cs0408003/path.tex
1: % 20-apr MANOR further typo fixing
2: % 7-Apr-2003 add l_p, add discussion about it. Took out the cube.
3: % 6-Apr-2003 Manor: Further typo fixing.
4: % 4-Apr-2003 HASVA to journal version. Removing the probab' path approx.
5: % 18-Dec-02. Small typo fix Manor
6: % Jul-2003. preparing for J. Algorithms
7: % Aug-7 2003 preparing for SICOMP & minor typos
8: % Feb 23 2004. Incorporating SICOMP referees' comments
9: % Jun 2004. Few stylistic improvements (MM)
10: 
11: 
12: \documentclass[11pt]{article}
13: 
14: \usepackage{amsmath,amssymb,amsthm,verbatim,fullpage}
15: \usepackage{epsfig,color}
16: 
17: 
18: 
19: \theoremstyle{plain}
20: \newtheorem{theorem}{Theorem}
21: \newtheorem{lemma}{Lemma}
22: \newtheorem{corollary}[lemma]{Corollary}
23: \newtheorem{remark}{Remark}
24: % \newtheorem{claim}[lemma]{Claim}
25: \newtheorem{proposition}[lemma]{Proposition}
26: \newtheorem{conjecture}{Conjecture}
27: \newtheorem{property}{Property}
28: \theoremstyle{definition}
29: \newtheorem{definition}{Definition}
30: % \newtheorem{algorithm}{Algorithm}
31: \newtheorem{observation}[lemma]{Observation}
32: 
33: % \newcommand{\qed}{\ensuremath{\Box}}
34: 
35: \newcommand{\E}{\mathbb{E}}
36: \newcommand{\la}{\langle}
37: \newcommand{\ra} {\rangle}
38: \newcommand{\eps}{\varepsilon}
39: 
40: %\newcommand{\algon}{\ensuremath{\text{\textsc{On}}}}
41: \newcommand{\algon}{\ensuremath{\mathcal{A}}}
42: \newcommand{\A}{\ensuremath{\mathcal{A}}}
43: \newcommand{\opt}{\ensuremath{\text{\textsc{Opt}}}}
44: \newcommand{\cost}{\text{cost}}
45: \newcommand{\bi}{\begin{itemize}}
46: \newcommand{\ei}{\end{itemize}}
47: \newcommand{\be}{\begin{enumerate}}
48: \newcommand{\ee}{\end{enumerate}}
49: % \newcommand{\etal}{\textsl{et.~al.}}
50: \newcommand{\ie}{\textsl{i.e.}}
51: \newcommand{\Ie}{\textsl{I.e.}}
52: \newcommand{\eg}{\textsl{e.g.}}
53: \newcommand{\dist}{\mathsf{d}}
54: \newcommand{\mS}{\mathcal{S}}
55: \newcommand{\poly}{\rm poly}
56: \newcommand{\f}{\mathsf{f}}
57: 
58: % \newcommand{\\Hst}{Ultrametric}
59: % \newcommand{\\Hsts}{Ultrametrics}
60: % \newcommand{\hst}{ultrametric}
61: % \newcommand{\\hsts}{ultrametrics}
62: % \newcommand{\ahst}{an ultrametric}
63: \newcommand{\mommit}[1]{}
64: \newcommand{\madd}[1]{#1}
65: \newcommand{\mnote}[1]{ $\ll$ \textsf{#1} --- Manor $\gg$ 
66: \marginpar{\textsc{mm}}}
67: \newcommand{\ynote}[1]{ $\ll$ \textsf{#1} --- Yair $\gg$ 
68: \marginpar{\textsc{yy}}}
69: 
70: \DeclareMathOperator{\lca}{lca}
71: 
72: % \newenvironment{proof}{\noindent\emph{Proof:}}{\hfill$\Box$}
73: 
74: 
75: 
76: \title{Multi-Embedding of Metric Spaces%
77: \thanks{A preliminary version of this paper appeared in \cite{BM03}.}}
78: 
79: 
80: \author{Yair Bartal\thanks{School of Computer Science, Hebrew University, 
81: Jerusalem 91904, Israel.
82: email: \texttt{yair@cs.huji.ac.il}.
83: {Supported in part by a grant from the Israeli Science Foundation 
84: (195/02).}}
85: \and
86: Manor Mendel\thanks{School of Computer Science, Hebrew University, Jerusalem 
87: 91904, Israel.
88: Email: \texttt{mendelma@gmail.com}.
89: {Supported by the Landau Center and a grant from the Israeli Science 
90: Foundation
91: (195/02).}}}
92: 
93: \begin{document}
94: 
95: \maketitle
96: 
97: \begin{abstract}
98: Metric embedding has become a common technique  in the design of
99: algorithms. Its applicability is often dependent on how high the
100: embedding's distortion is. For example, embedding finite metric space into
101: trees may require linear distortion as a function of its size. Using
102: probabilistic metric embeddings, the bound on the distortion
103: reduces to logarithmic in the size.
104: 
105: We make a step in the direction of bypassing the lower bound on
106: the distortion in terms of the size of the metric. We define
107: ``multi-embeddings" of metric spaces in which a point is mapped
108: onto a set of points, while keeping the target metric of
109: polynomial size and preserving the distortion of paths. The
110: distortion obtained with such multi-embeddings into ultrametrics
111: is at most $O(\log \Delta\log\log \Delta)$ where $\Delta$ is the
112: \emph{aspect ratio} of the metric. In particular, for expander
113: graphs, we are able to obtain {\em constant} distortion embeddings
114: into trees in contrast with the $\Omega(\log n)$ lower bound for
115: all previous notions of embeddings.
116: 
117: We demonstrate the algorithmic application of the new embeddings
118: for two optimization problems: \emph{group Steiner tree} and \emph{metrical 
119: task
120: systems}.
121: \end{abstract}
122: 
123: \begin{comment}
124: \begin{keywords}
125: Metric embeddings, Group Steiner tree, Metrical task systems
126: \end{keywords}
127: 
128: \begin{AMS}
129: 68W25, 68R10
130: \end{AMS}
131: 
132: \pagestyle{myheadings}
133: \thispagestyle{plain}
134: \markboth{Y. BARTAL AND M. MENDEL}{MULTI-EMBEDDING OF METRIC SPACES}
135: \end{comment}
136: 
137: %\clearpage
138: 
139: 
140: \section{Introduction}
141: 
142: Finite metric spaces and their analysis play a significant role in
143: the design of combinatorial algorithms. Many algorithmic
144: techniques were introduced in recent years concerning and using
145: metric spaces and their approximate embedding in other spaces, see
146: the surveys \cite{Ind01,IM04} for an overview of this topic.
147: 
148: %A metric space $M$ is a pair $(S,d)$, where $S$ is a set of points and
149: %$d:S\times S \rightarrow \mathbb{R}^+$ is a distance function that 
150: % satisfies:
151: %(i) $d(u,v)\geq 0$ for $u,v \in S$; (ii) $d(u,v)=0$ if and only if $u=v$;
152: %(iii) $d(u,v)=d(v,u)$ for all $u,v\in S$; (iv) $d(u,v)+d(v,w)\geq d(u,w)$ 
153: % for
154: %all $u,v,w\in S$. $d$ is also called the metric of $M$.
155: 
156: \begin{definition}
157: \label{def:embedding}
158: An embedding of a metric space $M=(V_M,d_M)$
159: into a metric space $N=(V_N,d_N)$ is a mapping $\phi:V_M
160: \rightarrow V_N$. The embedding is called \emph{non-contractive}
161: if for all $u,v \in V_M$, $d_M(u,v) \leq d_N(\phi(u),\phi(v))$ and
162: has distortion at most $\alpha$ if in addition for all $u,v \in
163: V_M$, $d_N(\phi(u),\phi(v)) \leq \alpha \cdot d_M(u,v)$. A
164: non-contractive embedding whose distortion is at most $\alpha$ is
165: called $\alpha$-embedding.
166: \end{definition}
167: 
168: The general framework for
169: applying metric embeddings in optimization problems is to embed a
170: given metric spaces into a metric space from some ``nice" family
171: and then apply an algorithm for that space. As a result, the
172: approximation ratio increases by a factor equal to the embedding's
173: distortion.
174: 
175: Among others, embeddings into low dimensional normed spaces
176: \cite{Bou85,LLR95} as well as probabilistic embeddings into trees
177: \cite{Bar96,Bar98,FRT03,Bar03} have many algorithmic applications.
178: In both cases the distortions of the embeddings are logarithmic in
179: the size of the metric. Unfortunately, there is a matching lower
180: bound on the distortion of these embeddings as well, which sets a
181: limit to their applicability. This paper presents a partial remedy
182: for this problem.
183: 
184: Tree metrics, and in particular {ultrametrics}, seem a natural
185: choice as a target class of ``simple" metric spaces.
186: Unfortunately, standard embedding is not useful when the target
187: space is a tree metric. Embedding arbitrary metric spaces into
188: trees requires distortion linear in the size of the metric
189: space~\cite{RR98}. Probabilistic embedding~\cite{Bar96} provides a
190: way to bypass this problem:
191: 
192: \begin{definition}[Probabilistic Embeddings] \label{def:prob-approx}
193: A metric space $M=(V_M,d_M)$ is $\alpha$-\emph{probabilistically
194: embedded} in a set of metric spaces $\mathcal{S}$ if there exists
195: a distribution $\mathcal{D}$ over $\mathcal{S}$ and for every
196: $N\in\mathcal{S}$, a non-contractive embedding $\phi_N:V_M
197: \rightarrow V_N$, such that for all $u,v \in V_M$, $\E_{N\in
198: \mathcal{D}}[d_N(\phi_N(u),\phi_N(v))] \leq \alpha \cdot
199: d_M(u,v)$.
200: \end{definition}
201: 
202: Using probabilistic embeddings, it is possible to obtain much
203: better bounds on the distortion
204: \cite{AKPW95,Bar96,Bar98,FRT03,Bar03}. The following bound is
205: shown in \cite{FRT03,Bar03}:
206: 
207: \begin{theorem}  \label{thm:bar98}
208: Any metric space on $n$ points can be  $O(\log n )$
209: probabilistically embedded in a set of $n$-point {ultrametrics}.
210: Moreover, the distribution can be sampled efficiently.
211: \end{theorem}
212: 
213: Theorem~\ref{thm:bar98} found many algorithmic applications in approximation
214: algorithms, online algorithms, and distributed algorithms, see for
215: example~\cite{Bar96,GKR98,FM00,KT99,BCR01}. The bound on the distortion in
216: Theorem~\ref{thm:bar98} is tight even for probabilistic embeddings into tree
217: metrics for which there is an $\Omega(\log n)$ lower bound \cite{Bar96}.
218: 
219: Theorem~\ref{thm:bar98} was originally formulated for a class of metric
220: spaces defined by the following natural generalization of ultrametrics:
221: 
222: \begin{definition}[\cite{Bar96}] \label{def:hst}
223: For $k\geq 1$, a $k$-\emph{hierarchically well-separated tree}
224: ($k$-HST) is a metric space defined on the leaves of a rooted tree
225: $T$. To each vertex $u\in T$ there is associated a label
226: $\Delta(u) \ge 0$ such that $\Delta(u)=0$ if and only if $u$ is a leaf of
227: $T$. The labels are such that if a vertex $v$ is a child of a
228: vertex $u$ then $\Delta(v)\leq \Delta(u)/k$. The distance between
229: two leaves $x,y\in T$ is defined as $\Delta(\lca(x,y))$, where
230: $\lca(x,y)$ is the least common ancestor of $x$ and $y$ in $T$.
231: \end{definition}
232: 
233: The definition of finite ultrametric is the same as a 1-HST. Any $k$-HST
234: is therefore, in particular, an ultrametric and any finite ultrametric
235: can be $k$-embedded in some $k$-HST \cite{Bar98}. We can
236: therefore restrict our attention to ultrametrics, while all
237: results generalize to $k$-HSTs.
238: 
239: The main contribution of this paper is in offering a new type of metric
240: embedding that makes it possible to bypass lower bounds for the standard and
241: even probabilistic metric embeddings. There are two key observations that
242: lead to this new type of embedding. The first is that in some applications 
243: it
244: is natural to match a point onto a set of points in the target metric space.
245: Motivated by two applications of Theorem~\ref{thm:bar98}: the
246: \emph{group Steiner tree problem} (henceforth, GST), and the
247: \emph{metrical task systems problem} (henceforth, MTS),
248: we propose the following definition:
249: 
250: \begin{definition}[Multi Embedding] \label{def:multi-embedding}
251: A multi embedding of $M$ in $N$ is a partial surjective function
252: $\f$ from $N$ on $M$, i.e. each point $x\in M$ is embedded into a
253: non-empty set $\f^{-1}(x)$. Points in $\f^{-1}(x)$ are called
254: \emph{representatives} of $x$ in $N$.
255: \end{definition}
256: 
257: The role of $\f^{-1}$ in Definition~\ref{def:multi-embedding} is
258: analogous to the role of $\phi$ in the
259: Definitions~\ref{def:embedding} and \ref{def:prob-approx} of
260: embedding and probabilistic embedding. Another way to define multi
261: embedding is by $\phi:M \to 2^N$, in which $\phi(u)\cap
262: \phi(v)=\emptyset$ for every $u\neq v$. In our notation we have
263: $\phi(x) = \f^{-1}(x)$. Since the $\f$ notation will be more
264: convenient, henceforth we will exclusively use it.
265: 
266: The second observation is that for many applications, including
267: those mentioned above, there is no need to approximate the
268: original distance for every pair of representatives. What is
269: really needed is that every path in the original space will be
270: approximated well by \emph{some} path in the target space.
271: 
272: A path in a metric space is an arbitrary finite sequence of points in the
273: space. The length of a simple path $p=\la u_1,u_2,\ldots,u_m \ra$ in a 
274: metric space
275: $M=(V,d)$ is defined as \( \ell(p)=\sum_{i=1}^{m-1} d(u_i,u_{i+1}) \).
276: 
277: \begin{definition}[Path Distortion] \label{def:path-approx}
278: A multi-embedding $\f$ of $M$ in $N$, is called \emph{non-contractive}
279: if for any $u,v\in N$, $d_N(u,v) \geq d_M(\f(u),\f(v))$.
280: The path-distortion of a non-contractive multi-embedding of $M$ in $N$, 
281: $\f:N\to M$,  is
282: the infimum over $\alpha$, for which
283: any path $p=\la
284: u_1,u_2,\ldots,u_m \ra$ in $M$, has a path $p'=\la
285: u'_1,u'_2,\ldots,u'_m \ra $ in $N$ such that $\f(u'_i)=u_i$ and
286: $\ell(p') \leq \alpha \cdot \ell(p)$.
287: 
288: A multi embedding  whose path-distortion is
289: at most $\alpha$ is called $\alpha$-path embedding.
290: \end{definition}
291: 
292: A crucial parameter for
293: multi-embeddings is the size of the target space
294: $\Gamma$.
295: In general, it will be desirable that
296: $\Gamma$ will be polynomial in the size of the source space.
297: In fact, if $\Gamma = \infty$ then there is a simple
298: $1$ path embedding of any finite metric space by trees:
299: Take all finite paths, convert each path to a simple path (by duplicating
300: points, if necessary), and put them under a single root with an edge of
301: length half the diameter.
302: This motivates a study of
303: the trade-off between $\Gamma$ and the path distortion of arbitrary metric
304: spaces by tree metrics. In Section~\ref{sec:trees} we study path
305: embedding of expander graphs and the hypercube into tree metrics. We show
306: e.g. that an $n$-point Ramanujan graphs have 3-path-embedding
307: into tree metrics of size
308: $\Gamma(n)\leq\text{poly}(n)$. This is in sharp contrast to the
309: status of expander graphs for previous notions of embeddings for which they
310: are considered ``worst case" examples with $\Omega(\log n)$ distortion
311: \cite{LLR95}. These results directly imply nearly tight results on the
312: approximation ratio for GST on expander graphs and hypercubes.
313: 
314: We consider multi embeddings when the class of target metric
315: spaces are {ultrametrics}. First, we observe that probabilistic
316: embedding into {ultrametrics} directly implies a bound for path
317: embedding by putting all the trees in $\mathcal{S}$ (the set used
318: in the probabilistic embedding) under a common new root. This
319: results with an $\alpha$-path embedding into {an ultrametric} of
320: size $ |\mathcal{S}|n$. Using the bound of \cite{CCGGP98} on the
321: number of {ultrametrics} needed in Theorem~\ref{thm:bar98} we
322: obtain an $O(\log n )$ path embedding into {an
323: ultrametric} of size $O(n^2 \log n)$.%
324: \footnote{In a preliminary version of this paper~\cite{BM03}, we
325: also introduced the notion of \emph{probabilistic
326: multi-embedding}. Using that notion we were able to show
327: probabilistic multi-embedding into {ultrametrics} of polynomial
328: size and path distortion $O(\log n \log\log \log n)$. Since an
329: $O(\log n)$ bound now follows from
330: Theorem~\ref{thm:bar98}~\cite{FRT03,Bar03}, we have decided to
331: drop the probabilistic multi-embedding result from this version.}
332: 
333: An important parameter of the metric spaces appearing in practice
334: is the aspect ratio of the metric, which is the ratio between the
335: diameter and minimum non-zero distance in the metric space. It
336: will be convenient for us to assume that the minimum distance is
337: 1, and so the aspect ratio becomes the diameter. It turns out that
338: the aspect ratio of the metric plays a significant role in the
339: path distortion of multi-embeddings. In
340: Section~\ref{sec:path-approx} we prove:
341: 
342: \begin{theorem} \label{thm:path-approx}
343: Fix $\beta>1$. For any metric space $M=(V,d)$ on $|V|=n$ points
344: and aspect ratio $\Delta$, there exists an efficiently
345: constructible multi-embedding into an ultrametric of size
346: $n^\beta$, whose path distortion is at most
347: \[
348: O_\beta( \min\{\log n \cdot \log \log n, \;\log \Delta \cdot \log\log \Delta 
349: \}). \]
350: \end{theorem}
351: 
352: Our construction beats the probabilistic embedding based
353: constructions on metrics with small aspect ratio. Expander graphs
354: are examples where a lower bound of $\Omega(\Delta)$ exists on
355: probabilistic embedding using trees \cite{LLR95}.
356: 
357: The constructions of multi-embeddings are in a sense dual to
358: Ramsey-type theorems for metric spaces \cite{BBM01,BLMN03}, where
359: the goal is to find a large subset which is well approximated by some
360: {ultrametric}.
361: % , and some of the techniques in the proof of
362: % Theorem~\ref{thm:path-approx} are dual to those of
363: % \cite{BBM01,BLMN03}.
364: 
365: We also provide a simple example in which
366: Theorem~\ref{thm:path-approx} is almost tight: Any $\alpha$-path
367: embedding into {ultrametrics} of a simple unweighted path of
368: length $n$ has $\alpha=\Omega(\log n)$. It follows, in particular,
369: that any $\alpha$-path embedding into {ultrametrics} of the metric
370: defined by an unweighted graph of diameter $\Delta$ has
371: $\alpha=\Omega(\log \Delta)$. 
372: Path embedding is motivated by two
373: intensively-studied algorithmic minimization problems: GST and
374: MTS, mentioned above. For both, the best known algorithms use
375: probabilistic embedding into trees/ultametrics. In
376: Section~\ref{sec:appl} we prove that in order to reduce these
377: problems to other metric spaces it is sufficient to use path
378: embedding. We therefore achieve improved algorithms for these
379: problems whenever the path embedding distortion beats that of
380: probabilistic embedding, and in particular, when the underlying
381: metric is of small aspect ratio.
382: 
383: \section{Applications} \label{sec:appl}
384: 
385: In this section we define MTS and GST
386: show that path distortion of multi-embeddings reduces
387: these problems to similar problems with different underlying metrics.
388: 
389: Metrical Task Systems (MTS)~\cite{BLS92} was introduced as a
390: framework for many online minimization problems. A MTS on metric
391: space $M=(S,d)$, $|S|=n$, is defined as follows. A ``system" has a
392: set of $n$ possible internal states $S$.  It receives a sequence
393: of \emph{tasks} $\sigma=\tau_1\tau_2\cdots \tau_m$. Each task
394: $\tau$ is a vector $\tau:S\rightarrow \mathbb{R}^+ \cup \{
395: \infty\}$ of nonnegative costs for serving $\tau$ in each of the
396: internal states of the system. The system may switch states (say
397: from $u$ to $v$), paying a cost equal to the distance $d(u,v)$ in
398: $M$, and then pays the service cost $\tau(v)$ associated with the
399: new state. The major limiting factor for the system is the
400: requirement to process the sequence in an online fashion, \ie,
401: serving each task without knowing the future tasks.
402: 
403: As customary in the analysis of online algorithms, MTS
404: is analyzed using the notion of \emph{competitive ratio}. A randomized 
405: online
406: algorithm $A$ is called $r$-competitive if there exists some constant $c$ 
407: such that for any
408: task sequence $\sigma$, $\E[\cost_A(\sigma)]\leq r\cdot 
409: \cost_{\opt}(\sigma)+c$,
410: where $\cost_A(\sigma)$ is the random variable of the cost for serving
411: $\sigma$ by $A$, and $\cost_{\opt}(\sigma)$ is the optimal (offline) cost 
412: for
413: serving $\sigma$. The current best online algorithm for the MTS problem in
414: $n$-point metric spaces is $O(\log^2 n \log \log n)$ competitive
415: \cite{FM00,FRT03} (an improvement of \cite{BBBT97,Bar98}). Both papers
416: \cite{BBBT97,FM00} actually solve the MTS problem for {ultrametrics}, and
417: then reduce arbitrary metric spaces to {ultrametrics} using
418: Theorem~\ref{thm:bar98}. We next show that path embedding suffices:
419: 
420: 
421: \begin{proposition} \label{prop:MTS-appl}
422: Assume that a metric space $M$ is $\alpha$-path
423: embedded in $N$.
424: Assume also that $N$
425: has an $r$-competitive MTS algorithm. Then
426: there is an $\alpha r$-competitive algorithm
427: for $M$.
428: \end{proposition}
429: \begin{proof}
430: We construct an online algorithm $\algon$ for $M$ as
431: follows:
432: Let $\algon_{N}$ be an
433: $r$-competitive online algorithm for $N$, and $\f:N\to M$ an $\alpha$ path 
434: embedding of $M$ in $N$.
435: The task
436: sequence $\sigma$ is translated to a task sequence $\sigma^N$ for $N$ task 
437: by
438: task as follows.
439: A task $\tau$ for $M$ is translated into a task $\tau^N$ for
440: $N$ such that $\tau^N(u')=\tau(\f(u'))$. $\algon$ maintains the
441: invariant that if $\algon_{N}$ is in state $v'$, then $\algon$ is in
442: state $\f(v')$.
443: 
444: It is easy to verify that $\cost_{\algon}(\sigma)\leq
445: \cost_{\algon_{N}}(\sigma^N)$, since the service costs are the same,
446: and the distances in $N$ are larger. Consider $\opt(\sigma)$, it
447: defines a path $p$ of serving $\sigma$ in $M$. Thus there exists a path 
448: $p^N$ as
449: in the statement of Definition~\ref{def:path-approx}. The path
450: $p^N$ is the way $\sigma^N$ would be served in $N$. In this way, since
451: $\f(p^N)=p$, the service costs in $N$ are the same as the services costs
452: in $M$, and $\ell(p^N) \leq \alpha
453: \ell(p)$. Thus $\cost_{\opt_{N}}(\sigma^N)
454: \leq \alpha \cdot \cost_{\opt}(\sigma)$. Summarizing:
455: \begin{equation*}
456: \E[\cost_{\algon}(\sigma) ] \leq
457: \E[\cost_{\algon_{N}}(\sigma^N)]
458: \leq r\cdot \cost_{\opt_{N}}(\sigma^N) +c
459: \leq \alpha r \cdot \cost_{\opt}(\sigma) +c.
460: \end{equation*}
461: \end{proof}
462: 
463: 
464: \begin{corollary} \label{corol:MTS1}
465: There is an $O(\log\Delta \log\log \Delta \cdot \log n\log\log
466: n)$-competitive randomized MTS algorithm for MTS defined on metric
467: spaces with diameter $\Delta$.
468: \end{corollary}
469: \begin{proof}
470: Apply Theorem~\ref{thm:path-approx} on the original metric and obtain an
471: $O(\log \Delta \log\log
472: \Delta)$ path embedding into an ultrametric of size  $\Gamma(n) = \poly(n)$.
473: This ultrametric has $O(\log
474: \Gamma(n)\, \log\log \Gamma(n))$ competitive algorithm  \cite{FM00}. Now
475: apply  Proposition~\ref{prop:MTS-appl} to obtain the claim.
476: \end{proof}
477: 
478: \medskip
479: 
480: The Group Steiner Tree Problem (GST)~\cite{RW90} can be stated as
481: follows: Given a graph $G=(V,E)$ on $n$ vertices with a weight
482: function $c:E\rightarrow \mathbb{R}_+$, and subsets of the
483: vertices $g_1,\ldots, g_k \subset V$ (called \emph{groups}), the
484: objective is to find a minimum weight subtree $T$ of $G$ that
485: contains at least one vertex from each $g_i$, $i\in [k]$. Under
486: certain standard complexity assumptions, this is hard to
487: approximate by a factor better than $\max\{ \log^{2-\eps} k,
488: \log^{2-\eps} n\}$~\cite{HK03}. The current best upper bound on
489: the approximation factor is $O(\log^2n \log k )$
490: \cite{GKR98,FRT03}. In \cite{GKR98}, an $O(\log n \log k)$
491: approximation algorithm for tree metrics is given, and the general
492: case is reduced to tree metrics using Theorem~\ref{thm:bar98}.
493: Again, we show that it is actually sufficient to use multi
494: embedding for this problem.
495: 
496: As a first step we observe that the problem can be easily cast in terms of
497: metric spaces instead of graphs: Given a graph $G=(V,E)$ with weights
498: $w:E\to \mathbb{R}_+$, let
499: $M=(V,d)$ be the shortest path metric induced by $G$ and $w$ on $V$.
500: A tree
501: $T$ in $M$ can be transformed into a tree $\hat{T}$ in $G$ such that the
502: total weight in $\hat{T}$ is not larger than the total weight in $T$, and
503: $\hat{T}$ contains all the vertices in $T$. This is done by replacing each
504: edge in $T$ by the shortest path between its endpoints in $G$, and taking a
505: spanning tree of the resulting subgraph. It therefore suffices to solve GST
506: on metric spaces.
507: 
508: \begin{proposition} \label{prop:gsp-appl}
509: Assume that a metric space $M$ is $\alpha$ path embedded in a metric space 
510: $N$.
511: Assume in addition that there is a [randomized] polynomial time
512: $r$ approximation algorithm for any GST instance with $k$ groups
513: defined on $N$. Then there exists a [randomized] polynomial time
514: $2\alpha r$-approximation algorithm for any GST instance with $k$ groups
515: defined on $M$.
516: \end{proposition}
517: \begin{proof}
518: We construct an approximation algorithm $\A$ for the instance
519: \linebreak
520: $\sigma=(M; g_1, \ldots, g_k)$ as follows. Denote by
521: $\f:N\to M$ the $\alpha$ path embedding of $M$ in $N$. Consider
522: the following instance of GST: $\sigma_N=(N;\f^{-1}(g_1), \ldots
523: \f^{-1}(g_k))$. Let $\A_{N}$ be an $r$-approximation algorithm for
524: $\sigma_N$. Let $T_N= \A_N(\sigma_N)$ be the tree constructed by
525: $A_N$. Denote by $\f(T_N)$ the image graph of $T_N$. \Ie, if
526: $T_N=(V_N,E_N)$, then $\f(T_N)=(\f(V_N), \{ \f(u) \f(v)|\; u v\in
527: E_N\})$. The graph $\f(T_N)$ is a connected and its weight is at
528: most the weight of $T_N$. It also spans at least one
529: representative form each group. Algorithm $\A$ returns a spanning
530: tree of $\f(T_N)$. This tree is a feasible solution and it
531: satisfies $\cost_{\A}(\sigma)\leq \cost_{\A_{N}}(\sigma_N)$.
532: 
533: Consider the tree $\opt(\sigma)$, double each edge in $\opt(\sigma)$ and 
534: take
535: an Euler tour $p$ of this graph. There exists a path in $N$, $p_N$, as in 
536: the
537: statement of Definition~\ref{def:path-approx}, such that $\f(p_N)=p$. The
538: path $p_N$ is a connected graph and spans at least one representative from
539: each group $\f^{-1}(g_j)$. As the weight of $p$ is twice the weight of
540: $\opt(\sigma)$, we have
541: \[\cost_{\opt_{N}}(\sigma_N) \leq  \ell(p_N) \leq
542: \alpha \ell(p)\leq 2\alpha \cost_{\opt}(\sigma). \]
543: Summarizing:
544: \begin{equation*}
545: \E[\cost_{\A}(\sigma) ]\leq \E[\cost_{\A_{N}}(\sigma_N)]
546: \leq r\, \cost_{\opt_{N}}(\sigma_N)  \leq 2\alpha r \,
547: \cost_{\opt}(\sigma) .
548: \end{equation*}
549: \end{proof}
550: 
551: \begin{corollary} \label{corol:gsp1}
552: There is an polynomial time $O(\log\Delta \log\log \Delta \log n\log k)$ 
553: approximation
554: algorithm for GST on metric spaces with diameter $\Delta$.
555: \end{corollary}
556: \begin{comment}
557: \begin{proof}
558: We apply Proposition~\ref{prop:gsp-appl} to
559: Corollary~\ref{cor:path-approx} and so we obtain $\alpha=\lambda$
560: and $\Gamma = \poly(n)$. Using the algorithm for trees from
561: \cite{GKR98} with $r(m)=O(\log m \log k)$, we obtain the claim.
562: \end{proof}
563: \end{comment}
564: 
565: 
566: \section{Multi Embedding into Ultrametrics} \label{sec:path-approx}
567: 
568: The following theorem is a restatement of
569: Theorem~\ref{thm:path-approx} in a more general form.
570: 
571: \begin{theorem}
572: Given any metric space $M=(V,d)$ on $|V|=n$ points and  diameter
573: $\Delta$, for any $t\in \mathbb{N}$, $M$ is $O(t\min\{\log \Delta,
574: \log n\})$ path embedded into an efficiently constructible
575: {ultrametric} of size  $\Gamma \leq n^\beta$, where \( \beta = \min \{(\log
576: n)^{1/t}, [t \log (4\Delta)]^{2/t}\}. \)
577: \end{theorem}
578: \begin{proof}
579: The construction of the multi-embedding is motivated by the
580: construction of subspaces approximating {ultrametric} in
581: \cite{BBM01,BLMN03}, but instead of deleting points, we
582: \emph{duplicate} them. We then prove the bounds on the
583: path distortion.
584: 
585: Let $\Delta$ be the diameter of $M$. Let $x$ and $\bar{x}$ be two points
586: realizing the diameter of $M$, and assume without loss of generality that
587: $|\{ y\in M:\ d(x,y)<\Delta/4\}|\leq n/2$ (otherwise, switch the roles of 
588: $x$
589: and $\bar{x}$). Define a series of sets $A_0=\{x\}$, and for $i\in
590: \{1,2,\ldots,t \}$, $A_i=\{y\in M |\ d(x,y)< i \Delta /4t\}$, and ``shells"
591: $S_i=A_i\setminus A_{i-1}$. Let $|V|=n$ and let $\eps_i= |A_i|/n$.
592: 
593: The algorithm for constructing the multi-embedding works as
594: follows: Choose a shell $S_{i}$, $i\in [t]$. Recursively,
595: construct a multi-embedding of the sub space $A_{i}$ into {an
596: ultrametric} $T_1$ and a multi-embedding of the subspace
597: $V\setminus A_{i-1}$ into {an ultrametric} $T_2$. To construct the
598: multi-embedding for $M$, we construct {an ultrametric} $T$ with
599: root labelled with $\Delta$, and two children, one is $T_1$ and
600: the other is $T_2$. This is a multi-embedding since the points in
601: $S_i$ are essentially being ``duplicated" at this stage. Note that
602: this is a non-contractive multi-embedding.
603: 
604: Next we prove an upper bound on the size of the resulting {ultrametric}
605: $T$, assuming that the shell was chosen carefully enough. The bound we
606: prove is $n^\beta$, where $\beta = \min \{(\log n)^{1/t}, [t \log
607: (4\Delta)]^{2/t}\}$.
608: 
609: We begin with the first bound. Let
610: $\beta=\beta(n)= (\log n)^{1/t}$. The proof proceeds by induction
611: on $n$ (whereas $t$ is fixed).
612: There must exist $i\in [t]$ such that $\eps_{i-1} \geq \eps_i^\beta$. 
613: Indeed,
614: note that $n^{-1}\leq \eps_0 \leq \eps_{t+1} \leq 1/2$. Assume for the
615: contrary that $\eps_{i-1}<\eps_i^\beta$ for all $i\in[t]$, then
616: \[ \eps_0 <\eps_1^\beta <\cdots \eps_t^{\beta^t}\leq
617: (\frac{1}{2})^{\log n} =\frac{1}{n} ,\]
618: which is a contradiction. Therefore we can fix $i$ such that $\eps_{i-1} 
619: \geq
620: \eps_i^\beta$. Inductively, assume that the recursive process results in at
621: most $(\eps_{i}n)^{\beta(\eps_i n)} \leq (\eps_{i}n)^\beta$
622: leaves in $T_1$ and at most
623: $((1-\eps_{i-1})n)^{\beta((1-\eps_{i-1})n)}\leq ((1-\eps_{i-1})n)^{\beta}$
624: leaves in $T_2$. So \( |T|\leq (\eps_{i}^{\beta}
625: + (1-\eps_{i-1})^{\beta}) n^{\beta}\). Since $\eps_{i-1} \geq
626: \eps_{i}^\beta$, we have \( \eps_{i}^{\beta}+ (1-\eps_{i-1})^{\beta} \leq
627: \eps_{i-1} + (1- \eps_{i-1})=1 \) and we are done.
628: 
629: We next prove the second bound. Let $\beta=\beta(\Delta) = [t \log
630: (4\Delta)]^{2/t}$. The proof is by induction on (the rounded value of)
631: $\Delta$. We claim that
632: \begin{eqnarray}\label{eq:condition}
633: \exists i\in [t] \qquad \eps_{i-1}\ge \eps_{i}^{\beta(\Delta/2)}
634: n^{\beta(\Delta/2)- \beta(\Delta)}.
635: \end{eqnarray}
636: Indeed, assume for the contrary that no such $i$ exists. Set $a=\log
637: (2\Delta)\ge 1$, so that $\beta(\Delta/2)=(ta)^{2/t}$ and $\beta(\Delta) =
638: [t(a+1)]^{2/t}$. Denote $b=n^{(ta)^{2/t}-[t(a+1)]^{2/t}}$ and 
639: $c=(ta)^{2/t}$.
640: The opposite of \eqref{eq:condition} then becomes
641: $\eps_{i-1}<\eps_i^c b$, for any $i\in[t]$. Iterating this $t$ times
642: we get:
643: \[
644: \frac{1}{n}=\eps_0 < \eps_t^{c^t} b^{1+c+c^2+\ldots+c^{t-1}} \le
645: \eps_t^{c^t} b^{c^{t-1}} \le b^{c^{t-1}} .
646: \]
647: So that:
648: \[
649: n^{(ta)^{2-2/t}\left[[t(a+1)]^{2/t}-(ta)^{2/t}\right]}<n,
650: \]
651: but this is a contradiction, since an application of the mean value theorem
652: implies the existence of $\xi \in [a,a+1]$, for which
653: \begin{multline*}
654: (ta)^{2-2/t}\left[[t(a+1)]^{2/t}-(ta)^{2/t}\right] = % \\ =
655:    (ta)^{2-2/t} [2 t^{-1+2/t}\xi^{-1+2/t}]
656: =  2 (ta) \left(\frac{a}{\xi}\right)^{1-2/t} \ge ta \ge 1 .
657: \end{multline*}
658: 
659: Choose an index $i\in \{1,\ldots, t\}$ satisfying (\ref{eq:condition}). 
660: Since
661: $i\le t$, $\Delta(A_{i})\le \Delta/2$. The choice of the index $i$, and 
662: using
663: the inductive hypothesis, gives the required lower bound on the cardinality
664: of $T$ since:
665: {%\setlength\arraycolsep{1pt}
666: \begin{align*}
667: |T|
668: % &\le |A_{i}|^{\beta(\Delta(A_{i}))}+(n-|A_{i-1}|)^{\beta(\Delta(M\setminus 
669: % A_{i-1}))}\\
670: \le (\eps_{i}n)^{\beta(\Delta/2)}+[(1-\eps_{i-1})n]^{\beta(\Delta)} \le
671: \eps_{i-1}n^{\beta(\Delta)}+(1-\eps_{i-1})n^{\beta(\Delta)}\le
672: n^{\beta(\Delta)}.
673: \end{align*}}
674: 
675: 
676: We note that the running time of the algorithm above is $O(n^2)$
677: on each vertex in the tree and therefore $O(n^{\beta+2})$ for the
678: whole tree . A slight variation on this algorithm (and a more
679: careful analysis) has an ${O}(n^{\max\{\beta,2\}})$ running time.
680: 
681: \begin{comment}
682: To reach path-distortion of $O(n^\beta \log \log n)$,
683: we don't need to compute the diameter.
684: Instead
685: it suffices to take arbitrary $x$ and the furthest point $\bar{x}$ . So for 
686: each vertex in the HST
687: we only need to invest $O(n)$ time. This gives $O(n^{\beta+1})$.
688: To improve upon this we do a more careful inductive argument
689: \begin{multline*}
690: T(n)\leq T(\eps_i n)+ T((1-\eps_{i-1})n) + O(n) \\
691:    \leq c_\beta \left(\eps_{i}^{\max\{2,\beta\}} 
692: +(1-\eps_{i-1})^{\max\{2,\beta\}}\right) n^{\max\{2,\beta\}} +O(n) \\
693:   \leq \c_\beta \left(\eps_{i-1}^{\max\{2/\beta,1\}} 
694: +(1-\eps_{i-1})^{\max\{2,\beta\}}\right ) n^{\max\{2,\beta\}} +O(n) \\
695: \leq c_\beta n^{\max\{\beta,2\}}
696: \end{multline*}
697: \end{comment}
698: 
699: 
700: The multi-embedding described above has the following properties:
701: \begin{enumerate}
702: \item The multi-embedding is non-contractive.
703: \item The tree structure defining the {ultrametric} is a binary
704: tree.\footnote{Note that more generally, any {ultrametric} can be
705: defined by a binary tree.}
706: \end{enumerate}
707: Let $u$ be an internal vertex in the binary tree defining the ultrametric,
708: and $T$ the subtree rooted at $u$.
709: We can rename the subtrees rooted with the children of $u$, as $T_1$
710: and $T_2$ such that:
711: \begin{enumerate}
712: \setcounter{enumi}{2}
713: \item
714: \label{item:stam} Let $x$ and $y$ two points in $M$. If $\emptyset \neq 
715: \f^{-1}(x) \cap T \subset T_1$
716: and $\emptyset \neq \f^{-1}(y) \cap T \subset T_2$, then
717: $d(x,y)\geq \Delta(T)/4t$.
718: \item $|\f(T_1)| \leq |\f(T)|/2$.
719: \item
720: $\Delta(T_1)\leq \Delta(T)/2$.
721: \end{enumerate}
722: 
723: We next show, using the properties above, that the path distortion
724: of this multi embedding is at most $8t \log \min\{n,\Delta\}$. Let
725: $p=\la u_1, u_2,\ldots, u_{m} \ra$ be a path in $M$ whose length
726: is $\ell(p)$ . We construct a path $\bar{p}$ on the leaves of $T$
727: whose length satisfies $\ell(\bar{p}) \leq 8t \log
728: \min\{n,\Delta\} \cdot \ell(p)$. The proof proceeds by induction
729: on the height of the tree defining the {ultrametric}.
730: 
731: \begin{figure}[ht]
732: \centering
733: \input{path.pstex_t}
734: \caption{A partition of the path to subpaths.}
735: \label{figure}
736: \end{figure}
737: 
738: We partition $p$ into sub-paths as follows. Define a sequence of indices and 
739: a
740: sequence of sub-trees of the root: Let $j_1=1$ and let
741: $\hat{T}_1\in\{T_1,T_2\}$ be the subtree of the root that includes the
742: longest prefix of $p$. Assume inductively that we have already defined
743: $j_{i-1}$ and $\hat{T}_{i-1}\in\{T_1,T_2\}$. Define $j_i$ to be the minimum 
744: index
745: such that $u_{j_i}$ is the first point in $p$ after $u_{j_{i-1}}$ with no
746: representative in $\hat{T}_{i-1}$. Let $\hat{T}_i\in\{T_1,T_2\}$ be the 
747: other
748: subtree of the root. Assume this process is finished with $j_{s}$,
749: $\hat{T}_{s}$. Next we define another sequence of indexes: $k_{s}=m$, for
750: $i<m$ we define $k_i$ to be the largest number, smaller than $j_{i+1}$, such
751: that $u_{k_i}$ does not have a representative in $\hat{T}_{i+1}$. By the
752: construction of $\hat{T}_i$, we have that $j_i \leq k_i$ and $u_{k_i}$ has a
753: representative in $\hat{T}_i$.  See Figure~\ref{figure} for example of such 
754: partition.
755: We have partitioned $p$ into sub-paths $(\la
756: u_{j_i},\ldots, u_{k_i} \ra)_i$ and $(\la u_{k_i},\ldots, u_{j_{i+1}}
757: \ra)_i$. Informally, a sub-path $\la u_{j_i},\ldots, u_{k_i} \ra$ will be
758: realized in $\hat{T}_i$, while sub-path $\la u_{k_i},\ldots, u_{j_{i+1}} 
759: \ra$
760: will be realized in $T_1$.
761: 
762: 
763: More formally,  let $L=\ell(p)$,  $L_{i_1,i_2}=\ell(\la
764: u_{i_1},u_{i_1+1},\ldots, u_{i_2} \ra)$, $n=|\f(T)|$, $n_1=|\f(T_1)|$,
765: $n_2=|\f(T_2)|$, and $\Delta=\Delta(T)$, $\Delta_1=\Delta(T_1)$,
766: and $\Delta_2=\Delta(T_2)$. We construct by induction on the tree structure
767: $T$ a path $\bar{p}$ in $T$ whose length satisfies \( \bar{L}=
768: \ell(\bar{p}) \leq 8t \log \min\{\Delta,n\} \cdot L . \)
769: 
770: By the induction hypothesis it is possible to construct for any
771: $i$, a path in $\hat{T}_i$ of representatives of $\la
772: u_{j_i},\ldots,u_{k_i} \ra$ whose length is
773: \[
774: \bar{L}_{j_i,k_i} %\leq L _{j_i,k_i} \cdot 8t \log \min \{n_i,\Delta_i\}
775: \leq L_{j_i,k_i} \cdot 8t  \log \min\{n,\Delta\}.
776: \]
777: Next, for any $i$, we construct a path of representatives of $ \la
778: u_{k_i}, \ldots ,u_{j_{i+1}}\ra$. Note that $u_{k_{i}+1},\ldots,
779: u_{j_{i+1}-1}$ have representatives in both $\hat{T}_i$ and
780: $\hat{T}_{i+1}$. Therefore, we construct inductively a path from a
781: representative of $u_{k_i+1}$ to a representative of
782: $u_{j_{i+1}-1}$ in $T_1$, so $\bar{L}_{k_i,j_{i+1}} \leq L
783: _{k_i,j_{i+1}} 8t \log\min\{n_1,\Delta_1\}$. We then connect the
784: representative of $u_{k_i}$ with the representative of $u_{k_i+1}$
785: and the representative of $u_{j_{i+1}-1}$ with the representative
786: of $u_{x_{i+1}}$, each such edge is of length at most the diameter
787: of $T$, $\Delta$. We have therefore constructed a path of
788: representatives of $ \la u_{k_i}, \ldots, u_{j_{i+1}}\ra$ whose
789: length is \( \bar{L}_{k_i,j_{i+1}} +2\Delta\).
790: 
791: Since $u_{k_i}$ does not have a representative in $\hat{T}_{i+1}$
792: and $u_{j_{i+1}}$ does not have representative in $\hat{T}_i$, we conclude
793: using
794: property~(\ref{item:stam}) above, that $d_M(u_{k_i},u_{j_{i+1}})
795: \geq \Delta /4 t$, and so $\Delta \leq 4t \cdot L_{k_i,j_{i+1}}$. To
796: summarize,
797: \begin{align*}
798: \bar{L}_{k_i,j_{i+1}} & \leq   L_{k_i,j_{i+1}} 8t \log
799: \min\{n_1,\Delta_1\}
800: +2\Delta \\
801: & \leq L_{k_i,j_{i+1}} 8t \bigl( \log \min\{n/2, \Delta/2\} + 1\bigr) \\
802: &= L_{k_i,j_{i+1}} 8t  \log \min\{n,\Delta\}
803:    .
804: \end{align*}
805: We conclude,
806: \begin{align*}
807: \bar{L} &= \sum_{i=1}^s \bar{L}_{j_i,k_i} + \sum_{i=1}^{s-1}
808: \bar{L}_{k_i,j_{i+1}} \\
809: & \leq 8t \log \min\{n,\Delta\} \cdot  \Bigl(\sum_{i=1}^s 
810: L_{j_i,k_i}+\sum_{i=1}^{s-1} L_{k_i,j_{i+1}} \Bigr )
811: \\ & =  8t \log \min\{n,\Delta\} L  .
812: \end{align*}
813: \end{proof}
814: 
815: 
816: We end the discussion on multi embedding into
817: {ultrametrics} with the following impossibility result.
818: \begin{proposition} \label{prop:inapprox}
819: Consider the metric defined by a simple $N$-point path. Then any
820: $\alpha$ path-embedding of this metric in an
821: {ultrametric} must have $\alpha =\Omega( \log n)$.
822: \end{proposition}
823: \begin{proof}
824: Let $M=\{v_1,v_2,\ldots,v_n\}$ be the metric space on $n$ points
825: such that $d_M(v_i,v_j)=|i-j|$. We prove that for any non-contractive
826: multi-embedding into {an ultrametric} $T$, any path of representatives
827: of $\la v_1, v_2,\ldots,v_n \ra$ is of length at least
828: $g(n)=\tfrac{n}{2} \log n$.
829: 
830: The proof proceeds by induction on $n$. For $n=1$ the claim is trivial. For
831: $n>1$, let $\la v'_1,v'_2,\ldots, v'_n \ra$ be a path of representatives in 
832: $T$.
833: Let $u=\lca_T(v'_1,v'_n)$, $\Delta(u)=d_T(v'_1,v'_n)\geq
834: d_M(v_1,v_n)=n-1$. Let $T_1$ be the subtree of the child of $u$ that 
835: contains
836: $v'_1$. $T_1$ does not contains $v'_n$. Let $i_1<n$ be the maximal $i$ such
837: that $\{v'_1,\ldots,v'_{i_1}\}\subset T_1$. As
838: $i_1+1$ is not contained in $T_1$, it must be that
839: $d_T(v'_{i_1},v'_{i_1+1})\geq \Delta(u)\geq n-1$. By the induction 
840: hypothesis
841: \begin{align*}
842: \ell_T(\la v'_1,\ldots, v'_{i_1} \ra) &\geq g(i_1), & \ell_T(\la
843: v'_{i_1+1},\ldots, v'_{n} \ra) &\geq g(n-i_1).
844: \end{align*}
845: Since $g$ is a convex function, $(g(i_1)+g(n-i_1))/2 \geq g( (i_1
846: +(n-i_1))/2)= g(n/2)$. We conclude
847: \begin{multline*}
848: \ell_T(\la v'_1,\ldots, v'_{n} \ra)  \geq g(i_1)+ (n-1)+ g(n-i_1)
849: \\     \geq 2 g(n/2) + n-1
850:      = 2 \tfrac{n}{4} \log \tfrac{n}{2} +(n-1)
851:      \geq \tfrac{n}{2} \log n .
852: \end{multline*}
853: \end{proof}
854: 
855: \section{Multi-Embedding into Trees} \label{sec:trees}
856: 
857: In this section we consider multi embeddings into \emph{arbitrary
858: tree metrics}. We only have preliminary results. Specifically, we
859: only consider two important types of metric spaces: expander
860: graphs  and the discrete cube with the Hamming metric, for which
861: we obtain better results. For both of them the preceding sections
862: proved an upper bound of $O(\log\log n \log\log\log n)$ and a
863: lower bound of $\Omega(\log \log n)$ on the path distortion of 
864: multi-embeddings
865: into {ultrametrics}  (the lower bound follows since both metrics contain
866: a path of length $\Omega(\log  n)$).
867: 
868: We begin with the observation that for $\Gamma = \infty$ it is
869: easy to obtain $1$ path embedding of any finite metric space into
870: trees. This is achieved by defining an infinite tree
871: metric as follows: joining all possible finite paths with a common
872: root, where the first node in the path is connected with an edge
873: weight of $\Delta/2$ to the root. Moreover,
874: 
875: \begin{proposition} \label{prop:trees}
876: Given a metric space $M$ defined by an unweighted graph of maximum
877: degree $d$ and diameter $\Delta$, and let $s\in\mathbb{N}$. Then $M$ can be
878: $(2+\frac{\Delta}{s})$ path-embedded into a tree metric of size
879: $n d^s$.
880: \end{proposition}
881: \begin{proof}
882: Along the lines of the construction described above, we take all paths of 
883: length
884: $s$, and join these with a common root, where the first node in the path is
885: connected with an edge weight of $\Delta/2$ to the root. Obviously, there 
886: are
887: at most $n d^s$ such paths.
888: Notice
889: that our choice  of weights to the edges adjacent to the root guarantees
890: that distances in
891: the resulting tree are no smaller than the original distances.
892: We next claim that the path distortion is at most
893: $(2+\frac{\Delta}{s})$. To see this,
894: consider a path
895: $p=\langle v_1, \ldots v_\ell \rangle$ of length $\ell$. We partition $p$
896: into sub-paths of length $s$: $p=p_1p_2\cdots p_t$, where $t=\lceil \ell/s
897: \rceil$, $p_j=\langle v_{(j-1)s+1}, \ldots, v_{js} \rangle$ for $j<t$, and
898: $p_t=\langle v_{(t-1)s+1},\ldots, v_\ell \rangle$. Now the sub-path $p_j$ is
899: mapped to the appropriate path in the tree. Note that the length of the 
900: image
901: path is $2\ell+  (t-1)\Delta$.
902: \end{proof}
903: 
904: This simple fact is particularly interesting for its implication
905: for expander graphs. Let $G$ be an $(n,d,\gamma d)$ graph, i.e., a
906: $d$-regular, $n$-vertex graph whose second eigenvalue in absolute
907: value is at most $\gamma d$. It is known \cite{Ch89} that such a
908: graph  has diameter at most $1+\log_{1/\gamma}n$, and so we obtain:
909: \begin{corollary}
910: Any $(n,d,\gamma d)$-graph has $3$ path embedding in a tree of size
911: $d n^{1+\log_{1/\gamma}d}$.
912: \end{corollary}
913: 
914: We also note that for the trees constructed in the proof of
915: Proposition~\ref{prop:trees}, it is particularly easy to obtain
916: a better approximation algorithm  for GST.
917: \begin{lemma} \label{lem:approx-group}
918: Consider a tree metric $M=(V,d)$, where $V=P_1\cup P_2\cup \cdots \cup
919: P_\ell$,  $P_i=\langle v^i_1,v^i_2,\ldots, v_s^i\rangle$ is an unweighted
920: simple path of length $s$, and $d(v^i_1,v^j_1)=\Delta$ for $i\neq j$. Then 
921: an
922: instance of the GST with $k$ groups defined on $M$ has
923: $(1 + \frac{2s}{\Delta})(1+ \ln k)$ approximation algorithm.
924: \end{lemma}
925: \begin{proof}
926: Consider a GST instance $g_1,\ldots g_k$ defined on
927: $M$. We first check whether there is a solution that is completely
928: contained in one $P_i$. This can be checked in polynomial time by
929: noting that if an optimal solution is contained in one $P_i$ then
930: it is an interval. Thus all is needed to be checked are  $\ell
931: \binom{s+1}{2}$ intervals.
932: 
933: Otherwise, the optimal solution intersects $t>1$ of the paths $P_1,\ldots
934: P_\ell$. Define a Hitting Set instance whose ground set is
935: $\{P_1,\ldots,P_\ell\}$ and the subsets are $g'_1,\ldots, g'_k$, where
936: $g'_i=\{P_j;\; P_j\cap g_i \neq \emptyset \}$. It follows that the optimal
937: cost of the hitting set problem is at least $t-1$.
938: The Hitting Set problem has a polynomial time $1+\ln
939: k$ approximation algorithm~\cite{Jo74,L75}. Let $S$ be the approximate
940: solution for the hitting set. We define a solution for the GST
941: instance
942: by taking a natural path over $\cup \{P_i;\; P_i\in
943: S\}$. Note that its length is at most $(\Delta+2s) |S| \le
944: (\Delta+2s)(t-1)(1+\ln k)$. But the cost of the optimal GST
945: algorithm is at least $(t-1)\Delta$.
946: \end{proof}
947: 
948: \begin{corollary}
949: For fixed $d>\lambda$,
950: There exist constants $c=c_{d,\lambda}$, $C=C_{d,\lambda}$, and polynomial
951: $p(t)=p_{d,\lambda}(t)$ such that GST on
952: $(n,d,\lambda)$ graphs has $p(n)$-time $(C\log k)$ approximation algorithm, 
953: and it
954: is NP-hard to approximate within a factor of $c\log k$.
955: \end{corollary}
956: \begin{proof}
957: The approximation algorithm follows from Proposition~\ref{prop:trees} and
958: Lemma~\ref{lem:approx-group}, by setting $s=\Delta$.
959: The hardness result follows since an $(n,d,\lambda)$
960: graph contains a subset of $n^{\Omega_{d,\lambda}(1)}$ points that is 
961: $O_{d,\lambda}(1)$
962: approximated by an equilateral space~\cite{BLMN03}.
963: GST on equilateral space is equivalent to a standard Hitting Set problem,
964: which is NP-hard to
965: approximate within a factor of $c\ln k$ \cite{RS97,LY94}.
966: Usage of points not in this
967: subspace (``Steiner points") can improve the approximation factor by at most
968: a factor of two~\cite{RW90,G01}.
969: \end{proof}
970: 
971: \medskip
972: 
973: We next examine multi-embedding of the $h$-dimensional hypercube with 
974: $n=2^h$
975: vertices. Using Proposition~\ref{prop:trees} with $s=h/\log h$, and using $d
976: = h$ and $\Delta = h$, we obtain $(\log \log n+2)$ path embedding into trees
977: of size
978: $\Gamma(n)\leq n^2$. Using Lemma~\ref{lem:approx-group} it also implies a
979: polynomial time $O(\log k \log \log n)$ approximation algorithm to GST
980: on the cube. Similarly to the expander graphs, it is
981: hard to approximate instances of GST on the cube
982: to within a $c\log k$ factor, for some constant $c>0$, since the cube
983: contains a subset of $n^{\Omega(1)}$ points that is $O(1)$ approximated by 
984: an
985: equilateral space.
986: 
987: 
988: \begin{comment}
989: Our goal in the next proposition is to obtain a better path
990: approximation of the cube by trees. Denote by $\log^{(0)}x =x$,
991: and $\log^{(i+1)}x= \log \log^{(i)}x$. Define $\log^* x=\min\{i\in
992: \mathbb{N}:\; \log^{(i)}x\leq 2\}$.
993: 
994: \begin{proposition} \label{prop:cube}
995: Let $\Omega$ be the $d$-dimensional cube with the Hamming metric,
996: and $n=2^d$. There exists an efficiently constructible multi
997: embedding of $\Omega$ into a tree metric of size
998: $\Gamma(n)\leq \mathrm{poly}(n)$ and path approximation of $O(\log^*
999: n)$.
1000: \end{proposition}
1001: \begin{proof}
1002: The idea of the construction is to have paths in different scales:
1003: short paths with small steps and long paths with large steps. In
1004: this proof we drop all ceiling and floor signs for the sake of
1005: simplicity of presentation. Let $\alpha_0=d$, $\alpha_i= (\log
1006: \alpha_{i-1})^2$, and $s=\min\{i:\; \alpha_i\leq 2\}$. Note that
1007: $s\leq 2\log^*d$. Let $t_i= d/ \alpha_i $. For a point $u$ in the
1008: cube, and $i\in \{1,\ldots,s\}$, let $S_i(u)$ be the union of all
1009: paths that start at $u$, and contain $\frac{2 t_i}{t_{i-1}}$
1010: edges, each edge of length $\frac{1}{2}t_{i-1}$ (\emph{i.e.}, the
1011: total length of each path is roughly $t_i$). The paths are united
1012: at a single vertex corresponding to $u$. Thus $S_i(u)$ can be
1013: thought of all the paths beginning at $u$ \emph{at scale
1014: $(t_i,t_{i-1})$}. We now define a tree containing all the scales
1015: together. $T_0(u)$ is the tree containing a single vertex
1016: (corresponding to $u$). $T_i(u)$ for $i\geq 1$ is the tree
1017: resulting by starting with $S_i(u)$ and for each vertex $x$ in
1018: $S_i(u)$, Let $v=\f(x)$ ($v\in\Omega$ is the original point in the
1019: cube) we attach a copy of $T_{i-1}(v)$ by identifying the root of
1020: this copy with $x$.
1021: 
1022: The final tree, $T$, is constructed by connecting all the trees
1023: $T_s(u)$, $u\in \Omega$, to a single new vertex $v_0$, with edges
1024: of length $d/2$.
1025: 
1026: We begin with estimating the blow-up factor. From symmetry
1027: $|S_i(u)|=|S_i(v)|$ for any $u,v\in\Omega$, and by a simple
1028: induction argument, $|T_i(u)|=\prod_{j=1}^i |S_j(u)|$. Therefore,
1029: $|T_s(u)|=\prod_{i=1}^s |S_i(u)|$. To estimate $|S_i(u)|$, note
1030: that there are $\binom{d}{0.5 t_{i-1}}$ points at distance $0.5
1031: t_{i-1}$ from a given point. Using the fact that $\binom{a}{b}\leq
1032: (\frac{ea}{b})^b$, we obtain
1033: \begin{equation*}
1034: |S_i(u)|\leq  {\binom{d}{\tfrac{1}{2} t_{i-1}}}^\frac{2 t_i}{t_{i-1}}\leq
1035: \left(\frac{2e d}{t_{i-1}}\right)^{t_i}
1036: \leq  e^{d \frac{\ln
1037: (2e\alpha_{i-1})}{\alpha_i}}.
1038: \end{equation*}
1039: Therefore,
1040: \begin{equation*}
1041:   |T_s(u)| \leq \prod_{i=1}^s |S_i(u)|\leq
1042: \exp\Bigl( d \bigl( \sum_{i=1}^s \tfrac{\ln(2e)+\ln \alpha_{i-1}}{\alpha_i}
1043: \bigr  ) \Bigr )
1044: \leq
1045: \exp \Bigl( d \bigl( \sum_{i=1}^s \tfrac{3}{\sqrt{\alpha_i}} \bigr
1046: ) \Bigr ) \leq dn^{O(1)}.
1047: \end{equation*}
1048: So $|T|=1+n|T_s(u)|=n^{O(1)}$.
1049: 
1050: We next estimate the path approximation of the cube by $T$. Let
1051: $p=\langle v_1, v_2,\ldots \rangle$ be a path in the cube. We
1052: define $\hat{p}^s$, the \emph{$s$-level skeleton of $p$} as a
1053: (non-contiguous) subsequence of $p$ defined as follows. The first
1054: vertex is $v_1$. The second vertex is $v_{i_1}$ such that $i_1$ is
1055: the smallest $i$ of distance $\frac{1}{2}t_{s-1}$ from $v_1$. The
1056: $j$ vertex $v_{i_j}$ is the smallest $i$ larger than $i_{j-1}$
1057: such that $v_{i_j}$ is a distance $\frac{1}{2}t_{s-1}$ from
1058: $v_{i_{j-1}}$. We define a path of representatives of  $\hat{p}^s$
1059: in $\cup_u S_s(u)$ in the same way as in
1060: Proposition~\ref{prop:trees}. Since $t_s=\Omega(d)$, this is an
1061: $O(1)$ path approximation of $\hat{p}^s$.
1062: 
1063: Notice that $\hat{p}^s$ induces a natural partition of $p$ into
1064: (contiguous) sub-paths. Consider such a sub-path $q=\langle
1065: w=v_{i_j},v_{i_j+1},\ldots,v_{i_{j+1}-1}\rangle$, and repeat the
1066: same process at level $s-1$, i.e., Let $\hat{q}^{s-1}$ be the
1067: $(s-1)$-level skeleton of $q$, obtained by taking $w$ and
1068: inductively the first point at distance $\frac{1}{2}t_{s-2}$ from
1069: the previous point in the skeleton. Again, $\hat{q}^{s-1}$ is
1070: $O(1)$ path approximated by a path in $S_{s-1}(w)$. This is
1071: possible since all the points in $\hat{q}$ are at distance at most
1072: $\frac{1}{2}t_{s-1}$ from $w$, while the paths in $S_{s-1}(w)$ are
1073: of length $t_{s-1}$. Continue this in all levels down to level
1074: $1$. The concatenation of all paths from all levels is a path of
1075: representatives of $p$ in $T$. The construction of the paths was
1076: done is such a way their concatenation increase the total distance
1077: of the path of duplicates by at most a factor of $2$. Each level
1078: has a total length which is $O(1)$ times the length of $p$.
1079: Therefore we obtain $O(s)$ path approximation.
1080: \end{proof}
1081: 
1082: We stress that Proposition~\ref{prop:cube} does not immediately yield an
1083: improved bound on
1084: the approximation ratio of GST on the cube, since we do not
1085: know of an algorithm better than the $O(\log k \log n)$ approximation 
1086: algorithm
1087: of \cite{GKR98} for the resulting trees.
1088: \end{comment}
1089: 
1090: \section{Discussion} \label{sec:conclusion}
1091: 
1092: An interesting open problem is to determine worst case bounds for
1093: path distortion of multi embedding into trees of polynomial size.
1094: As indicated by the case of expander graphs, such bounds may be
1095: better than those for {ultrametrics}. 
1096: 
1097: Results on multi-embedding into trees directly reflect on the
1098: approximability of GST. As shown for expanders and hypercubes, it
1099: is possible that for special classes of metric spaces, a
1100: combination of improved path embedding and a specialized solution
1101: would yield (nearly) tight upper bounds. Our approach to show the
1102: (near) tightness of the results in those cases stems from metric
1103: Ramsey-type considerations (i.e. the existence of large
1104: approximately equilateral subspace). Such considerations are in
1105: fact more general and may lead to more results of this flavor.
1106: \footnote{In \cite{BLMN03} it is shown that any metric space
1107: contains a ``large" subspace which is approximately an
1108: ultrametric, or a $k$-HST. Such trees were used in \cite{HK03} to
1109: prove inapproximability results for GST. It is plausible that
1110: these techniques can be combined to obtain tight bounds for GST in
1111: specific metric spaces.} 
1112: 
1113: Multi-embedding into {ultrametrics}, also implies multi embedding into
1114: $\ell^d_p$, where $d=O(\log n)$ with similar
1115: path distortion~\cite{blmn-lowdim}. It is natural to ask whether better
1116: path distortion is possible for multi-embedding into $\ell_1$ or $\ell_2$.
1117: Further study of multi-embeddings in other settings and their applications
1118: seems an attractive direction for future research.
1119: 
1120: \subsubsection*{Acknowledgments} We thank Robi Krau\-th\-gamer
1121: and Assaf Naor for fruitful discussions.
1122: 
1123: \bibliographystyle{siam}
1124: \bibliography{path}
1125: \end{document}
1126: 
1127: