cs0609097/main.tex
1: %% Ketan, Francesco, Emilio
2: 
3: \documentclass[12pt,draftcls,onecolumn]{IEEEtran}
4: %\documentclass[10pt,final]{IEEEtran}
5: %\documentclass[12pt]{article}
6: %\documentclass[conference,10pt]{ieeeconf}
7: \IEEEoverridecommandlockouts \overrideIEEEmargins
8: % \documentclass[conference]{cssconf}
9: %\documentclass[12pt]{article}
10: \usepackage{amssymb,amsmath,color,psfrag}
11: \usepackage{graphicx}
12: \usepackage{algorithm,algorithmic,xspace}
13: 
14: \renewcommand{\natural}{{\mathbb{N}}}
15: \newcommand{\irr}{{\mathbb{I}}}
16: \newcommand{\real}{{\mathbb{R}}}
17: \newcommand{\subscr}[2]{{#1}_{\textup{#2}}}
18: \newcommand{\union}{\cup}
19: \newcommand{\intersection}{\ensuremath{\operatorname{\cap}}}
20: \newcommand{\bigintersection}{\ensuremath{\operatorname{\bigcap}}}
21: \newcommand{\intersect}{\ensuremath{\operatorname{\cap}}}
22: \newcommand{\margin}[1]{\marginpar{\tiny #1}}
23: \newcommand{\map}[3]{#1: #2 \rightarrow #3}
24: %\definecolor{Iblue}{rgb}{.1289,.1601,.4179}
25: \newcommand{\setdef}[2]{\{#1 \; | \; #2\}}
26: \newcommand{\Bigcball}[2]{\overline{B}\Big(#2,#1\Big)}
27: 
28: \newcommand{\todo}[1]{\par\noindent{\color{red}\raggedright\sc{#1}
29:     \par\marginpar{\Large \bf $\star$}}}
30: \newcommand{\DD}{\mathcal{D}}
31: \newcommand{\PP}{\mathcal{P}}
32: \newcommand{\umax}{\subscr{r}{ctr}}
33: \newcommand{\vmax}{\subscr{r}{vel}}
34: %\newcommand{\SOTSP}{\textsc{Second Order Traveling Salesperson Problem}\xspace}
35: \newcommand{\pointset}{P}
36: \newcommand{\vinit}{v_0}
37: \newcommand{\setpointsets}[1]{\PP_{#1}}
38: \newcommand{\domain}{\ensuremath{\mathcal{Q}}}
39: \newcommand{\ETSP}[1]{\operatorname{ETSP}(#1)}
40: \newcommand{\SOTSP}[1]{\operatorname{DITSP}(#1)}
41: \newcommand{\dist}[2]{\operatorname{d}(#1,#2)}
42: \newcommand{\diam}{\operatorname{diam}}
43: \newcommand{\norm}[1]{\|#1\|}
44: \newcommand{\maximize}{\text{maximize}}
45: \newcommand{\subj}{\text{subj. to}}
46: \newcommand{\E}{\operatorname{E}}
47: \newcommand{\card}{\operatorname{card}}
48: \newcommand{\Length}{\operatorname{Length}}
49: \newcommand{\Area}{\operatorname{Area}}
50: \newcommand{\Volume}{\operatorname{Volume}}
51: \newcommand{\TT}{\mathcal{T}}
52: \newcommand{\DTSP}[2]{\operatorname{DTSP}_{#2}(#1)}
53: \newcommand{\AltAlgo}{\textsc{Alternating Algorithm}\xspace}
54: \newcommand{\BeadTilingAlgo}{\textsc{Bead-Tiling Algorithm}\xspace}
55: \newcommand{\BTA}{\textsc{Bead Tiling Algorithm}\xspace}
56: \newcommand{\CFA}{\textsc{Cylinder Covering Algorithm}\xspace}
57: \newcommand{\RecBTA}{\textsc{RecBTA}\xspace}
58: \newcommand{\RecCCA}{\textsc{RecCCA}\xspace}
59: \newcommand{\shortBTA}{\textsc{BTA}\xspace}
60: \newcommand{\shortCCA}{\textsc{CCA}\xspace}
61: \newcommand{\RecBeadTilingAlgo}{\textsc{Recursive Bead-Tiling Algorithm}\xspace}
62: \newcommand{\RecCylFillingAlgo}{\textsc{Recursive Cylinder-Covering Algorithm}\xspace}
63: \newcommand{\LenRBTA}[2]{\operatorname{L_{\mathrm{RBTA}, #2}}(#1)}
64: \newcommand{\LenRCFA}[2]{\operatorname{L_{\mathrm{RCFA}, #2}}(#1)}
65: \newcommand{\LenRecBeadAlgo}[1]{\operatorname{\mathcal{T}_{\mathrm{RecBTA}}(#1)}}
66: \newcommand{\LenRecCylAlgo}[1]{\operatorname{\mathcal{T}_{\mathrm{RecCFA}}(#1)}}
67: \newcommand{\RHRecBTA}{\textsc{Receding Horizon Recursive Bead-Tiling Algorithm}\xspace}
68: \newcommand{\RHRecCFA}{\textsc{Receding Horizon Recursive Cylinder-Filling Algorithm}\xspace}
69: \newcommand{\tile}{\ensuremath{\mathcal{B}_{\rho}}}
70: \newcommand{\cyl}{\ensuremath{\mathcal{C}_{\rho}}}
71: \newcommand{\speed}{s}
72: \newcommand{\DoubIntTSP}{\text{DITSP}}
73: \newcommand{\width}{W}
74: \newcommand{\height}{H}
75: \newcommand{\depth}{D}
76: 
77: %% theorems (always good to have some)
78: \newtheorem{theorem}{Theorem}[section]
79: \newtheorem{proposition}[theorem]{Proposition}
80: \newtheorem{corollary}[theorem]{Corollary}
81: \newtheorem{definition}[theorem]{Definition}
82: \newtheorem{lemma}[theorem]{Lemma}
83: \newtheorem{remark}[theorem]{Remark}
84: \newtheorem{remarks}[theorem]{Remarks}
85: \newtheorem{example}[theorem]{Example}
86: \newtheorem{algo}[theorem]{Algorithm}
87: \newtheorem{problem}[theorem]{Problem}
88: 
89: % "box" symbols at end of proofs
90: \def\QEDclosed{\mbox{\rule[0pt]{1.3ex}{1.3ex}}} % for a filled box
91: % V1.6 some journals use an open box instead that will just fit around a closed one
92: \def\QEDopen{{\setlength{\fboxsep}{0pt}\setlength{\fboxrule}{0.2pt}\fbox{\rule[0pt]{0pt}{1.3ex}\rule[0pt]{1.3ex}{0pt}}}}
93: \def\QED{\QEDclosed} % default to closed
94: 
95: % Procend
96: \newcommand\oprocendsymbol{\hbox{$\square$}}
97: \newcommand\oprocend{\relax\ifmmode\else\unskip\hfill\fi\oprocendsymbol}
98: \def\eqoprocend{\tag*{$\bullet$}}
99: 
100: %% Enumerate environment
101: \renewcommand{\theenumi}{(\roman{enumi})}
102: \renewcommand{\labelenumi}{\theenumi}
103: 
104: %\renewcommand{\baselinestretch}{0.965}
105: %\renewcommand{\baselinestretch}{1.4}
106: % \renewcommand{\clearpage}{}
107: 
108: \begin{document}
109: 
110: \title{\LARGE \bf Traveling Salesperson Problems for a double integrator}
111: 
112: \author{Ketan Savla \and Francesco Bullo\thanks{Ketan Savla and Francesco
113:     Bullo are with the Center for Control, Dynamical Systems and Computation,
114:     University of California at Santa Barbara,
115:     \texttt{ketansavla@umail.ucsb.edu, bullo@engineering.ucsb.edu}} \and
116:  Emilio Frazzoli\thanks{Emilio Frazzoli is with the Aeronautics and Astronautics Department, Massachusetts Institute of Technology,  \texttt{frazzoli@mit.edu}}}
117:   
118: \maketitle
119: 
120: 
121: 
122: \begin{abstract}
123:   In this paper we propose some novel path planning strategies for a double
124:   integrator with bounded velocity and bounded control inputs. First, we
125:   study the following version of the Traveling Salesperson Problem (TSP):
126:   given a set of points in $\real^d$, find the fastest tour over the point
127:   set for a double integrator. We first give asymptotic bounds on the time
128:   taken to complete such a tour in the worst-case. Then, we study a
129:   stochastic version of the TSP for double integrator where the points are
130:   randomly sampled from a uniform distribution in a compact environment in
131:   $\real^2$ and $\real^3$. We propose novel algorithms that perform within
132:   a constant factor of the optimal strategy with high probability. Lastly,
133:   we study a dynamic TSP: given a stochastic process that generates
134:   targets, is there a policy which guarantees that the number of unvisited
135:   targets does not diverge over time? If such stable policies exist, what
136:   is the minimum wait for a target?  We propose novel stabilizing
137:   receding-horizon algorithms whose performances are within a constant
138:   factor from the optimum with high probability, in $\real^2$ as well as
139:   $\real^3$. We also argue that these algorithms give identical
140:   performances for a particular nonholonomic vehicle, Dubins vehicle.
141: \end{abstract}  
142: 
143: \section{Introduction}
144: %problem description
145: The Traveling Salesperson Problem (TSP) with its variations is one of the
146: most widely known combinatorial optimization problems. While extensively
147: studied in the literature, these problems continue to attract great
148: interest from a wide range of fields, including Operations Research,
149: Mathematics and Computer Science. The Euclidean TSP (ETSP)
150: \cite{JB-JH-JH:59,JMS:90} is formulated as follows: given a finite point
151: set $\pointset$ in $\real^d$ for $d \in \natural$, find the minimum-length
152: closed path through all points in $\pointset$.  It is quite natural to
153: formulate this problem in the context of other dynamical vehicles.  The
154: focus of this paper is the analysis of the TSP for a vehicle with double
155: integrator dynamics or simply a double integrator; we shall refer to it as
156: $\DoubIntTSP$. Specifically, $\DoubIntTSP$ will involve finding the
157: \emph{fastest} tour for a double integrator through a set of points.
158: 
159: %literature review
160: Exact algorithms, heuristics and polynomial-time constant factor
161: approximation algorithms are available for the Euclidean TSP,
162: see~\cite{DA-RB-VC-WC:98,SA:97,SL-BWK:73}. However, unlike most other
163: variations of the TSP, it is believed that the $\DoubIntTSP$ cannot be
164: formulated as a problem on a finite-dimensional graph, thus preventing the
165: use of well-established tools in combinatorial optimization. On the other
166: hand, it is reasonable to expect that exploiting the geometric structure of
167: feasible paths for a double integrator, one can gain insight into the
168: nature of the solution, and possibly provide polynomial-time approximation
169: algorithms.
170: 
171: %problem motivation 
172: The motivation to study the $\DoubIntTSP$ arises in robotics and
173: uninhabited aerial vehicles (UAVs) applications. In particular, we envision
174: applying our algorithm to the setting of an UAV monitoring a collection of
175: spatially distributed points of interest. Additionally, from a purely
176: scientific viewpoint, it is of general interest to bring together the work
177: on dynamical vehicles and that on TSP. UAV applications also motivate us to
178: study the Dynamic Traveling Repairperson Problem (DTRP), in which the
179: aerial vehicle is required to visit a dynamically generated set of targets.
180: This problem was introduced by Bertsimas and van Ryzin in
181: \cite{DJS-GJvR:91} and then decentralized policies achieving the same
182: performances were proposed in \cite{EF-FB:03r}. Variants of these problems
183: have attracted much attention recently
184: \cite{EF-FB:03r,ZT-UO:05,RWB-TWM-MAG-EPA:02,SD:05,SR-RS-SD:05}. However, as
185: with the TSP, the study of DTRP in conjunction with vehicle dynamics has
186: eluded attention from the research community.
187: 
188: 
189: %contributions
190: The contributions of this paper are threefold.  First, we analyze the minimum time to traverse
191: $\DoubIntTSP$ in $\real^d$ for $d \in \natural$. We show that the minimum time to traverse
192: $\DoubIntTSP$ belongs to $O(n^{1-\frac{1}{2d}})$ and in the worst case,
193: it also belongs\footnote{For
194:   $\map{f,g}{\natural}{\real}$, we say that $f \in O(g)$ (respectively, $f
195:   \in \Omega(g)$) if there exist $N_0 \in \natural$ and $k \in \real_+$
196:   such that $|f(N)| \le k |g(N)|$ for all $N \ge N_0$ (respectively,
197:   $|f(N)| \ge k |g(N)|$ for all $N \ge N_0$).  If $f \in O(g)$ and $f \in
198:   \Omega(g)$, then we use the notation $f \in \Theta (g)$.} to $\Omega(n^{1-\frac{1}{d}})$.  Second, we
199: study the \emph{stochastic} $\DoubIntTSP$, i.e., the problem of finding the
200: fastest tour through a set of target points that are uniformly randomly
201: generated. We show that the minimum time to traverse the tour for the stochastic $\DoubIntTSP$ belongs to
202: $\Omega(n^{2/3})$ in $\real^2$ and $\Omega(n^4/5)$ in $\real^3$.  Drawing
203: inspiration from our earlier work~\cite{KS-EF-FB:06h-tmp}, we propose two
204: novel algorithms for the stochastic $\DoubIntTSP$: the \RecBeadTilingAlgo
205: for $\real^2$ and the \RecCylFillingAlgo for $\real^3$.  We prove that
206: these algorithms provide a constant-factor approximation to the optimal
207: $\DoubIntTSP$ solution with high probability.  Third, we propose two
208: algorithms for the DTRP in the heavy load case based on the
209: fixed-resolution versions of the corresponding algorithms for stochastic
210: $\DoubIntTSP$. We show that the performance guarantees for the stochastic
211: $\DoubIntTSP$ translate into stability guarantees for the average
212: performance of the DTRP problem for a double integrator.  Specifically, the
213: performances of the algorithms for the DTRP are within a constant factor of
214: the optimal policies.  We contend that the successful application to the
215: DTRP problem does indeed demonstrate the significance of the $\DoubIntTSP$
216: problem from a control viewpoint. As a final minor contribution, we also
217: show that the results obtained for stochastic $\DoubIntTSP$ carry over to
218: the stochastic TSP for the Dubins vehicle, i.e., for a nonholonomic vehicle
219: moving along paths with bounded curvature, without reversing direction.
220: 
221: This work completes the generalization of the known combinatorial results
222: on the ETSP and DTRP (applicable to systems with single integrator
223: dynamics) to double integrators and Dubins vehicle models.  It is
224: interesting to compare our results with the setting where the vehicle is
225: modeled by a single integrator; this setting corresponds to the so-called
226: Euclidean case in combinatorial optimization. The results are summarized as
227: follows: \smallskip
228: 
229: \begin{center}
230: \noindent\resizebox{.63\linewidth}{!}{
231: \begin{tabular}{|c|c|c|c|}
232: \hline
233: &Single& Double& Dubins\\
234: &integrator & integrator & vehicle \\
235: \hline \hline
236: Min. time for & $\Theta(n^{1-\frac{1}{d}})$ \cite{JMS:90} & $\Omega(n^{1-\frac{1}{d}})$, & $\Theta(n)$ \cite{KS-EF-FB:04l}\\
237: TSP tour & & $O(n^{1-\frac{1}{2d}})$& ($d=2,3$)\\
238: (worst-case) & & &\\
239: \hline
240: Exp. min. time & $\Theta(n^{1-\frac{1}{d}})$ \cite{JMS:90} & $\Theta(n^{1-\frac{1}{2d-1}})$ & $\Theta(n^{1-\frac{1}{2d-1}})$\\
241: for TSP tour & & w.h.p.& w.h.p. \\
242: (stochastic) & &($d=2,3$)&($d=2,3$)\\
243: \hline
244: System time & $\Theta(\lambda^{d-1})$ \cite{DJS-GJvR:91}& $\Theta(\lambda^{2(d-1)})$ & $\Theta(\lambda^{2(d-1)})$\\
245: for DTRP & ($d=1$)&($d=2,3$)&($d=2,3$) \\
246: \hline
247: \end{tabular}}
248: \end{center}
249: 
250: %% \margin{can something be said at all for all $d \in \natural$?} 
251: 
252: \smallskip Remarkably, the differences between these various bounds for the
253: TSP play a crucial role when studying the DTRP problem; e.g., stable
254: policies exist only when the minimum time for traversing the TSP tour grows
255: strictly sub-linearly with $n$.  For the DTRP problem we propose novel
256: policies and show their stability for a uniform target-generation process
257: with intensity $\lambda$.  It is clear from the table that motion
258: constraints make the system much more sensitive to increases in the target
259: generation rate $\lambda$.
260: 
261: \section{Setup and worst-case DITSP}
262: For $d \in \natural$, consider a vehicle with double integrator dynamics:
263: \begin{equation}
264:   \ddot{p}(t)=u(t),\quad \| u(t)\| \leq \umax, \quad \| \dot{p}(t)\| \leq \vmax, 
265:   \label{eq:second-order-model}
266: \end{equation}
267: where $p \in \real^d$ and $u \in \real^d$ are the position and control
268: input of the vehicle, $\vmax \in \real_+$ and $\umax \in \real_+$ are the
269: bounds on the attainable speed and control inputs.  Let
270: $\domain\subset\real^d$ be a unit hypercube.  Let
271: $\pointset=\{q_1,\ldots,q_n\}$ be a set of $n$ points in $\domain$ and
272: $\setpointsets{n}$ be the collection of all point sets $\pointset \subset
273: \domain$ with cardinality $n$.  Let $\ETSP{\pointset}$ denote the cost of
274: the Euclidean TSP over $\pointset$ and let $\SOTSP{\pointset}$ denote the
275: cost of the TSP for double integrator over $\pointset$, i.e., the time
276: taken to traverse the fastest closed path for a double integrator through
277: all points in $\pointset$. We assume $\vmax$ and $\umax$ to be constant and
278: we study the dependence of $\map{$\DoubIntTSP$}{\setpointsets{n}}{\real_+}$
279: on $n$.  Without loss of generality, we assume the vehicle starts
280: traversing the TSP tour at $t=0$ with initial position~$q_1$.
281: \begin{lemma}\textit{(Worst-case Lower Bound on the TSP for Double Integrator)}
282:   \label{lem:lower-bound-deterministic}
283:   For $\vmax,\umax\in\real_+$ and $d \in \natural$, there exists a point
284:   set $\pointset \in \setpointsets{n}$ in $\domain \subset \real^d$ such
285:   that $\SOTSP{\pointset}$ belongs to $\Omega(n^{1-\frac{1}{d}})$.
286: \end{lemma}
287: \begin{proof}
288:   We consider the class of point sets that give rise to the worst case
289:   scenario for the ETSP; we refer the reader to~\cite{JMS:90}.  It suffice
290:   to note that, for such a point set of cardinality $n$ in $\real^d$, the
291:   minimum distance between any two points belongs to
292:   $\Omega(n^{-\frac{1}{d}})$. The minimum time required for a double
293:   integrator with initial speed $\tilde{v}$ to go from one point to another
294:   at a distance $\tilde{\delta}$ is lower bounded by
295:   $\sqrt{(\tilde{v}/\umax)^2+2 (\tilde{\delta}/\umax)}-\tilde{v}/\umax$.
296:   However, $\tilde{v} \leq \vmax$ and for the point set under
297:   consideration, $\tilde{\delta}$ belongs to $\Omega(n^{-\frac{1}{d}})$.
298:   This implies that the minimum time required for a double integrator to
299:   travel between two points of the given point set belongs to
300:   $\Omega(n^{-\frac{1}{d}})$.  Hence, the minimum time required for the
301:   vehicle to complete the tour over this point set belongs to $n
302:   \Omega(n^{-\frac{1}{d}})$, i.e., $\Omega(n^{1-\frac{1}{d}})$.
303: \end{proof}
304:  
305: We now propose a simple strategy for the $\DoubIntTSP$ and analyze its
306: performance. The STOP-GO-STOP strategy can be described as follows: The
307: vehicle visits the points in the same order as in the optimal ETSP tour
308: over the same set of points.  Between any pair of points, the vehicle
309: starts at the initial point at rest, follows the shortest-time path to
310: reach the final point with zero velocity.
311: 
312: \begin{theorem}\textit{(Upper Bound on the TSP for Double Integrator)}
313:   \label{theorem:upper-bound-deterministic}
314:   For any point set $\pointset \in \setpointsets{n}$ in $\domain \subset
315:   \real^d$ and $\umax > 0$, $\vmax > 0$ and $d \in \natural$,
316:   $\SOTSP{\pointset}$ belongs to $O(n^{1-\frac{1}{2d}})$.
317: \end{theorem}
318: \begin{proof}
319:   Without any loss of generality, let
320:   $(q_1,\ldots,q_n,q_1)$ be the optimal order of points for the Euclidean
321:   TSP over $\pointset$.  For $1 \leq i \leq n-1$, let
322:   $\delta_i=\norm{q_i-q_{i+1}}$ and $\delta_n=\norm{q_n-q_1}$. If
323:   $\delta_i$ is the distance between a set of points, then the time $t_i$
324:   required to traverse that distance by a double integrator following the
325:   STOP-GO-STOP strategy is given by:
326:   \begin{equation*}
327:     t_i=  \begin{cases}
328:       2\sqrt{\frac{\delta_i}{\umax}}, & \quad\textrm{if $\delta_i \leq
329:       \frac{\vmax^2}{\umax}$},\\      
330:     \frac{\vmax}{\umax} + \frac{\delta_i}{\vmax}, & \quad
331:       \textrm{otherwise}. 
332:     \end{cases}
333:   \end{equation*}
334:   Let $\mathcal{I}=\setdef{1 \leq i \leq n}{\delta_i \leq \vmax^2/\umax}$
335:   and $\mathcal{I}^c=\{1,\ldots,n\} \setminus \mathcal{I}$. Also, let
336:   $n_{\mathcal{I}}$ be the cardinality of the set $\mathcal{I}$ and let
337:   $n_{\mathcal{I}^c}=n-n_{\mathcal{I}}$.  Therefore, an upper bound on the
338:   minimum time taken to complete the tour as obtained from this strategy is
339:   \begin{equation}\label{eq:first-expr-for-SOTSP}
340:     \begin{split}
341:       \SOTSP{\pointset} & \leq \sum_{i=1}^n t_i = \sum_{i \in \mathcal{I}} t_i + \sum_{i \in \mathcal{I}^c} t_i  
342:       =\frac{2}{\sqrt{\umax}} \sum_{i \in \mathcal{I}} \sqrt{\delta_i} +
343:       n_{\mathcal{I}^c} \frac{\vmax}{\umax} + \frac{1}{\vmax} \sum_{i \in
344:       \mathcal{I}^c} \delta_i  \\ 
345:     &\leq \frac{2}{\sqrt{\umax}} \sum_{i \in \mathcal{I}} \sqrt{\delta_i} +
346:       n_{\mathcal{I}^c} \Big(\frac{\vmax}{\umax} +
347:       \frac{\diam(\domain)}{\vmax} \Big),  
348:     \end{split}
349:   \end{equation}
350:   where $\diam(\domain)$ is the length of the largest segment lying
351:   completely inside $\domain$. From the well known upper bound
352:   \cite{JMS:90} on the tour length of optimal ETSP, there exists a constant
353:   $\beta(\domain)$ such that $\sum_{i \in \mathcal{I}} \delta_i \leq
354:   \sum_{i=1}^n \delta_i \leq \beta(\domain) n^{1-\frac{1}{d}}$. Hence an
355:   upper bound on the term $\sum_{i \in \mathcal{I}} \sqrt{\delta_i}$ in
356:   eqn. \eqref{eq:first-expr-for-SOTSP} can be obtained by solving the
357:   following optimization problem:
358:   \begin{equation*}
359:     \begin{split}
360:       \maximize \quad & \sum_{i \in \mathcal{I}} \sqrt{\delta_i}, \qquad
361:       \subj \quad 
362: \sum_{i \in \mathcal{I}} \delta_i \leq \beta(\domain)
363:       n^{1-\frac{1}{d}}.
364:     \end{split}
365:   \end{equation*}
366: By employing the method of Lagrange multipliers, one can see that the
367: maximum is achieved when $\delta_i = \beta(\domain)
368: \frac{n^{1-\frac{1}{d}}}{n_{\mathcal{I}}} \quad \forall i \in \mathcal{I}$.
369: Hence $\sum_{i \in \mathcal{I}} \sqrt{\delta_i} \leq \sqrt{\beta(\domain)}
370: \sqrt{n_{\mathcal{I}} n^{1-\frac{1}{d}}}$. Substituting this in eqn.
371: \eqref{eq:first-expr-for-SOTSP}, we get that
372: \begin{equation}
373:   \label{eq:final-expr-for-SOTSP}
374:   \SOTSP{\pointset} \leq \frac{2 \sqrt{\beta(\domain)}}{\sqrt{\umax}}
375:   \sqrt{n_{\mathcal{I}}} n^{\frac{1}{2}-\frac{1}{2d}} + n_{\mathcal{I}^c}
376:   \Big(\frac{\vmax}{\umax} + \frac{\diam(\domain)}{\vmax} \Big).
377: \end{equation}
378: However, $n_{\mathcal{I}} \leq n$ and Lemma~\ref{lem:num-of-dist-pts-bound}
379: implies that $n_{\mathcal{I}^c}$ belongs to $O(n^{1-\frac{1}{d}})$.
380: Incorporating these facts into eqn. \eqref{eq:final-expr-for-SOTSP}, one
381: arrives at the final result.
382: \end{proof}
383: 
384: The above theorem relies on the following key result.
385: \begin{lemma}
386: \label{lem:num-of-dist-pts-bound}
387: Given any point set $\pointset \in \setpointsets{n}$ in $\domain \subset
388: \real^d$, if $(q_1,q_2,\ldots,q_n,q_1)$ is the order of points for the
389: optimal ETSP tour over $\pointset$, then for any $\eta \in \real_+$, the
390: cardinality of the set $\setdef{q_i \in \pointset}{\norm{q_i-q_{i+1}} >
391:   \eta}$ belongs to $O(n^{1-\frac{1}{d}})$.
392: \end{lemma}
393: \begin{proof}
394:   By contradiction, assume there exists $\tilde{\eta} \in \real_+$ such
395:   that the cardinality of $\setdef{p_i \in \pointset}{\norm{q_i-q_{i+1}} >
396:     \tilde{\eta}}$ belongs to $\Omega(n^{1-\frac{1}{d}+\epsilon})$ for some
397:   $\epsilon > 0$. This implies that $\ETSP{P}$ belongs to $\tilde{\eta}
398:   \times \Omega(n^{1-\frac{1}{d}+\epsilon}) =
399:   \Omega(n^{1-\frac{1}{d}+\epsilon})$.  However, we know from~\cite{JMS:90}
400:   that $\ETSP{P}\in O(n^{1-\frac{1}{d}})$.
401: \end{proof}
402: 
403: \section{The stochastic $\DoubIntTSP$}
404: \label{sec:stochastic-SOTSP}
405: The results in the previous section showed that based on a simple strategy, the STOP-GO-STOP strategy, we are already guaranteed to have sublinear cost for the $\DoubIntTSP$ when the point sets are considered on an individual basis. However, it is reasonable to argue that
406: there might be better algorithms when one is concerned with \emph{average} performance. In particular, one can expect that when
407: $n$ target points are stochastically generated in $\domain$ according to a
408: uniform probability distribution function, the cost of $\DoubIntTSP$ should be lower than the one given by the STOP-GO-STOP strategy. We shall refer to the problem of studying the average performance of $\DoubIntTSP$ over this class of point sets as \emph{stochastic} $\DoubIntTSP$. In this section, we present novel algorithms for stochastic $\DoubIntTSP$ and then establish bounds on their performances.
409: 
410: We make the following assumptions: in $\real^2$, $\domain$ is a rectangle of width
411: $W$ and height $H$ with $W\geq H$; in $\real^3$, $\domain$ is a rectangular box of
412: width $W$, height $H$ and depth $\depth$ with $W\geq H\geq \depth$.
413: Different choices for the shape of $\domain$ affect our conclusions only by
414: a constant.  The axes of the reference frame are parallel to the sides of
415: $\domain$.  The points $P = (p_1,\ldots,p_n)$ are randomly generated
416: according to a uniform distribution in $\domain$.
417: 
418: \subsection{Lower bounds}
419: First we provide lower bounds on the expected length of stochastic $\DoubIntTSP$ for the 2 and 3 dimensional case.
420: \begin{theorem} \textit{(Lower bounds on stochastic $\DoubIntTSP$)}
421: \label{theorem:lower-bounds}
422: For all $\vmax>0$ and $\umax>0$, the expected cost of a stochastic $\DoubIntTSP$ visiting a set of $n$ uniformly-randomly-generated points satisfies the following inequalities:
423:   \begin{eqnarray*}
424:     \lim_{n\to+\infty} \frac{\E[\SOTSP{\pointset \subset \domain \subset \real^2}]}{n^{2/3}} \ge
425:     \frac{3}{4} \Big(\frac{6 \width \height}{\vmax \umax}\Big)^{1/3} \quad \text{and} \\
426:     \lim_{n\to+\infty} \frac{\E[\SOTSP{\pointset \subset \domain \subset \real^3}]}{n^{4/5}} \ge
427:     \frac{5}{6} \Big(\frac{20 \width \height \depth}{\pi \vmax \umax^2}\Big)^{1/5}.
428:   \end{eqnarray*}
429: \end{theorem} 
430: 
431: \begin{proof}
432: We first prove the first inequality.
433: Choose a random point $q_i \in \pointset$ as the initial position and $v_i$ as the initial speed of the
434:   vehicle on the tour, and choose the heading randomly. We would like
435:   to compute a bound on the expected time to the closest next point in
436:   the tour; let us call such a time $t^*$.
437: To this purpose, consider the set $R_\mathrm{t}$ of points that are
438:   reachable by a second order vehicle within time $t$ .
439:   It can be verified that the area of such a set can be bounded, as
440:   $t\to0^+$, by
441:   \begin{equation}
442: %    \label{eq:area}
443:     \mathrm{Area}(R_t) \leq \frac{\umax v_i t^3}{6} + o(t^3) \leq \frac{\umax \vmax t^3}{6} + o(t^3).
444:   \end{equation}
445: Given time $t$, the probability that $t^* > t$
446:   is no less than the probability that there is no other target
447:   reachable within a time at most $t$; in other words,
448:   \begin{equation*}
449:     \mathrm{Pr}[t^* > t] \ge 1 - n \frac{\mathrm{Area}(R_t)}{\mathrm{Area}(\domain)}
450:     \ge 1 - n \frac{\umax \vmax t^3}{6 \width \height} - o (t^3).    
451:   \end{equation*}
452:  In terms of expectation, defining $c = \frac{n \umax \vmax}{6 \width \height}$,
453:   \begin{eqnarray*}
454:     \E[t^*] &=& \int_0^{+\infty} \mathrm{Pr}(t^*> \xi) \;
455:     d\xi\\ &\ge & \int_0^{+\infty} \max\left\{0, 1-\frac{n \umax \vmax}{6 \width \height}\xi^3 - o
456:     (\xi^3)\right\}\; d\xi\\ & \ge & \int_0^{c^{-1/3}} (1- c\xi^3) \; d \xi
457:     - n \int_0^{c^{-1/3}} o(\xi^3) \; d \xi\\ 
458:     &=& \frac{3}{4}
459:     \left(\frac{6 \width \height}{\vmax \umax n}\right)^{1/3} - o (n^{-1/3}).
460:   \end{eqnarray*}
461: The expected total tour time will be no smaller than $n$ times the
462:   expected shortest time between two points, i.e.,
463:   $$\E[\SOTSP{\pointset}{\vmax,\umax}{2}] \geq \frac{3}{4} \left(\frac{6 n^2 \width \height}{\vmax \umax}\right)^{1/3} - o (n^{2/3}).$$
464:   Dividing both sides by $n^{2/3}$ and
465:   taking the limit as $n\to+\infty$, we get the first result.
466: 
467: We now prove the second inequality.
468: Choose a random point $q_i \in \pointset$ as the initial position and $v_i$ as the initial speed of the
469:   vehicle on the tour, and choose the heading randomly. We would like
470:   to compute a bound on the expected time to the closest next point in
471:   the tour; let us call such a time $t^*$.
472: To this purpose, consider the set $R_\mathrm{t}$ of points that are
473:   reachable by a second order vehicle within time $t$ .
474:   It can be verified that the volume of such a set can be bounded, as
475:   $t\to0^+$, by
476:   \begin{equation}
477:  %   \label{eq:volume}
478:     \mathrm{Volume}(R_t) \leq \frac{\pi \umax^2 v_i t^5}{20} + o(t^5) \leq  \frac{\pi \umax^2 \vmax t^5}{20} + o(t^5).
479:   \end{equation}
480: Given time $t$, the probability that $t^* > t$
481:   is no less than the probability that there is no other target
482:   reachable within a time at most $t$; in other words,
483:   \begin{equation*}
484:     \mathrm{Pr}[t^* > t] \ge 1 - n \frac{\mathrm{Volume}(R_t)}{\mathrm{Volume}(\domain)}
485:     \ge 1 - n \frac{\pi \umax^2 \vmax t^5}{20 \width \height \depth} - o (t^5).    
486:   \end{equation*}
487: In terms of expectation, defining $c = \frac{n \pi \umax^2 \vmax}{20 \width \height \depth}$,
488:   \begin{eqnarray*}
489:     \E[t^*] &=& \int_0^{+\infty} \mathrm{Pr}(t^*> \xi) \;
490:     d\xi\\ &\ge & \int_0^{+\infty} \max\left\{0, 1-\frac{n \pi \umax^2 \vmax}{20 \width \height \depth} \xi^5 - o
491:     (\xi^5)\right\}\; d\xi\\ & \ge & \int_0^{c^{-1/5}} (1- c\xi^5) \; d \xi
492:     - n \int_0^{c^{-1/5}} o(\xi^5) \; d \xi\\ 
493:     &=& \frac{5}{6}
494:     \left(\frac{20 \width \height \depth}{\vmax \umax^2 n}\right)^{1/5} - o (n^{-1/5}).
495:   \end{eqnarray*}
496: 
497: 
498: The expected total tour time will be no smaller than $n$ times the
499:   expected shortest time between two points, i.e.,
500:   $$\E[\SOTSP{\pointset}{\vmax,\umax}{3}] \geq \frac{5}{6} \left(\frac{20 n^4 \width \height \depth}{\vmax \umax^2}\right)^{1/5} - o (n^{4/5}).$$
501:   Dividing both sides by $n^{4/5}$ and
502:   taking the limit as $n\to+\infty$, we get the second result.
503: 
504: \end{proof}
505: 
506: 
507: 
508: \subsection{Constructive upper bounds}
509: In this section, we first recall the \RecBeadTilingAlgo from our earlier work \cite{KS-EF-FB:06a} on \emph{Dubins} vehicle and use it to propose novel algorithms for the stochastic $\DoubIntTSP$: the \RecBeadTilingAlgo for $\real^2$ and \RecCylFillingAlgo for $\real^3$. The performances of these algorithms will be shown to be within a constant factor of the optimal with high probability. 
510: 
511: In \cite{KS-EF-FB:06a}, we studied stochastic versions of TSP for Dubins vehicle. Though conventionally Dubins vehicle is restricted to be a \emph{planar} vehicle, one can easily generalize the  model even for the three (and higher) dimensional case. Correspondingly, Dubins vehicle can be defined as a vehicle that is constrained to move with a constant speed along paths of bounded curvature, without reversing direction. Accordingly, a \emph{feasible curve for Dubins vehicle} or a \emph{Dubins path} is defined as a curve that is twice differentiable almost everywhere, and such that the magnitude of its curvature is bounded above
512: by $1/\rho$, where $\rho>0$ is the minimum turn radius. Based on this, one can immediately come up with the following analogy between feasible curves for Dubins vehicle and a double integrator.
513: 
514: \begin{lemma}\textit{(Trajectories of Dubins and double integrators)}
515:   \label{lem:feasible-curves}
516:   A feasible curve for Dubins vehicle with minimum turn radius $\rho>0$ is
517:   a feasible curve for a double integrator (modeled in
518:   equation~\eqref{eq:second-order-model}) moving with a constant speed
519:   $\sqrt{\rho \umax}$. Conversely, a feasible curve for a double integrator
520:   moving with a constant speed $s \leq \vmax$ is a feasible curve for Dubins
521:   vehicle with minimum turn radius $\frac{s^2}{\umax}$.
522: \end{lemma}\smallskip
523: 
524: In \cite{KS-EF-FB:06h-tmp}, we proposed a novel algorithm, the
525: \RecBeadTilingAlgo for the stochastic version of the Dubins TSP (DTSP) in
526: $\real^2$; we showed that this algorithm performed within a constant factor
527: of the optimal with high probability. In this paper, taking inspiration
528: from those ideas, we propose algorithms to compute feasible curves for a
529: double integrator moving with a constant speed.
530: %
531: %% As it shall become clear, this does not affect our results since we are
532: %% interested in quantifying the dependence of the algorithms on $n$. At this
533: %% time, we would like to point out the obvious fact 
534: Note that moving at the maximum speed $\vmax$ is not necessarily the best
535: strategy since it restricts the maneuvering capability of the vehicle.
536: Nonetheless, this strategy leads to efficient algorithms.
537: %
538: In what follows we assume that the double integrator is moving with some
539: constant speed $s \leq \vmax$. Next, we proceed towards devising strategies
540: which perform within a constant factor of the optimal for stochastic
541: $\DoubIntTSP$ in $\real^2$ as well as $\real^3$, both with high
542: probability.
543: 
544: \subsubsection{The basic geometric construction}
545: \label{sec:basic}
546: Here we define useful geometric objects and study their properties.  Given
547: the constant speed $s$ for the double integrator let
548: $\rho=\frac{s^2}{\umax}$; from Lemma~\ref{lem:feasible-curves} this
549: constant corresponds to the minimum turning radius of the \emph{analogous}
550: Dubins vehicle. Consider two points $p_-$ and $p_+$ on the plane, with
551: $\ell = \|p_+-p_-\|_2 \le 4\rho$, and construct the bead $\tile(\ell)$ as
552: detailed in Figure~\ref{fig:tile}.
553: 
554: \begin{figure}[htb]
555:   \centerline{\includegraphics[width=0.5\columnwidth]{tile_constr_rectangle}}
556:   \caption{Construction of the ``bead'' $\tile(\ell)$. The figure shows how the
557:     upper half of the boundary is constructed, the bottom half is
558:     symmetric.The figure shows the rectangle $efgh$ which is used to
559:     construct the "cylinder" $\cyl(\ell)$.}
560:   \label{fig:tile}
561: \end{figure}
562: 
563: Associated with the bead is also the rectangle $efgh$. Rotating this
564: rectangle about the line passing through $p_-$ and $p_+$ gives rise to a
565: cylinder $\cyl(\ell)$. The regions $\tile(\ell)$ and $\cyl(\ell)$ enjoy the
566: following asymptotic properties as $(l/\rho)\to{0^+}$:
567: \begin{enumerate}
568: \item [(P1)] The maximum ``thickness'' of $\tile(\ell)$ is equal to
569:   \begin{equation*}
570:     w(\ell) = 4 \rho
571:     \left(1-\sqrt{1-\frac{\ell^2}{16\rho^2}}\right)=\frac{\ell^2}{8\rho}+
572:     \rho \cdot o\left(\frac{\ell^3}{\rho^3}\right) .
573:   \end{equation*}
574: The radius of cross-section of $\cyl(\ell)$ is $w(\ell)/4$ and the length of $\cyl(\ell)$ is $\ell$.
575: 
576: \item [(P2)] The area of $\tile(\ell)$ is equal to
577:   \begin{equation*}
578:     \Area(\tile(\ell)) = \frac{\ell w(\ell)}{2} = \frac{\ell^3}{16\rho} +
579:     \rho^2 \cdot o\left(\frac{\ell^4}{\rho^4}\right).
580:   \end{equation*}
581:   The volume of $\cyl(\ell)$ is equal to
582:   \begin{equation*}
583:   \Volume[\cyl(\ell)] = \pi \Big(\frac{w(\ell)}{4}\Big)^2 \frac{\ell}{2} = \frac{\pi \ell^5}{2048\rho^2} +
584:   \rho^3 \cdot o\left(\frac{\ell^6}{\rho^6}\right).
585:   \end{equation*}
586:   
587: \item [(P3)] For any $p \in \tile$, there is at least one feasible curve
588:   $\gamma_p$ through the points $\{p_-, p, p_+\}$, entirely contained
589:   within $\tile$.  The length of any such path is at most
590:   \begin{equation*}
591:     \Length(\gamma_p) \le 4\rho \arcsin\left(\frac{\ell}{4\rho}\right)
592:     = \ell + \rho \cdot o\left(\frac{\ell^3}{\rho^3}\right).
593:   \end{equation*}
594:   Analogously, for any $\tilde{p} \in \cyl$, there is at least one feasible
595:   curve $\gamma_{\tilde{p}}$ through the points $\{p_-, \tilde{p}, p_+\}$,
596:   entirely contained within the region obtained by rotating $\tile(\ell)$
597:   about the line passing through $p_-$ and $p_+$. The length of
598:   $\gamma_{\tilde{p}}$ satisfies the same upper bound as the one
599:   established for $\gamma_p$.
600: 
601: \end{enumerate}
602: The geometric shapes introduced above can be used to cover $\real^2$ and
603: $\real^3$ in an \emph{organized} way. The plane can be periodically
604: \emph{tiled}\footnote{A tiling of the plane is a collection of sets whose
605:   intersection has measure zero and whose union covers the plane.} by
606: identical copies of $\tile(\ell)$, for any $\ell\in]0,4\rho]$.  The
607: cylinder, however does not enjoy any such special property.  For our
608: purpose, we consider a particular covering of $\real^3$ by cylinders
609: described as follows.
610: 
611: \begin{figure}[htbp]
612:   \centerline{\includegraphics[width=0.5\linewidth]{LayerofCylinders}}
613:   \caption{A typical layer of cylinders formed by stacking rows of cylinders}
614:   \label{fig:layer-of-cylinders}
615: \end{figure}
616: 
617: A \emph{row of cylinders} is formed by joining cylinders end to end along
618: their length. A layer of cylinders is formed by placing rows of cylinders
619: parallel and on top of each other as shown in
620: Figure~\ref{fig:layer-of-cylinders}. For covering $\real^3$, these layers
621: are arranged next to each other and with offsets as shown in
622: Figure~\ref{fig:cross-section}(a), where the cross section of this
623: arrangement is shown.  We refer to this construction as the
624: \emph{covering of $\real^3$}. 
625: \begin{figure}[htbp]
626: \begin{center}
627:   \includegraphics[width=0.42\linewidth]{CrossSection}
628:   \includegraphics[width=0.42\linewidth]{PhaseTransition}\\
629:   (a) \phantom{a big big b blank space} (b)
630: \end{center}
631: \caption{(a): Cross section of the arrangement of the layers of cylinders
632:   used for covering $\domain \subset \real^3$, (b): The relative position
633:   of the bigger cylinder relative to smaller ones of the prior phase during
634:   the phase transition.}
635: \label{fig:cross-section}  
636: \end{figure}
637: 
638: %% Having defined the appropriate geometric objects, we now move towards
639: %% designing constant factor approximation algorithms for stochastic
640: %% $\DoubIntTSP$ in $\real^2$ and $\real^3$.
641: 
642: % We now proceed towards describing the details of the algorithms.
643: 
644: \subsubsection{The 2D case: The \RecBeadTilingAlgo (\RecBTA)}
645: Consider a tiling of the plane such that $\mathrm{Area}[\tile(\ell)] =
646: \mathrm{Area}[\domain \subset \real^2]/(2n) = WH/(2n)$; to obtain this
647: equality we assume $\ell$ to be a decreasing function of $n$ such that
648: $\ell(n) \leq 4\rho$. Furthermore, we assume the tiling is chosen to be
649: aligned with the sides of $\domain \subset \real^2$, see
650: Figure~\ref{fig:sweep4}.
651:  
652: The proposed algorithm consists of a sequence of phases; during each of
653: these phases, a feasible curve will be constructed that ``sweeps'' the set
654: $\domain$.  In the first phase, a feasible curve is constructed with the
655: following properties:
656:  \begin{enumerate}
657:  \item it visits all non-empty beads once,
658:  \item it visits all rows\footnote{A row is a maximal string of beads with
659:      non-empty intersection with~$\domain$.} in sequence top-to-down,
660:    alternating between left-to-right and right-to-left passes, and visiting
661:    all non-empty beads in a row,
662:  \item when visiting a non-empty bead, it services at least one target in
663:    it.
664: \end{enumerate}
665: %It is a consequence of bead's property (P3) that there exists a Dubins path
666: %visiting at least one target in any non-empty bead. 
667: 
668: In order to visit the outstanding targets, a new phase is initiated. In
669: this phase, instead of considering single beads, we will consider
670: ``meta-beads'' composed of two beads each, as shown in
671: Figure~\ref{fig:sweep4}, and proceed in a similar way as the first phase,
672: i.e., a feasible curve is constructed with the following properties:
673:  \begin{enumerate}
674:   \item the curve visits all non-empty meta-beads once,
675:   \item it visits all (meta-bead) rows in sequence top-to-down,
676:     alternating between left-to-right and right-to-left passes, and
677:     visiting all non-empty meta-beads in a row,
678:   \item when visiting a non-empty meta-bead, it services at least one target in
679:     it.
680: \end{enumerate}
681: 
682: This process is iterated at most $\log_2 n+1$ times, and at each phase
683: meta-beads composed of two neighboring meta-beads from the previous phase
684: are considered; in other words, the meta-beads at the $i$-th phase are
685: composed of $2^{i-1}$ neighboring beads. After the last phase, the leftover
686: targets will be visited using, for example, a greedy strategy.
687:  
688: \begin{figure*}[htbp]
689:   \centerline{\includegraphics[width=0.95\linewidth]{MetabeadPhases}}
690:   \caption{Sketch of ``meta-beads" at successive phases in the recursive bead tiling algorithm.}
691:   \label{fig:sweep4}
692: \end{figure*}
693: 
694: The following result is related to a similar result in
695: \cite{YA-AZB-ARK-EU:99}.
696: 
697: \begin{theorem}[Targets remaining after recursive phases]
698:   \label{theorem:RecBeadTilingAlgo}
699:   Let $\pointset\in\setpointsets{n}$ be uniformly randomly generated in
700:   $\domain \in \real^2$. The number of unvisited targets after the last
701:   recursive phase of the \RecBTA is less than $24 \log_2 n$ with high
702:   probability, i.e., with probability approaching one as $n\to+\infty$.
703: \end{theorem}\smallskip
704: 
705: \begin{proof}
706:   Associate a unique identifier to each bead, let $b(t)$ be the identifier
707:   of the bead in which the $t^{\text{th}}$ target is sampled, and let $h(t)
708:   \in \natural$ be the phase at which the $t^{\text{th}}$ target is
709:   visited.  Without loss of generality, assume that targets within a single
710:   bead are visited in the same order in which they are generated, i.e., if
711:   $b(t_1) = b(t_2)$ and $t_1 < t_2$, then $h(t_1) < h(t_2)$.  Let $v_i(t)$
712:   be the number of beads that contain unvisited targets at the inception of
713:   the $i^{\text{th}}$ phase, computed after the insertion of the
714:   $t^{\text{th}}$ target. Furthermore, let $m_i$ be the number of
715:   $i^{\text{th}}$ phase meta-beads (i.e., meta-beads containing $2^{i-1}$
716:   neighboring beads) with a non-empty intersection with $\domain$. Clearly,
717:   $v_i(t) \le v_i(n)$, $m_i \le 2m_{i+1}$, and $v_1(n) \le n \le m_1/2$
718:   with certainty.  The $t^{\text{th}}$ target will not be visited during
719:   the first phase if it is sampled in a bead that already contains other
720:   targets. In other words,
721:   \begin{equation*}
722:     \Pr\big[h(t) \ge 2|\; v_1(t)\big] = \frac{v_1(t)}{m_1} \le \frac{v_1(n)}{2n}
723:     \le \frac{1}{2}.     
724:   \end{equation*}
725:   Similarly, the $t^{\text{th}}$ target will not be visited during the
726:   $i^{\text{th}}$ phase if (i) it has not been visited before the
727:   $i^{\text{th}}$ pass, and (ii) it belongs to a meta-bead that already
728:   contains other targets not visited before the $i^{\text{th}}$ phase:
729:   \begin{align*}
730:     \Pr\big[h(t)  \ge i+1 &|\; (v_i(t-1), v_{i-1}(t-1), v_1(t-1))\big] \\
731:     &= \Pr\big[h(t) \ge i+1 |\;  h(t) \ge i, v_i(t-1)\big]  
732:      \; \cdot\; \Pr\big[h(t) \ge i|\; (v_{i-1}(t-1), \ldots, v_1(t-1))\big] \\
733:     &\le  \frac{v_i(t-1)}{m_i} \Pr[h(t) \ge i |\; (v_{i-1}(t-1), \ldots,
734:     v_1(t-1))] \\
735:     &= \prod_{j=1}^i \frac{v_j(t-1)}{m_j} \le \prod_{j=1}^i \frac{2^{j-1}
736:       v_j(n)}{2n} = \left(\frac{2^{\frac{i-3}{2}}}{n}\right)^i \prod_{j=1}^i
737:     v_j(n).
738:   \end{align*}
739:   
740:   Given a sequence $\{\beta_i\}_{i\in\natural}\subset\real_+$ and given a
741:   fixed $i \ge 1$, define a sequence of binary random variables
742:   \begin{equation*}
743:     Y_t =
744:     \begin{cases}
745:       1,\quad & \text{if}\;  h(t) \ge i+1 \mbox{ and } v_i(t-1) \le \beta_i n,\\
746:       0, &\text{otherwise.}
747:     \end{cases}
748:   \end{equation*}
749:   In other words, $Y_t=1$ if the $t^{\text{th}}$ target is not visited
750:   during the first $i$ phases even though the number of beads still
751:   containing unvisited targets at the inception of the $i^{\text{th}}$
752:   phase is less than $\beta_i n$.  Even though the random variable $Y_t$
753:   depends on the targets generated before the $t^{\text{th}}$ target, the
754:   probability that it takes the value 1 is bounded by
755:   \begin{equation*}
756:     \Pr[Y_t = 1 |\; b(1), b(2), \ldots, b(t-1)] \le 2^{\frac{i(i-3)}{2}}
757:     \prod_{j=1}^i \beta_j =: q_i,
758:   \end{equation*}
759:   regardless of the actual values of $b(1), \ldots, b(t-1)$.  It is
760:   known~\cite{YA-AZB-ARK-EU:99} that if the random variables $Y_t$ satisfy
761:   such a condition, the sum $\sum_t Y_t$ is stochastically dominated by a
762:   binomially distributed random variable, namely,
763:   \begin{equation*}
764:     \Pr\left[\sum_{t=1}^n Y_t  > k\right] \le \Pr[B(n,q_i) > k].
765:   \end{equation*}
766:   In particular,
767:   \begin{equation}
768:     \label{eq:chernoff}
769:     \Pr\left[\sum_{t=1}^n Y_t > 2nq_i\right] \le \Pr[B(n,q_i) > 2 np_i] <
770:     2^{-nq_i/3}, 
771:   \end{equation}
772:   where the last inequality follows from Chernoff's Bound~\cite{RM-PR:95}.
773:   Now, it is convenient to define $\{\beta_i\}_{i\in\natural}$ by
774:   \begin{equation*}
775:     \beta_1 = 1,\quad
776:     \beta_{i+1} =  2 q_i = 2^{\frac{i(i-3)}{2}+1}
777:     \prod_{j=1}^i \beta_j = 2^{i-2} \; \beta_i^2,
778:   \end{equation*}
779:   which leads to $\beta_i = 2^{1-i}$. In turn, this implies that
780:   equation~\eqref{eq:chernoff} can be rewritten as
781:   \begin{equation*}
782:     \Pr\left[\sum_{t=1}^n Y_t > \beta_{i+1}n\right]  
783:     < 2^{-\beta_{i+1}n/6}=2^{-\frac{n}{3 \cdot  2^{i}}}, 
784:   \end{equation*}
785:   which is less than $1/n^2$ for $i \le i^*(n) := \lfloor \log_2 n - \log_2
786:   \log_2 n - \log_2 6 \rfloor \le \log_2 n$.  Note that $\beta_{i} \le 12
787:   \; \frac{\log_2 n}{n}$, for all $i > i^*(n)$.
788:   
789:   Let $\mathcal{E}_i$ be the event that $v_i(n) \le \beta_i n$.  Note that
790:   if $\mathcal{E}_i$ is true, then $v_{i+1}(n) \le \sum_{t=1}^{n} Y_t$: the
791:   right hand side represents the number of targets that will be visited
792:   after the $i^{\text{th}}$ phase, whereas the left hand side counts the
793:   number of beads containing such targets. We have, for all $i \le i^*(n)$:
794:   \begin{equation*}
795:     \Pr\Big[ v_{i+1} > \beta_{i+1}n |\; \mathcal{E}_i \Big] \cdot
796:     \Pr[\mathcal{E}_i]  \le \Pr\left[\sum_{t=1}^n Y_t > \beta_{i+1} n\right]
797:     \le \frac{1}{n^2},
798:   \end{equation*}
799:   that is, $ \displaystyle \Pr\left[\neg \mathcal{E}_{i+1} |\;
800:     \mathcal{E}_i \right] \le \frac{1}{n^2\; \Pr[\mathcal{E}_i]},$
801:   and thus (recall that $\mathcal{E}_1$ is true with certainty):
802:   \begin{equation*}
803:     \Pr\left[\neg \mathcal{E}_{i+1}\right]  \le \frac{1}{n^2} + \Pr[\neg
804:     \mathcal{E}_i] \le \frac{i}{n^2}.
805:   \end{equation*}
806:   In other words, for all $i \le i^*(n)$, $v_{i}(n) \le \beta_in$ with high
807:   probability.
808:   
809:   Let us now turn our attention to the phases such that $i>i^*(n)$. The
810:   total number of targets visited after the $(i^*)^{\text{th}}$ phase is
811:   dominated by a binomial variable $B(n, 12 \log_2 n/n)$; in particular,
812:   \begin{align*}
813:     \Pr \Big[ v_{i^*+1} > 24 \log_2 n |\; \mathcal{E}_{i^*}\Big] \cdot
814:     \Pr[\mathcal{E}_{i^*}] &\le \Pr \Big[ \sum_{t=1}^n Y_t > 24 \log_2 n\Big] \\
815:     &\le \Pr \big[B(n, 12 \log_2 n/n) > 24 \log_2 n\big] \le 2^{-12 \log_2
816:       n}.
817:   \end{align*}
818:   Dealing with conditioning as before, we obtain
819:   \begin{equation*}
820:     \Pr \left[v_{i^*+1} > 24 \log_2 n\right] \le \frac{1}{n^{12}} + \Pr[\neg
821:     \mathcal{E}_{i^*}] \le \frac{1}{n^{12}} + \frac{\log_2 n}{n^2}.
822:   \end{equation*}
823:   In other words, the number of targets that are left unvisited after the
824:   $(i^*)^{\text{th}}$ phase is bounded by a logarithmic function of $n$
825:   with high probability.
826: \end{proof}
827: 
828: In summary, Theorem~\ref{theorem:RecBeadTilingAlgo} says that after a
829: sufficiently large number of phases, almost all targets will be visited,
830: with high probability.  The second key point is to recognize that (i) the
831: length of the first phase is of order $n^{2/3}$ and (ii) the length of each
832: phase is decreasing at such a rate that the sum of the lengths of the
833: $\lceil\log_2 n\rceil$ recursive phases remains bounded and proportional to
834: the length of the first phase.  (Since we are considering the asymptotic
835: case in which the number of targets is very large, the length of the beads
836: will be very small; in the remainder of this section we will tacitly
837: consider the asymptotic behavior as $\ell/\rho \to 0^+$.)
838: 
839: \begin{lemma}[Path length for the first phase]
840:   \label{lemma:first_phase} 
841:   Consider a tiling of the plane with beads of length $\ell$. For any $\rho
842:   > 0$ and for any set of target points, the length $L_1$ of a path
843:   visiting once and only once each bead with a non-empty intersection with
844:   a rectangle $\domain$ of width $W$ and length $H$ satisfies
845:   \begin{equation*}
846:     L_1 \le \frac{16 \rho WH}{\ell^2} \left( 1+ \frac{7}{3} \pi \frac{\rho}{W}
847:     \right)+ \rho \cdot o\left( \frac{\rho}{\ell}\right).
848:   \end{equation*}
849: \end{lemma}\smallskip
850: \begin{proof}
851:   A path visiting each bead once can be constructed by a sequence of
852:   passes, during which all beads in a row are visited in a left-to-right or
853:   right-to-left order. In each row, there are at most $\lceil W/\ell\rceil
854:   +1$ beads with a non-empty intersection with $\domain$. Hence, the cost
855:   of each pass is at most:
856:   \begin{equation*}
857:   L^\mathrm{pass}_1 \le W + 2 \ell + \rho \cdot
858:   o\left(\frac{\ell^2}{\rho^2}\right).
859:   \end{equation*} 
860:   
861:   Two passes are connected by a U-turn maneuver, in which the direction of
862:   travel is reversed, and the path moves to the next row, at distance equal
863:   to one half the width of a bead. The length of the shortest path to
864:   reverse the heading of the vehicle with co-located initial and final
865:   points is $(7/3) \pi \rho$, the length of the U-turn satisfies
866:   \begin{equation*}
867:     L^\mathrm{U-turn}_1 \le \frac{7}{3}\pi \rho + \frac{1}{2}w(\ell) \le
868:     \frac{7}{3} \pi \rho + \frac{\ell^2}{16 \rho} + \rho \cdot o\left(
869:       \frac{\ell^3}{\rho^3}\right).    
870:   \end{equation*}  
871:   The total number of passes, i.e., the total number of rows of beads with
872:   non-empty intersection with $\domain$, satisfies 
873:   \begin{equation*}
874:     N^\mathrm{pass}_1 \le \left\lceil \frac{2H}{w(\ell)} \right\rceil + 1
875:     \le \frac{16\rho H}{\ell^2} + 2 +
876:     o\left(\frac{\rho}{\ell}\right).
877:   \end{equation*}  
878:   A simple upper bound on the cost of closing the tour is given by 
879:   \begin{equation*}
880:     L^\mathrm{close}_1 \le (W + 2 \ell) + (H + 2 w(\ell)) + 2\pi \rho = W +
881:     H + 2\pi\rho + 2 \ell + \rho \cdot o(\ell/\rho).
882:   \end{equation*}  
883:   In summary, the total length of the path followed during the first phase
884:   is
885:   \begin{align*}
886:     L_1 & \le N^\mathrm{pass}_1 \left(L^\mathrm{pass}_1 +
887:       L^\mathrm{U-turn}_1\right) + L^\mathrm{close}
888:     \\
889:     &\le \left(\frac{16\rho H}{\ell^2} + 2 +
890:       o\left(\frac{\rho}{\ell}\right)\right) \left( W + 2 \ell +
891:       \frac{7}{3} \pi \rho + \frac{\ell^2}{16 \rho}+\rho \cdot o\left(
892:         \frac{\ell^2}{\rho^2}\right) \right) + W + H + 2 \pi \rho + 2 \ell
893:     + \rho \cdot o(\ell/\rho)
894:     \\
895:     &\le \frac{16 \rho WH}{\ell^2} \left( 1+ \frac{7}{3} \pi \frac{\rho}{W}
896:     \right)+ \rho \cdot o \left( \frac{\rho}{\ell}\right).
897: \end{align*}
898: \end{proof}
899: 
900: Based on the calculation for the first phase, we can estimate the length of the paths in
901: generic phases of the algorithm. Since the total number of phases in the
902: algorithm depends on the number of targets $n$, as does the length of the
903: beads $\ell$, we will retain explicitly the dependency on the phase number.
904:  
905: \begin{lemma}[Path length at odd-numbered phases]
906:   \label{lemma:odd}
907:   Consider a tiling of the plane with beads of length $\ell$. For any $\rho
908:   > 0$ and for any set of target points, the length $L_{2j-1}$ of a path
909:   visiting once and only once each meta-bead with a non-empty intersection
910:   with a rectangle $\domain$ of width $W$ and length $H$ at phase number
911:   $(2j-1)$, $j\in\natural$ satisfies
912:  \begin{equation*}
913:     L_{2j-1} 
914:     %%   \leq 2^{5-j} \left(\frac{\rho W
915:     %%        H}{\ell^2} \left(1 + \frac{7}{3}\frac{\pi \rho}{W} \right) + \rho
916:     %%      \cdot o\left(\frac{\rho^2}{\ell^2} \right) \right).
917:       \leq 2^{5-j} \left[\frac{\rho W H}{\ell^2} \left(1 +
918:         \frac{7}{3}\frac{\pi \rho}{W} \right) + \rho \cdot
919:       o\left(\frac{\rho}{\ell} \right) \right] + 32 \frac{\rho H}{\ell} 
920:       + \rho \cdot o\left(\frac{\rho}{\ell} \right) 
921:       + 2^j \left[3 \ell+ \rho \cdot  o\left(\frac{\ell}{\rho} \right) \right]. 
922:   \end{equation*}
923: \end{lemma}\smallskip
924: \begin{proof}
925:   During odd-numbered phases, the number of beads in a meta-bead is a
926:   perfect square and the considerations made in the proof of
927:   Lemma~\ref{lemma:first_phase} can be readily adapted.  The length of each
928:   pass satisfies
929:   \begin{equation*}
930:     L^\mathrm{pass}_{2j-1} \le \left(W + 2^{j} \ell\right) \left[ 1 +
931:     o\left(\frac{\ell}{\rho}\right)\right].      
932:   \end{equation*}
933:   The length of each U-turn maneuver is bounded as
934:   \begin{equation*}
935:     L^\mathrm{U-turn}_{2j-1} \le \frac{7}{3}\pi \rho + 2^{j-2}w(\ell) \le
936:     \frac{7}{3}\pi \rho + 2^{j-2}\;\left[ \frac{\ell^2}{8\rho} +  \rho \cdot
937:     o\left(\frac{\ell^3}{\rho^3}\right)\right],
938:   \end{equation*}  
939:   from which 
940:   \begin{equation*}
941:     L^\mathrm{pass}_{2j-1} + L^\mathrm{U-turn}_{2j-1} = W+ \frac{7}{3}\pi
942:     \rho + o \left(\frac{\ell}{\rho}\right) + 2^j \left[ \ell + \rho \cdot o
943:       \left(\frac{\ell}{\rho}\right)\right]. 
944:   \end{equation*}
945:     
946:   The number of passes satisfies:
947:   \begin{equation*}
948:     N^\mathrm{pass}_{2j-1} \le 2^{5-j} \left[\frac{\rho
949:     H}{\ell^2}+o\left(\frac{\rho}{\ell}\right)\right] + 2.    
950:   \end{equation*}
951:   Finally, the cost of closing the tour is bounded by
952:   \begin{equation*}
953:     L^\mathrm{close}_{2j-1} \le W + H +  2\pi \rho + 2^{j} \left[\ell +
954:       \rho \cdot o(\ell/\rho)\right].
955:   \end{equation*}
956:   Therefore, a bound on the total length of the path is 
957:   \begin{multline*}
958:     L_{2j-1} = N^\mathrm{pass}_{2j-1} (
959:     L^\mathrm{pass}_{2j-1}+L^\mathrm{U-turn}_{2j-1}) +
960:     L^\mathrm{close}_{2j-1}\\
961:     \leq 2^{5-j} \left[\frac{\rho W H}{\ell^2} \left(1 +
962:         \frac{7}{3}\frac{\pi \rho}{W} \right) + \rho \cdot
963:       o\left(\frac{\rho}{\ell} \right) \right] + 32 \frac{\rho H}{\ell} 
964:       + \rho \cdot o\left(\frac{\rho}{\ell} \right) 
965:       + 2^j \left[3 \ell+ \rho \cdot  o\left(\frac{\ell}{\rho} \right) \right].
966: % \\    
967: %    \le \left( 2^{5-j} \left(\frac{\rho H}{\ell^2} +
968: %        o\left(\frac{\rho}{\ell}\right)\right)+2\right)\cdot \left(W+2^j
969: %      \ell + W\cdot o\left(\frac{\ell^2}{\rho^2}\right)+2^j \rho \cdot
970: %      o\left(\frac{\ell^3}{\rho^3}\right) +\frac{7}{3} \pi \rho + 2^{j-5}
971: %      \frac{\ell^2}{\rho} + 2^{j-2} \rho \cdot
972: %      o\left(\frac{\ell^3}{\rho^3}\right) \right)
973: %    \\
974: %    + W+H+2\pi \rho + 2^j \left( \ell + \rho \cdot o(\ell/\rho)\right),
975: %    \\
976:   \end{multline*}
977: \end{proof}
978: 
979: \begin{lemma}[Path length at even-numbered phases]
980:   \label{lemma:even} 
981:   Consider a tiling of the plane with beads of length $\ell$. For any $\rho
982:   > 0$, a rectangle $\domain$ of width $W$ and length $H$ and any set of
983:   target points, paths in each phase of the $\BeadTilingAlgo$ can be chosen
984:   such that $L_{2j} \le 2 L_{2j+1}$, for all $j \in \natural$.
985: \end{lemma}\smallskip
986: \begin{proof}
987:   Consider a generic meta-bead $B_{2j+1}$ traversed in the
988:   $(2j+1)^\mathrm{th}$ phase, and let $l_3$ be the length of the path
989:   segment within $B_{2j+1}$. The same meta-bead is traversed at most twice
990:   during the $(2j)^\mathrm{th}$ phase; let $l_1$, $l_2$ be the lengths of
991:   the two path segments of the $(2j)^\mathrm{th}$ phase within $B_{2j+1}$.
992:   By convention, for $i\in\{1,2,3\}$, we let $l_i=0$ if the $i^\mathrm{th}$
993:   path does not intersect $B_{2j+1}$. Without loss of generality, the order
994:   of target points can be chosen in such a way that $l_1 \le l_2 \le l_3$,
995:   and hence $l_1+l_2 \le 2 l_3$. Repeating the same argument for all
996:   non-empty meta-beads, we prove the claim.
997: \end{proof}
998: 
999: Finally, we can summarize these intermediate bounds into the main result of
1000: this section. We let $\LenRBTA{\pointset}{\rho}$ denote the length of the
1001: path computed by the \RecBeadTilingAlgo for a point set~$\pointset$.
1002: \begin{theorem}[Path length for the \RecBeadTilingAlgo]
1003:   \label{theorem:total-path-upper-bound}
1004:   Let $P\in\setpointsets{n}$ be uniformly randomly generated in the
1005:   rectangle of width $W$ and height $H$. For any $\rho>0$, with high
1006:   probability
1007:   \begin{equation*}
1008:     \lim_{n\to+\infty} \frac{\DTSP{P}{\rho}}{n^{2/3}} \,\le\,
1009:     \lim_{n\to+\infty} \frac{ \LenRBTA{\pointset}{\rho} }{n^{2/3}} 
1010:     \, \le \, 24 \sqrt[3]{\rho W H} \left( 1+ \frac{7}{3} \pi \frac{\rho}{W}
1011:     \right).    
1012:   \end{equation*}
1013: \end{theorem}\smallskip
1014: \begin{proof}
1015:   For simplicity we let $\LenRBTA{\pointset}{\rho}=L_\mathrm{RBTA}$.
1016:   Clearly, $L_\mathrm{RBTA} = L'_\mathrm{RBTA} + L''_\mathrm{RBTA}$, where
1017:   $L'_\mathrm{RBTA}$ is the path length of the first $\lceil\log_2n\rceil$
1018:   phases of the $\RecBeadTilingAlgo$ and $L''_\mathrm{BTA}$ is the length
1019:   of the path required to visit all remaining targets.  An immediate
1020:   consequence of Lemma~\ref{lemma:even}, is that
1021:   \begin{equation*}
1022:     L'_\mathrm{RBTA} = \sum_{i=1}^{\lceil\log_2(n)\rceil} L_i
1023:     %%
1024:     \leq 3\sum_{j=1}^{\left\lceil{\log_2(n)}/{2}\right\rceil}
1025:     L_{2j-1}.
1026:   \end{equation*}
1027:   The summation on the right hand side of this equation can be expanded
1028:   using Lemma~\ref{lemma:odd}, yielding
1029:   \begin{multline*}
1030:     L'_\mathrm{RBTA} \le 3 \left\{ \left[\frac{\rho W H}{\ell^2} \left(1 +
1031:       \frac{7}{3}\frac{\pi \rho}{W} \right) + \rho \cdot
1032:       o\left(\frac{\rho^2}{\ell^2} \right) \right] 
1033:       \sum_{j=1}^{\left\lceil\log_2(n)/2\right\rceil}2^{5-j} \right.\\ \left. 
1034:       + \left( 32 \frac{\rho H}{\ell} + \rho \cdot
1035:       o\left(\frac{\rho}{\ell}\right) \right) \left\lceil \frac{\log_2n}{2}
1036:       \right\rceil + 
1037:       \left[ 3 \ell + \rho \cdot
1038:         o(\ell/\rho)\right]\sum_{j=1}^{\left\lceil\log_2(n)/2\right\rceil}2^j\right\}.
1039:   \end{multline*}
1040:   Since $\sum_{j=1}^k 2^{-j} \le \sum_{j=1}^{+\infty} 2^{-j} = 1,$ and
1041:   $\sum_{j=1}^k 2^j = 2^{k+1}-2 \le 2^{k+1},$ the previous equation can be
1042:   simplified to
1043:   \begin{multline*}
1044:     L'_\mathrm{RBTA} \le 3 \left\{ 32 \left[\frac{\rho W H}{\ell^2} \left(1
1045:           + \frac{7}{3}\frac{\pi \rho}{W} \right) + \rho \cdot
1046:         o\left(\frac{\rho}{\ell} \right) \right] \right.
1047:     \\
1048:     \left.  + \left(32 \frac{\rho H}{\ell} + \rho \cdot o\left(
1049:           \frac{\ell}{\rho} \right) \right) \left\lceil \frac{\log_2n}{2}
1050:       \right\rceil + \left[3 \ell + \rho \cdot o(\ell/\rho)\right]\cdot (4
1051:       \sqrt{n}) \right\}.
1052:   \end{multline*}  
1053:   Recalling that $\ell = 2(\rho WH/n)^{1/3} + o(n^{-1/3})$ for large $n$,
1054:   the above can be rewritten as
1055:   \begin{equation*}
1056:     L'_\mathrm{RBTA} \le 24 \sqrt[3]{\rho W H n^2}
1057:     \left( 1+ \frac{7}{3} \pi \frac{\rho}{W} \right)+ o(n^{2/3}).
1058:   \end{equation*}
1059:   Now it suffices to show that $L''_\mathrm{RBTA}$ is negligible with
1060:   respect to $L'_\mathrm{RBTA}$ for large $n$ with high probability.  From
1061:   Theorem~\ref{theorem:RecBeadTilingAlgo}, we know that with high
1062:   probability there will be at most $24 \log_2 n$ unvisited targets after
1063:   the $\lceil\log_2n\rceil$ recursive phases. From
1064:   \cite{KS-EF-FB:04l} we know that, with high probability, the
1065:   length of a \AltAlgo tour through these points satisfies
1066:   \begin{equation*}
1067:     L''_\mathrm{RBTA} \le \kappa \lceil 12 \log_2 n \rceil \pi \rho+ o(\log_2
1068:     n).   
1069:   \end{equation*}
1070: \end{proof}
1071: 
1072: In order to obtain an upper bound on the $\SOTSP{\pointset}$, we derive the
1073: expression for time taken, $\mathcal{T}_\mathrm{RecBTA}$, by the \RecBTA to
1074: execute the path of length $\LenRBTA{\pointset}{\rho}$ and then optimize it
1075: with respect to $\rho$. Based on this calculation, we get the following
1076: result.
1077: 
1078: \begin{theorem}\textit{(Upper bound on the total time in $\real^2$)}
1079:   \label{theorem:total-time-upper-bound-2D}
1080:   Let $P\in\setpointsets{n}$ be uniformly randomly generated in the
1081:   rectangle of width $W$ and height $H$. For any double integrator
1082:   \eqref{eq:second-order-model}, with high probability
1083:   \begin{equation*}
1084:     \lim_{n\to+\infty} \frac{\mathcal{T}_\mathrm{RecBTA}}{n^{2/3}} 
1085:     \, \le \, 24 \left(\frac{WH}{\vmax \umax}\right)^{1/3} \left( 1+ \frac{7\pi\vmax^2}{3W}
1086:     \right). 
1087:   \end{equation*}
1088: \end{theorem}\smallskip
1089: 
1090: \begin{remark}
1091:   Theorems~\ref{theorem:lower-bounds}
1092:   and~\ref{theorem:total-time-upper-bound-2D} imply that, with high
1093:   probability, the \RecBTA is a
1094:   $\frac{32}{\sqrt[3]{6}}\left(1+\frac{7\pi\vmax^2}{3\umax W}
1095:   \right)$-factor approximation (with respect to $n$) to the optimal stochastic DITSP
1096:   in $\real^2$ and that $\E[\SOTSP{\pointset \subset \domain \subset \real^2}]$
1097:   belongs to $\Theta(n^{2/3})$.
1098: \end{remark}
1099: \subsubsection{The 3D case: The \RecCylFillingAlgo (\RecCCA)}
1100: Consider a covering of $\domain \in \real^3$ by cylinders such that
1101: $\mathrm{Volume}[\cyl(\ell)]= \mathrm{Volume}[\domain \subset \real^3]/(4n)
1102: = WH\depth/(4n)$ (Again implying that $n$ is sufficiently large).
1103: Furthermore, the covering is chosen in such a way that it is aligned with
1104: the sides of $\domain \subset \real^3$.
1105: 
1106: The proposed algorithm will consist of a sequence of phases; each phase
1107: will consist of five sub-phases, all similar in nature. For the first
1108: sub-phase of the first phase, a feasible curve is constructed with the
1109: following properties:
1110:  \begin{enumerate}
1111:   \item it visits all non-empty cylinders once,
1112:   \item it visits all rows of cylinders in a layer in sequence top-to-down in a layer,
1113:     alternating between left-to-right and right-to-left passes, and
1114:     visiting all non-empty cylinders in a row,
1115:   \item it visits all layers in sequence from one end of the region to the other,
1116:   \item when visiting a non-empty cylinder, it services at least one target in
1117:     it.
1118: \end{enumerate}
1119: 
1120: \begin{figure*}[htbp]
1121:   \centerline{\includegraphics[width=0.95\linewidth]{AllSubPhases}}
1122:   \caption{Starting from top left in the left-to-right, top-to bottom
1123:     direction, sketch of projection of ``meta-cylinders'' on the
1124:     corresponding side of $\domain \subset \real^3$ at second, third,
1125:     fourth and fifth sub-phases of a phase in the recursive cylinder
1126:     covering algorithm.} 
1127:   \label{fig:allsubphases}
1128: \end{figure*}
1129: 
1130: In subsequent sub-phases, instead of considering single cylinders, we will
1131: consider ``meta-cylinders'' composed of $2$, $4$, $8$ and $16$ beads each
1132: for the remaining four sub-phases, as shown in
1133: Figure~\ref{fig:allsubphases}, and proceed in a similar way as the first
1134: sub-phase, i.e., a feasible curve is constructed with the following
1135: properties:
1136:  \begin{enumerate}
1137:   \item the curve visits all non-empty meta-cylinders once,
1138:   \item it visits all (meta-cylinder) rows in sequence top-to-down in a
1139:     (meta-cylinder) layer, alternating between left-to-right and
1140:     right-to-left passes, and visiting all non-empty meta-cylinders in a
1141:     row,
1142:   \item it visits all (meta-cylinder) layers in sequence from one end of
1143:     the region to the other,
1144:   \item when visiting a non-empty meta-cylinder, it services at least one
1145:     target in it.
1146: \end{enumerate}
1147: 
1148: A meta-cylinder at the end of the fifth sub-phase, and hence at the end of
1149: the first phase will consist of 16 nearby cylinders. After this phase, the
1150: transitioning to the next phase will involve enlarging the cylinder to $32$
1151: times its current size by increasing the radius of its cross section by a
1152: factor of $4$ and doubling its length as outlined in
1153: Figure~\ref{fig:cross-section}(b). It is easy to see that this bigger
1154: cylinder will contain the union of $32$ nearby smaller cylinders. In other
1155: words, we are forming the object $\cyl(2\ell)$ using a conglomeration of $32$
1156: $\cyl(\ell)$ objects. This whole process is repeated at most $\log_2{n}+2$ times. After the last phase, the leftover
1157: targets will be visited using, for example, a greedy strategy.
1158: 
1159: %% We have the following result, which is similar to the one for
1160: %% \RecBTA.
1161: 
1162: %% \begin{lemma} 
1163: %% \label{theorem:RecCylFillingAlgo}
1164: %% Let $\pointset\in\setpointsets{n}$ be uniformly randomly generated in
1165: %% $\domain \subset \real^3$. Then, the number of unvisited targets after the
1166: %% last phase of the \RecCCA over $\pointset$ belongs to $O(\log
1167: %% n)$ with high probability.
1168: %% \end{lemma}\smallskip
1169: 
1170: %% We know at this point is that after a sufficiently large number of phases,
1171: %% almost all targets will be visited, with high probability. The key point is
1172: %% to recognize that the length of each phase is decreasing at such a rate
1173: %% that the sum of the lengths of all the phases remains bounded.  We first
1174: %% state and prove the following result which characterizes the time required
1175: %% to execute the \RecCCA.
1176: 
1177: %% \begin{lemma}\textit{(Time for the \RecCCA)}
1178: %%   \label{lem:len-for-phases-3D}
1179: %%   Let $\pointset\in\setpointsets{n}$ be uniformly randomly generated in
1180: %%   $\domain \subset \real^3$. Then the time required by double integrator to
1181: %%   execute $\log_2 n$ phases of the \RecCCA over $\pointset$
1182: %%   belongs to $O(n^{4/5})$.
1183: %% \end{lemma}\smallskip
1184: 
1185: %% % \begin{proof}
1186: %% %   Due to space limitations, we present only a short sketch of the proof.
1187: %% %   One can verify that the time required to execute the $i$-th phase,
1188: %% %   $\mathcal{T}_i$ belongs to $O(\frac{1}{(2^{i-1}l)^4})$, i.e., there
1189: %% %   exists a constant $c$ such that $\mathcal{T}_i \leq
1190: %% %   \frac{c}{(2^{i-1}l)^4}$. However $l$ was chosen such that
1191: %% %   $\mathrm{Volume}[\cyl(l)]=1/(4n)$. This choice of $l$ implies that
1192: %% %   $\mathcal{T}_i \leq \frac{\tilde{c}n^{4/5}}{2^{4(i-1)}}$ for appropriate
1193: %% %   constant $\tilde{c}$.  The total length of the path is upper bounded by
1194: %% %   $\sum_{i=1}^{+\infty} \mathcal{T}_i$ which belongs to $O(n^{4/5})$.
1195: %% % \end{proof}
1196: 
1197: %% Based on the results obtained so far, we are now ready to state an upper
1198: %% bound on the time required by double integrator to service all the $n$
1199: %% targets while executing the \RecCCA followed by some greedy
1200: %% heuristics; let $\LenRecCylAlgo{\pointset}$ represent the corresponding
1201: %% quantity.
1202: 
1203: We have the following results, which are similar to the one for the \RecBeadTilingAlgo.
1204: 
1205: \begin{theorem}[Targets remaining after recursive phases]
1206:   \label{theorem:RecCylFillingAlgo}
1207:   Let $\pointset\in\setpointsets{n}$ be uniformly randomly generated in
1208:   $\domain \subset \real^3$. The number of unvisited targets after the last recursive phase
1209:   of the \RecCylFillingAlgo over $\pointset$ is less than $24 \log_2 n$
1210:   with high probability, i.e., with probability approaching one as
1211:   $n\to+\infty$.
1212: \end{theorem}\smallskip
1213: 
1214: \begin{lemma}[Path length for the first sub phase]
1215:   \label{lemma:first_sub_phase} 
1216:   Consider a covering of the space with cylinders $\cyl(\ell)$. For any $\rho
1217:   > 0$ and for any set of target points, the length $L_I$ of a path executing the first sub-phase of the \RecCylFillingAlgo in a rectangular box $\domain$ of width $\width$, height $\height$ and depth $\depth$ satisfies
1218:   \begin{equation*}
1219:     L_I \le \frac{1024 \rho^2 \width \height \depth}{\ell^4} \left( 1+ \frac{7}{3} \pi \frac{\rho}{\width}
1220:     \right)+ \rho \cdot o\left( \frac{\rho^3}{\ell^3}\right).
1221:   \end{equation*}
1222: \end{lemma}\smallskip
1223: \begin{proof}
1224:   A path visiting each cylinder once can be constructed by a sequence of
1225:   passes, during which all cylinders in a row are visited by making left-to-right and then 
1226:   right-to-left passes. This is done for all the rows of cylinders. In each row, there are at most $\lceil \width/\ell\rceil
1227:   +1$ cylinders encountered in one pass. Hence, the cost
1228:   of each pass is at most:
1229:   \begin{equation*}
1230:   L^\mathrm{pass}_I \le \width + 2 \ell + \rho \cdot
1231:   o\left(\frac{\ell^2}{\rho^2}\right).
1232:   \end{equation*} 
1233:   In order to visit all cylinders in a row, the vehicle needs to make two passes through that row and the paths for these two passes are connected by a u-turn path whose length is $\frac{7}{3}\pi \rho$ + $\frac{\ell}{2}$. Therefore the length of the path required to visit all cylinders in one row is:
1234:   \begin{equation*}
1235:     L^\mathrm{row}_I \le 2 \width + \frac{9}{2} \ell + \frac{7}{3}\pi \rho + \rho \cdot
1236:   o\left(\frac{\ell^2}{\rho^2}\right).    
1237: \end{equation*}
1238:   During the transition from one row to another, the vehicle needs to make a U-turn maneuver, in which the direction of
1239:   travel is reversed, and the path moves to the next row, at distance equal
1240:   to the diameter of the cylinder. Since the length of the shortest path to
1241:   reverse the heading of the vehicle with co-located initial and final
1242:   points is $(7/3) \pi \rho$, the length of the U-turn satisfies
1243:   \begin{equation}
1244: \label{eq:u-turn}
1245:     L^\mathrm{U-turn}_I \le \frac{7}{3}\pi \rho + \frac{1}{2}w(\ell) \le
1246:     \frac{7}{3} \pi \rho + \frac{\ell^2}{16 \rho} + \rho \cdot o\left(
1247:       \frac{\ell^3}{\rho^3}\right).    
1248:   \end{equation}  
1249:   The total number of rows, i.e., the total number of rows of cylinders with
1250:   non-empty intersection with $\domain$, satisfies
1251:   \begin{equation*}
1252:     N^\mathrm{row}_I \le \left\lceil \frac{2 \height}{w(\ell)} \right\rceil + 1
1253:     \le \frac{16\rho H}{\ell^2} + 
1254:     o\left(\frac{\rho}{\ell}\right).
1255:   \end{equation*}  
1256: During the transition from one row to another, the vehicle needs to make a U-turn maneuver whose length satisfies the same bound as in Eq. \eqref{eq:u-turn}
1257:   The total number of layers of cylinders satisfies
1258: \begin{equation*}
1259:       N^\mathrm{layer}_I \le \left\lceil \frac{4 \depth}{w(\ell)} \right\rceil + 1
1260:     \le \frac{32\rho \depth}{\ell^2} + 
1261:     o\left(\frac{\rho}{\ell}\right).
1262: \end{equation*}  
1263:   A simple upper bound on the cost of closing the tour is given by 
1264:   \begin{equation*}
1265:     L^\mathrm{close}_I \le (W + 2 \ell) + (H + 2 w(\ell)) + (\depth + w(\ell))+ 2\pi \rho = W +
1266:     H + \depth + 2\pi\rho + 2 \ell + \rho \cdot o(\ell/\rho).
1267:   \end{equation*}  
1268:   In summary, the total length of the path followed during the first sub-phase
1269:   is
1270:   \begin{align*}
1271:     L_I & \le N^\mathrm{layer}_I \left(N^\mathrm{row}_I \left(L^\mathrm{row}_I +
1272:       L^\mathrm{U-turn}_I\right) + L^\mathrm{U-turn}_I\right) + L^\mathrm{close}
1273:     \\
1274:     &\le \frac{1024 \rho^2 \width \height \depth}{\ell^4} \left( 1+ \frac{7}{3} \pi \frac{\rho}{\width}
1275:     \right)+ \rho \cdot o \left( \frac{\rho^3}{\ell^3}\right).
1276: \end{align*}
1277: \end{proof}
1278: 
1279: Based on this calculation, we can estimate the length of the paths in
1280: subsequent sub-phases.
1281:  
1282: \begin{eqnarray*}
1283: L_{II} \le \frac{1024 \rho^2 \width \height \depth}{\ell^4} \left( 1+ \frac{7}{3} \pi \frac{\rho}{\width}
1284:     \right)+ \rho \cdot o \left( \frac{\rho^3}{\ell^3}\right),\\
1285: L_{III} \le \frac{512 \rho^2 \width \height \depth}{\ell^4} \left( 1+ \frac{7}{3} \pi \frac{\rho}{\width}
1286:     \right)+ \rho \cdot o \left( \frac{\rho^3}{\ell^3}\right),\\
1287: L_{IV} \le \frac{512 \rho^2 \width \height \depth}{\ell^4} \left( 1+ \frac{7}{3} \pi \frac{\rho}{\width}
1288:     \right)+ \rho \cdot o \left( \frac{\rho^3}{\ell^3}\right),\\
1289: L_{V} \le \frac{256 \rho^2 \width \height \depth}{\ell^4} \left( 1+ \frac{7}{3} \pi \frac{\rho}{\width}
1290:     \right)+ \rho \cdot o \left( \frac{\rho^3}{\ell^3}\right).
1291: \end{eqnarray*}
1292: 
1293: The length of path to execute the first phase is then the some of the path lengths for these five sub-phases.
1294: 
1295: \begin{lemma}[Path length at the first phase]
1296:   \label{lemma:first-phase-3D}
1297:   Consider a covering of the space with cylinders $\cyl(\ell)$. For any $\rho
1298:   > 0$ and for any set of target points, the length $L_1$ of a path visiting once and only once each cylinder with a non-empty intersection with a rectangular box $\domain$ of width $\width$, height $\height$ and depth $\depth$ satisfies
1299:   \begin{equation*}
1300:     L_1 \le \frac{3328 \rho^2 \width \height \depth}{\ell^4} \left( 1+ \frac{7}{3} \pi \frac{\rho}{\width}
1301:     \right)+ \rho \cdot o\left( \frac{\rho^3}{\ell^3}\right).
1302:   \end{equation*}
1303: \end{lemma}\smallskip
1304:   
1305: Since we increase the length of cylinders by a factor of two while doing the phase transtion from one phase to the another, the length of path for the subsequent $i^{\text{th}}$ phase is given by:
1306: \begin{equation*}
1307:     L_i \le \frac{3328 \rho^2 \width \height \depth}{16^i\ell^4} \left( 1+ \frac{7}{3} \pi \frac{\rho}{\width}
1308:     \right)+ \rho \cdot o\left( \frac{\rho^3}{\ell^3}\right).
1309: \end{equation*}
1310: 
1311: Finally, we can summarize these intermediate bounds into the main result of
1312: this section. We let $\LenRCFA{\pointset}{\rho}$ denote the length of the
1313: path computed by the \RecCylFillingAlgo for a point set~$\pointset$.
1314: \begin{theorem}[Path length for the \RecCylFillingAlgo]
1315:   \label{theorem:total-path-upper-bound-3D}
1316:   Let $P\in\setpointsets{n}$ be uniformly randomly generated in the
1317:   rectangle of width $W$, height $H$ and depth $\depth$. For any $\rho>0$, with high
1318:   probability
1319:   \begin{equation*}
1320:     \lim_{n\to+\infty} \frac{\SOTSP{\pointset \subset \domain \subset \real^3}}{n^{4/5}} \,\le\,
1321:     \lim_{n\to+\infty} \frac{ \LenRCFA{\pointset}{\rho} }{n^{4/5}} 
1322:     \, \le \, \frac{3328}{15}\left(\frac{\pi}{16}\right)^{4/5} (\rho^2 W H \depth)^{1/5}.    
1323:   \end{equation*}
1324: \end{theorem}\smallskip
1325: 
1326: \begin{proof}
1327: Clearly, 
1328: \begin{align*}
1329: L_\mathrm{RCFA} & = \sum_{i=1}^{\lceil \frac{\log[2]{n}+7}{5} \rceil} \left(\frac{3328 \rho^2 \width \height \depth}{16^i\ell^4} \left( 1+ \frac{7}{3} \pi \frac{\rho}{\width}
1330:     \right)+ \rho \cdot o\left( \frac{\rho^3}{\ell^3}\right)\right) \\
1331: & \le \frac{53248 \rho^2 \width \height \depth}{15\ell^4} \left( 1+ \frac{7}{3} \pi \frac{\rho}{\width}
1332:     \right)+ \rho \cdot o\left( \frac{\rho^3}{\ell^3}\right).
1333: \end{align*}
1334: Recalling that $\ell = 2\left(\frac{16 \rho^2 WH \depth}{\pi n}\right)^{1/5} + o(n^{-1/5})$ for large $n$,
1335:   the above can be rewritten as
1336:   \begin{equation*}
1337:     L_\mathrm{RCFA} \le \frac{3328}{15}\left(\frac{\pi}{16}\right)^{4/5} (\rho^2 W H \depth)^{1/5} \left( 1+ \frac{7}{3} \pi \frac{\rho}{\width}
1338:     \right) n^{4/5} + o(n^{4/5}).
1339:  \end{equation*}
1340: \end{proof}
1341: 
1342: \begin{theorem}\textit{(Upper bound on the total time in $\real^3$)}
1343:   \label{theorem:total-time-upper-bound-3D}
1344:   Let $P\in\setpointsets{n}$ be uniformly randomly generated in the
1345:   rectangular box of width $W$, height $H$ and depth $\depth$. For any
1346:   double integrator \eqref{eq:second-order-model}, with high probability
1347:   \begin{equation*}
1348:     \lim_{n\to+\infty} \frac{\mathcal{T}_\mathrm{RecCCA}}{n^{4/5}} 
1349:     \, \le \, 61 \left(\frac{W H \depth}{\umax^2\vmax}\right)^{1/5} 
1350:     \left(1+ \frac{7 \pi \vmax^2}{3 W \umax} \right). 
1351:   \end{equation*}
1352: \end{theorem}\smallskip
1353: 
1354: \begin{remark}
1355:   Theorems~\ref{theorem:lower-bounds}
1356:   and~\ref{theorem:total-time-upper-bound-3D} imply that, with high
1357:   probability, the \RecCCA is a $50\left(1+\frac{7\pi\vmax^2}{3\umax W}
1358:   \right)$-factor approximation (with respect to $n$) to the optimal stochastic DITSP
1359:   in $\real^3$ and that $\E[\SOTSP{\pointset \subset \domain \subset \real^3}]$
1360:   belongs to $\Theta(n^{4/5})$.
1361: \end{remark}
1362: 
1363: \section{The DTRP for double integrator}
1364: \label{sec:DTRP}
1365: We now turn our attention to the Dynamic Traveling Repairperson Problem
1366: (DTRP) that was introduced in~\cite{DJS-GJvR:91} and that we here tackle
1367: for a double integrator.
1368: 
1369: \subsection{Model and problem statement}
1370: In the DTRP the double integrator is required to visit a dynamically
1371: growing set of targets, generated by some stochastic process.  We assume
1372: that the double integrator has unlimited range and target-servicing
1373: capacity and that it moves at a unit speed with minimum turning radius
1374: $\rho>0$.
1375: 
1376: Information about the outstanding targets representing the demand at time
1377: $t$ is described by a finite set $n(t)$ of positions $\DD(t)$. Targets are
1378: generated, and inserted into $\DD$, according to a time-invariant
1379: spatio-temporal Poisson process, with time intensity $\lambda > 0$, and
1380: uniform spatial density inside the region $\domain$, which we continue to
1381: assume to be a rectangle for two dimensions and a rectangular box for three
1382: dimensions.
1383: %% In other words, given a set $\mathcal{S} \subseteq\domain$, the expected
1384: %% number of targets generated in $\mathcal{S}$ within the time interval
1385: %% $[t,t']$ is
1386: %% \begin{equation*}
1387: %%   \E\big[\card(D(t') \cap \mathcal{S}) - \card(D(t) \cap \mathcal{S})\big] =
1388: %%   \lambda(t'-t) \Area(\mathcal{S}).
1389: %% \end{equation*}
1390: %% (Strictly speaking, the above equation holds when targets are not being
1391: %% removed from the queue $D$.) 
1392: Servicing of a target and its removal from the set $\DD$, is achieved when
1393: the double integrator moves to the target position.  A control policy $\Phi$
1394: for the DTRP assigns a control input to the vehicle as a function of its
1395: configuration and of the current outstanding targets.
1396: %% We also consider policies that compute a control input based on a snapshot
1397: %% of the outstanding target configurations at certain time sequences. Let
1398: %% $\TT_{\Phi}=\{t_k\}_{k\in\natural}$ be a strictly increasing sequence of
1399: %% times at which such computations are started: with some abuse of
1400: %% terminology, we will say that $\Phi$ is a receding horizon strategy if it
1401: %% is based on the most recent target data $D_{\text{rh}}(t)$, where
1402: %% \begin{equation*}
1403: %%   D_{\text{rh}}(t) = D(\max\setdef{t_{\text{rh}}\in
1404: %%   \TT_{\Phi}}{t_{\text{rh}}\leq t}).
1405: %% \end{equation*}
1406: The policy $\Phi$ is a stable policy for the DTRP if, under its action
1407: \begin{equation*}
1408:   n_{\Phi} = \lim_{t\to+\infty} \E[n(t)|\;\dot{p}=\Phi(p,\DD)] < +\infty, 
1409: \end{equation*}
1410: that is, if the double integrator is able to service targets at a rate that
1411: is, on average, at least as fast as the rate at which new targets are
1412: generated.
1413: 
1414: Let $T_j$ be the time elapsed from the time the $j^{\text{th}}$ target is
1415: generated to the time it is serviced and let $T_{\Phi}:=\lim_{j\to+\infty}
1416: \E[T_j]$ be the steady-state system time for the DTRP under the policy
1417: $\Phi$. (Note that if the system is stable, then it is known~\cite{LK:75}
1418: that $n_{\Phi}=\lambda T_{\Phi}$.)  Clearly, our objective is to design a
1419: policy $\Phi$ with minimal system time $T_\Phi$.
1420: 
1421: \subsection{Lower and constructive upper bounds}
1422: In what follows, we design control policies that provide constant-factor
1423: approximation of the optimal achievable performance. Consistently with the
1424: theme of the paper, we consider the case of \emph{heavy load}, i.e., the
1425: problem as the time intensity $\lambda\to+\infty$. We first provide lower bounds for the system time, and then present
1426: novel approximation algorithms providing upper bound on the performance.
1427: 
1428: \begin{theorem}[Lower bound on the DTRP system time]
1429:   \label{theorem:lower-bound-system-time}
1430:   For a double integrator \eqref{eq:second-order-model}, the system time $T_{\mathrm{DTRP},2}$ and
1431:   $T_{\mathrm{DTRP},3}$ for the DTRP in two and three dimensions satisfy
1432:   \begin{align*}
1433:     \lim_{\lambda\to\infty}\!\! \frac{T_{\mathrm{DTRP},2}}{\lambda^2}&
1434:     \ge\frac{81}{32} \frac{W\!H}{\vmax \umax},
1435:     \enspace
1436:     \lim_{\lambda\to\infty}\!\! \frac{T_{\mathrm{DTRP},3}}{\lambda^4}
1437:     \ge\frac{7813}{972} \frac{W\!H\!\depth}{\vmax \umax^2}.
1438:   \end{align*}
1439: \end{theorem}\smallskip
1440: 
1441: \begin{proof}
1442: We prove the lower bound on $T_{\mathrm{DTRP},2}$; the bound on $T_{\mathrm{DTRP},3}$ follows on similar lines.
1443: Let us assume that a stabilizing policy is available. In such a case, the number of outstanding targets approaches a finite steady-state value, $n^*$, related to the system time by Little's formula, i.e., $n^*=\lambda T_{\mathrm{DTRP},2}$. In order for the policy to be stabilizing, the time needed, on average, to service $m$ targets must be no greater than the average time interval in which $m$ new targets are generated. The average tim needed by the double integrator to service one target is no gretaer than the expected minimum time from an arbitrarily placed vehicle to the closest target; in other words, we can write the stability condition $\E[t^*(n^*)]\leq 1/\lambda$. A bound on the expected value of $t^*$ has been computed in the proof of Theorem~\ref{theorem:lower-bounds}, yielding
1444: \begin{equation*}
1445: \frac{3}{4}\left(\frac{6 W H}{\vmax \umax n}\right)^{1/3} \leq \E[t^*(n^*)]\leq 1/\lambda.
1446: \end{equation*}
1447: Using Little's formula $n^*=\lambda T_{\mathrm{DTRP},2}$, and rearranging, we get the desired result.
1448: \end{proof}
1449: 
1450: We now propose simple strategies, the \BTA (for $\real^2$) and the \CFA (for $\real^3$), based on the concepts
1451: introduced in the previous section.  The \BTA (\shortBTA) strategy consists of the following
1452: steps:
1453: \begin{enumerate}
1454: \item Tile the plane with beads of length $\ell := \min\{
1455:   C_\mathrm{BTA}/\lambda,4\rho\}$, where
1456:   \begin{equation} 
1457:     \label{eq:C_BTA}
1458:     C_\mathrm{BTA}  =
1459:     0.5241 \vmax  \left(1 + \frac{7\pi\rho}{3W}\right)^{-1}.
1460:   \end{equation}
1461: \item \label{step2} Traverse all non-empty beads once, visiting one target
1462:   per non-empty bead. Repeat this step.
1463: \end{enumerate}
1464: 
1465: The \CFA (\shortCCA) strategy is akin to the \shortBTA, where the region is covered with
1466: cylinders constructed from beads of length $\ell := \min\{
1467: C_\mathrm{CFA}/\lambda,4\rho\}$, where
1468: \begin{equation*}
1469:   %% \label{eq:C_CFA}
1470:   C_\mathrm{CCA}  =
1471:   0.1615 \vmax  \left(1 +\frac{7\pi\rho}{3W}\right)^{-1}.
1472: \end{equation*} 
1473: The policy is then to traverse all non-empty cylinders once, visiting one
1474: target per non-empty cylinder.  The following result characterizes the
1475: system time for the closed loop system induced by these algorithms and is
1476: based on the bounds derived to arrive at
1477: Theorems~\ref{theorem:total-time-upper-bound-2D} and
1478: \ref{theorem:total-time-upper-bound-3D}.
1479: 
1480: \begin{theorem}[Upper bound on the DTRP system time]
1481:   \label{theorem:DTRP}
1482:   For a double integrator \eqref{eq:second-order-model} and $\lambda>0$, the \shortBTA and the \shortCCA are stable policies for the
1483:   DTRP and the resulting system times $T_\mathrm{BTA}$ and $T_\mathrm{CFA}$ satisfy:
1484:   \begin{gather*}
1485:     \lim_{\lambda\to\infty}\! \frac{T_{\mathrm{DTRP},2}}{\lambda^2} \le
1486:     \lim_{\lambda \to\infty}\! \frac{T_\mathrm{BTA}}{\lambda^2} \le 70.5 \frac{WH}{\vmax \umax} \left(1+\frac{7\pi\vmax^2}{3W\umax} \right)^3,
1487:     \\
1488:     \lim_{\lambda\to\infty}\! \frac{T_{\mathrm{DTRP},3}}{\lambda^4} \le
1489:     \lim_{\lambda \to\infty}\! \frac{T_\mathrm{CFA}}{\lambda^4} \le 2\cdot 10^7
1490:     \frac{W\!H\!\depth}{\vmax \umax^2}\!\left(\!1\! + \frac{7\pi\rho}{3W}\right)^5\!\!\!.
1491:   \end{gather*}
1492:   \end{theorem}\smallskip
1493: 
1494: \begin{proof}
1495: We prove the upper bound on $T_{\mathrm{DTRP},2}$; the upper bound on $T_{\mathrm{DTRP},3}$ follows on similar lines.
1496:   Consider a generic bead $B$, with non-empty intersection with $\domain$.
1497:   Target points within $B$ will be generated according to a Poisson process
1498:   with rate $\lambda_B$ satisfying
1499:   \begin{equation*}
1500:     \lambda_B = \lambda \frac {\Area(B \cap \domain)}{WH}\le \lambda\frac{
1501:       \Area(B)}{WH} = \frac{C_\mathrm{BTA}^3}{16 \rho WH\lambda^2} +
1502:     o\left(\frac{1}{\lambda^2}\right).    
1503:   \end{equation*}
1504:   The vehicle will visit $B$ at least once every $T_{\RecBTA,1}$ time units, where
1505:   $T_{\RecBTA,1}$ is the bound on the time required to traverse a path of length $L_1$, as computed
1506:   in Lemma~\ref{lemma:first_phase}. As a consequence, targets in $B$ will
1507:   be visited at a rate no smaller than 
1508:   \begin{equation*}
1509:     \mu_B = \frac{C_\mathrm{BTA}^2 \vmax}{16 \rho WH \lambda^2}
1510:     \left(1+\frac{7}{3}\pi \frac{\rho}{W}\right)^{-1} + o
1511:     \left(\frac{1}{\lambda^2}\right).  
1512:   \end{equation*}
1513:   In summary, the expected time $T_\mathrm{B}$ between the appearance of a
1514:   target in $B$ and its servicing by the vehicle is no more than the system
1515:   time in a queue with Poisson arrivals at rate $\lambda_B$, and
1516:   deterministic service rate $\mu_B$. Such a queue is called a $M/D/1$
1517:   queue in the literature~\cite{LK:75}, and its system time is known to be
1518:   \begin{equation*}
1519:     T_{M/D/1} = \frac{1}{\mu_B} \left(1+\frac{1}{2}
1520:       \frac{\lambda_B}{\mu_B-\lambda_B}\right).  
1521:   \end{equation*}
1522:   Using the computed bounds on $\lambda_B$ and $\mu_B$, and taking the
1523:   limit as $\lambda\to+\infty$, we obtain
1524:   \begin{equation}
1525:     \label{eq:TB}
1526:     \lim_{\lambda\to+\infty} \frac{T_\mathrm{B}}{\lambda^2} \le
1527:     \lim_{\lambda\to+\infty} \frac{T_{M/D/1}}{\lambda^2} \le  
1528:     \frac{16 \rho W H}{C^2_\mathrm{BTA}\vmax \left(1+\frac{7}{3}\pi
1529:         \frac{\rho}{W}\right)^{-1}} \left(1 + \frac{1}{2}
1530:       \frac{C_\mathrm{BTA}}{\vmax \left(1+\frac{7}{3}\pi \frac{\rho}{W}\right)^{-1}-
1531:         C_\mathrm{BTA}}   \right). 
1532:   \end{equation}
1533:   Since equation~\eqref{eq:TB} holds for {\em any} bead intersecting
1534:   $\domain$, the bound derived for $T_B$ holds for all targets and is
1535:   therefore a bound on $T_\mathrm{DTRP,2}$. The expression on the right hand
1536:   side of~\eqref{eq:TB} is a constant that depends on problem parameters
1537:   $\rho$, $W$, and $H$, and on the design parameter $C_\mathrm{BTA}$, as
1538:   defined in equation~\eqref{eq:C_BTA}. Stability of the queue is
1539:   established by noting that $C_\mathrm{BTA} < \vmax (1+7/3\; \pi\;
1540:   \rho/W)^{-1}$.  Additionally, the choice of $C_\mathrm{BTA}$ in
1541:   equation~\eqref{eq:C_BTA} minimizes the right hand side of \eqref{eq:TB}
1542:   yielding the numerical bound in the statement. We then substitute $\rho=\vmax^2/\umax$ to yield the final result.
1543: \end{proof}
1544: 
1545: \begin{remark}
1546: Note that the achievable performances of the \shortBTA and the \shortCCA provide a
1547: constant-factor approximation to the lower bounds established
1548: in Theorem~\ref{theorem:lower-bound-system-time}.
1549: \end{remark}
1550: \section{Extension to the TSPs for the Dubins vehicle}
1551: In our earlier works \cite{KS-EF-FB:04l,KS-FB-EF:05j,KS-EF-FB:06h-tmp}, we
1552: have studied the TSP for the Dubins vehicle in the planar case. In
1553: \cite{KS-EF-FB:04l}, we proved that in the worst case, the time taken to
1554: complete a TSP tour by the Dubins vehicle will belong to $\Theta(n)$. One
1555: could shown that this result holds true even in $\real^3$. In
1556: \cite{KS-EF-FB:06h-tmp}, the first known algorithm with strictly sublinear
1557: asymptotic minimum time for tour traversal was proposed for the stochastic
1558: DTSP in $\real^2$. This algorithm was modified in \cite{KS-EF-FB:06h-tmp}
1559: to give a constant factor approximation to the optimal with high
1560: probability.  This naturally lead to a stable policy for the DTRP problem
1561: for the Dubins vehicle in $\real^2$ which also performed within a constant
1562: factor of the optimal with high probability. The $\RecCCA$ developed in
1563: this paper can naturally be extended to apply to the stochastic DTSP in
1564: $\real^3$. It follows directly from Lemma~\ref{lem:feasible-curves} that in
1565: order to use the $\RecCCA$ for a Dubins vehicle with minimum turning radius
1566: $\rho$, one has to simply compute feasible curves for double integrator
1567: moving with a constant speed $\sqrt{\rho \umax}$. Hence the results stated
1568: in Theorem~\ref{theorem:total-time-upper-bound-3D} and
1569: Theorem~\ref{theorem:DTRP} also hold true for the Dubins vehicle.
1570: 
1571: This equivalence between trajectories makes the \RecCCA the first known
1572: strategy with a strictly sublinear asymptotic minimum time for tour
1573: traversal for stochastic DTSP in $\real^3$. The fact that it performs
1574: within a constant factor of the optimal with high probability and that it
1575: gives rise to a constant factor approximation and stabilizing policy for
1576: DTRP for Dubins vehicle in $\real^3$ is also novel.
1577: 
1578: \section{Conclusions}
1579: \label{sec:conclusion}
1580: In this paper we have proposed novel algorithms for various TSP problems
1581: for vehicles with double integrator dynamics.  Future directions of
1582: research include extensive simulations to support the results obtained in
1583: this paper, study of centralized and decentralized versions of the DTRP,
1584: and more general task assignment and surveillance problems for vehicles
1585: with nonlinear dynamics.
1586: 
1587: %% \section*{Acknowledgment}
1588: %% This material is based upon work supported in part by ONR YIP Award
1589: %% N00014-03-1-0512 and AFOSR MURI Award F49620-02-1-0325. The authors thank
1590: %% the anonymous reviewers and editors for their helpful comments.
1591: 
1592: {\small
1593:   \bibliographystyle{ieeetr}%
1594:   \bibliography{alias,Main,FB,New}
1595: }
1596: 
1597: \end{document}
1598: 
1599: