Dec., 1979), pp. 867-880
666:
667:
668: Metastable: having or characterized by only a slight margin of stability
669: (Merriam--Webster Dictionary
670: