This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Reliable Spanners for Metric Spacesthanks: A preliminary version of this paper appeared in SoCG 2021 [HMO21].

Sariel Har-Peled Department of Computer Science; University of Illinois; 201 N. Goodwin Avenue; Urbana, IL, 61801, USA; [email protected]; http://sarielhp.org/. Work on this paper was partially supported by a NSF AF award CCF-1907400.    Manor Mendel Department of Mathematics and Computer Science, The Open University of Israel, [email protected]. Supported by BSF Grant no. 2018223.    Dániel Oláh Department of Mathematics and Computing Science, TU Eindhoven, P.O. Box 513, 5600 MB Eindhoven, The Netherlands.
Abstract

A spanner is reliable if it can withstand large, catastrophic failures in the network. More precisely, any failure of some nodes can only cause a small damage in the remaining graph in terms of the dilation. In other words, the spanner property is maintained for almost all nodes in the residual graph. Constructions of reliable spanners of near linear size are known in the low-dimensional Euclidean settings. Here, we present new constructions of reliable spanners for planar graphs, trees and (general) metric spaces.

1 Introduction

Let =(P,d){\mathcal{M}}=({P},{\mathrm{d}}) be a finite metric space. Let G=(P,E){G}=({P},{E}) be a sparse graph on the points of {\mathcal{M}} whose edges are weighted with the distances of their endpoints. The graph G{G} is a tt-spanner if for any pair of vertices u,vPu,v\in{P} we have dG(u,v)td(u,v){\mathrm{d}}_{G}\left({u,v}\right)\leq t\cdot{\mathrm{d}}\left({u,v}\right), where dG(u,v){\mathrm{d}}_{G}\left({u,v}\right) is the length of the shortest path between uu and vv in G{G}, and d(u,v){\mathrm{d}}\left({u,v}\right) is the distance in the metric space between uu and vv. Spanners were first introduced by Peleg and Schäffer [PS89] as a tool in distributed computing, but have since found use in many other areas, such as algorithms, networking, data structures, and metric geometry, see [Pel00, NS07].

Fault tolerant spanners

A desired property of spanners is the ability to withstand failures of some of their vertices. One such notion is provided by fault tolerance [CLNS15, LNS98, LNS02, Luk99, Sol14]. A graph G{G} is a kk-fault tolerant tt-spanner, if for any subset of vertices B{{B}}, with |B|k\left|{{{B}}}\right|\leq k, the graph GB{G}\setminus{{B}} is a tt-spanner. However, for kk-fault tolerant graphs there is no guarantee if more than kk vertices fail, and furthermore, the size of a fault tolerant graph grows (linearly) with the parameter kk. In particular, for fixed t1t\geq 1, the optimal size of kk-fault tolerant (2t1)(2t-1)-spanners on nn vertices is 𝒪(k11/tn1+1/t){\mathcal{O}}(k^{1-1/t}n^{1+1/t}) [BDPW18]. Note that vertex degrees must be at least Ω(k)\Omega(k) to avoid the possibility of isolating a vertex. Thus, it is not suitable for massive failures in the network.

Reliable spanners

An alternative notion was introduced by Bose et al. [BDMS13]. Here, for a parameter ϑ>0{\vartheta}>0, a ϑ{\vartheta}-reliable tt-spanner has the property that for any failure (or attack) set B{{B}}, the residual graph GB{G}\setminus{{B}} is a tt-spanner path for the points of VB+{V}\setminus{{B^{+}}}, where B+B{{B^{+}}}\supseteq{{B}} is some set, such that |B+|(1+ϑ)|B|\left|{{{B^{+}}}}\right|\leq(1+{\vartheta})\left|{{{B}}}\right|. We consider two variants:

  1. (A)

    Adaptive adversary (i.e., standard or “deterministic” model): The adversary knows the spanner G{G}, and the set B{{B}} is chosen as a worst case for G{G}.

  2. (B)

    Oblivious adversary (i.e., “randomized” model): Here, the spanner G{G} is drawn from a probability distribution χ\chi (over the same number of vertices). The adversary knows χ\chi in advance, but not the sampled spanner. In this oblivious model, we require that 𝔼[|B+|](1+ϑ)|B|\mathop{\mathbb{E}}\!\left[{|{{B^{+}}}|}\right]\leq(1+{\vartheta})|{{B}}|.

See Section 2 for precise definitions.

Previous work

Bose et al. [BDMS13] provided some lower bounds and constructions in the general settings, but the bounds on the size of the damaged set (i.e., B+{{B^{+}}}) are much larger. In the Euclidean settings, for any point set Pd{P}\subseteq\mathbb{R}^{d}, and for any constants ϑ,ε(0,1){\vartheta},\varepsilon\in(0,1), one can construct a ϑ{\vartheta}-reliable (1+ε)(1+\varepsilon)-spanner with only 𝒪(nlognloglog6n){\mathcal{O}}\bigl{(}n\log n\log\!\log^{6}n\bigr{)} edges [BHO19] (an alternative, but slightly inferior construction, was provided independently by Bose et al. [BCDM18]). The number of edges can be further reduced in the oblivious adversary case, where one can construct an oblivious (1+ε)(1+\varepsilon)-spanner that is ϑ{\vartheta}-reliable in expectation and has 𝒪(nloglog2nlogloglogn){\mathcal{O}}(n\log\!\log^{2}n\log\!\log\!\log n) edges [BHO20].

Covers.

A cover is a set of clusters (i.e., subsets of the point set) that covers the metric space with certain desirable properties. Awerbuch and Peleg [AP90] showed a cover, where (i) each cluster has diameter O(k)O(k), (ii) every vertex participates in O(kn1/k)O(kn^{1/k}) clusters, and (iii) for every vertex vv, the ball of radius kk centered at vv, is contained in a single cluster. Busch et al. [BLT14] show how to compute a cover of a planar graph, with diameter 16k\leq 16k per cluster, such that each pair in distance kk from each other belongs to some cluster, and every vertex participates in at most 1818 clusters. For graphs that excludes a minor of fixed size, they get a similar result, except that each vertex might participate in O(logn)O(\log n) vertices, where nn is the number of vertices in the input graph. Abraham et al. [AGMW10] presented a result with better sparsity when the graphs do not have Kr,rK_{r,r} as a minor.

Constructions of ϑ{\vartheta}-reliable Δ\Delta-spanners
metric Δ\Delta guarantee size ref
Uniform 22 expectation 𝒪(nϑ1logϑ1).{\mathcal{O}}(n{\vartheta}^{-1}\log{\vartheta}^{-1})\Bigr{.} Lemma 3.1
tt det. lower bound Ω(n1+1/t)\Omega(n^{1+{1}/{t}}) Lemma 3.2
2t12t-1 deterministic 𝒪(ϑ2n1+1/t).{\mathcal{O}}\left({{\vartheta}^{-2}n^{1+1/t}}\right)\Bigr{.} Theorem 3.6
2t2t deterministic 𝒪(ϑ1n1+1/t).{\mathcal{O}}\left({{\vartheta}^{-1}n^{1+1/t}}\right)\Bigr{.} Theorem 3.6
Finite metrics 𝒪(logn){\mathcal{O}}(\log n) expectation 𝒪(ϑ1npolylog){\mathcal{O}}\left({{\vartheta}^{-1}n\mathop{\mathrm{polylog}}}\right) Lemma 5.3
𝒪(t){\mathcal{O}}(t) expectation 𝒪(ϑ1n1+1/tpolylog){\mathcal{O}}\left({{\vartheta}^{-1}n^{1+1/t}\mathop{\mathrm{polylog}}}\right) Lemma 5.3
𝒪(tlogn){\mathcal{O}}(t\log n) deterministic 𝒪(ϑ1n1+1/tpolylog){\mathcal{O}}\left({{\vartheta}^{-1}n^{1+1/t}\mathop{\mathrm{polylog}}}\right) Lemma 5.4
𝒪(t2){\mathcal{O}}(t^{2}) deterministic 𝒪(ϑ1n1+1/tpolylog){\mathcal{O}}\left({{\vartheta}^{-1}n^{1+1/t}\mathop{\mathrm{polylog}}}\right) Lemma 5.4
Ultrametrics 2+ε2+\varepsilon expectation 𝒪(ϑ1ε2npolylog).{\mathcal{O}}\left({{\vartheta}^{-1}\varepsilon^{-2}n\mathop{\mathrm{polylog}}}\right)\Bigr{.} Lemma 5.5
(2+ε)t1(2+\varepsilon)t-1 deterministic 𝒪(ϑ2ε3n1+1/tpolylog).{\mathcal{O}}\left({{\vartheta}^{-2}\varepsilon^{-3}n^{1+1/t}\mathop{\mathrm{polylog}}}\right)\Bigr{.} Lemma 5.6
Trees 3+ε3+\varepsilon expectation 𝒪(ϑ1ε2npolylog).{\mathcal{O}}\left({{\vartheta}^{-1}\varepsilon^{-2}n\mathop{\mathrm{polylog}}}\right)\Bigr{.} Lemma 5.7
(4+ε)t3(4+\varepsilon)t-3 deterministic 𝒪(ϑ2ε3n1+1/tpolylog).{\mathcal{O}}\left({{\vartheta}^{-2}\varepsilon^{-3}n^{1+1/t}\mathop{\mathrm{polylog}}}\right)\Bigr{.} Lemma 5.8
Planar graphs 3+ε3+\varepsilon expectation 𝒪(ϑ1ε4npolylog).{\mathcal{O}}\left({{\vartheta}^{-1}\varepsilon^{-4}n\mathop{\mathrm{polylog}}}\right)\Bigr{.} Lemma 5.9
(4+ε)t3(4+\varepsilon)t-3 deterministic 𝒪(ϑ2ε6n1+1/tpolylog){\mathcal{O}}({\vartheta}^{-2}\varepsilon^{-6}n^{1+1/t}\mathop{\mathrm{polylog}}) Lemma 5.10
Table 1.1: Our results. Polylog factors are polynomial factors in logn\log n and logΦ\log{\Phi}, where Φ{\Phi} is the spread of the metric. For trees and planar graphs, these results are for graphs with weights on the edges. Here in expectation denotes that the spanner works against an oblivious adversary (here, the expectation is over the randomization in the construction), and the guarantee is on the expected size of the damaged set. Similarly, deterministic implies an adaptive adversary.

Our results.

We provide new constructions of reliable spanners for finite uniform metrics, ultrametrics, trees, planar graphs and finite metrics. Our new results are summarized in Table 1.1.

Technique

Our approach for constructing reliable spanners is in two steps: We first construct reliable spanners for uniform metrics and then reduce the problem of constructing reliable spanners for general metrics to uniform metrics using covers.

Spanners for uniform metrics

Uniform metrics have trivial classical 2-spanners – that is, star graphs. In the oblivious model one can simply use “constellation of stars” with a constant number of random centers, and the resulting spanner is linear in size. In the adaptive settings, we present a lower bound of Ω(n1+1/t)\Omega(n^{1+1/t}) edges for a reliable tt-spanner, and asymptotically “matching” construction of a deterministic (2t1)(2t-1)-reliable spanner with 𝒪(n1+1/t){\mathcal{O}}(n^{1+1/t}) edges. The construction is based on reliable expanders – these are expanders that remain expanding under the type of attacks described above. See Section 3 and Section 6 for details.

Covers

A tt-cover of a finite metric space =(P,𝖽){\mathcal{M}}=({P},{\mathsf{d}}) is a family of subsets 𝒞={S|SP}{\mathcal{C}}=\left\{{S}\;\middle|\;{S}\subseteq{P}\right\}, such that for each pair p,qP{p},{q}\in{P} of points there exists a subset in 𝒞{\mathcal{C}} that contains both points and whose diameter is at most td(p,q)t\cdot{\mathrm{d}}\left({{p},{q}}\right). Covers are used here to extend reliable spanners for uniform metrics into reliable spanners for general metrics. This is done by using spanners for uniform in each set of the cover and then taking a union of the edges of those graphs. See Section 5.

Naturally, the size S𝒞|S|\sum_{{S}\in{\mathcal{C}}}\left|{{S}}\right| of a tt-cover 𝒞{\mathcal{C}} is an important parameter in the resulting size of the spanner, so in Section 4 we study the problem of constructing good covers. For general nn-point spaces with spread at most Φ{\Phi}, we observe that the Ramsey partitions of [MN07] provide 𝒪(t){\mathcal{O}}(t) covers of size 𝒪(n1+1/tlogΦ){\mathcal{O}}(n^{1+1/t}\log{\Phi}), which is close to optimal, because of an Ω(n(n1/t+logtΦ))\Omega(n(n^{1/t}+\log_{t}{\Phi})) lower bound we provide. In more specific cases, like ultrametrics, trees and planar graphs, one can do better. For example, for trees and planar graphs one gets (2+ε)(2+\varepsilon)-covers of near linear size. For planar graphs, known partitions have much larger gap, which makes these results quite interesting.

New reliable spanners

Plugging the constructions of spanners for uniform metrics with the construction of covers yields reliable spanners for finite uniform, ultrametric, tree, planar, and general metrics. The results are summarized in Table 1.1.

Efficient construction

All our constructions relies on randomized constructions of expanders (over mm vertices), that succeeds with probability 11/mO(1)\geq 1-1/m^{O(1)}. As such, the constructions described can be done efficiently, if one wants constructions of spanners for nn vertices that succeeds with probability 11/nO(1)1-1/n^{O(1)}. See Remark 6.3 for details.

2 Preliminaries

2.1 Metric spaces

For a set 𝒳\mathcal{X}, a function d:𝒳2[0,){\mathrm{d}}:\mathcal{X}^{2}\rightarrow[0,\infty), is a metric if it is symmetric, complies with the triangle inequality, and d(p,q)=0{\mathrm{d}}\left({{p},{q}}\right)=0 \iff p=q{p}={q}. A metric space is a pair =(𝒳,d){\mathcal{M}}=(\mathcal{X},{\mathrm{d}}), where d{\mathrm{d}} is a metric. For a point p𝒳{p}\in\mathcal{X}, and a radius rr, the ball of radius rr is the set

𝖻(p,r)={q𝒳|d(p,q)r}.{\mathsf{b}}\left({{p},r}\right)=\left\{{q}\in\mathcal{X}\;\middle|\;{\mathrm{d}}\left({{p},{q}}\right)\leq r\right\}.

For a finite set X𝒳X\subseteq\mathcal{X}, the diameter of XX is

diam(X)=diam(X)=maxp,qXd(p,q),\mathrm{diam}\left({X}\right)={\mathrm{diam}}_{{\mathcal{M}}}\left({X}\right)=\max_{{p},{q}\in X}{\mathrm{d}}\left({{p},{q}}\right),

and the spread of XX is Φ(X)=diam(X)minp,qX,pqd(p,q).{{\Phi}}\left({X}\right)=\frac{\mathrm{diam}\left({X}\right)}{\min_{{p},{q}\in X,{p}\neq{q}}{\mathrm{d}}\left({{p},{q}}\right)}. A metric space =(𝒳,d){\mathcal{M}}=(\mathcal{X},{\mathrm{d}}) is finite, if 𝒳\mathcal{X} is a finite set. In this case, we use Φ=Φ(𝒳){\Phi}={{\Phi}}\left({\mathcal{X}}\right) to denote the spread of the (finite) metric.

A natural way to define a metric space is to consider an undirected connected graph G=(P,E){G}=({P},{E}) with positive weights on the edges. The shortest path metric of G{G}, denoted by dG{\mathrm{d}}_{G}, assigns for any two points p,qP{p},{q}\in{P} the length of the shortest path between p{p} and q{q} in the graph. Thus, any graph G{G} readily induces the finite metric space (V(G),dG)({V}\left({{G}}\right),{\mathrm{d}}_{G}). If the graph is unweighted, then all the edges have weight 11.

A tree metric is a shortest path metric defined over a graph that is a tree.

2.2 Reliable spanners

Definition 2.1.

For a metric space =(P,d){\mathcal{M}}=({P},{\mathrm{d}}), a graph H=(P,E){H}=({P},{E}) is a tt-spanner, if for any p,qP{p},{q}\in{P}, d(p,q)dH(p,q)td(p,q).{\mathrm{d}}\left({{p},{q}}\right)\leq{\mathrm{d}}_{{H}}\left({{p},{q}}\right)\leq t\cdot{\mathrm{d}}\left({{p},{q}}\right). Here dH{\mathrm{d}}_{H} is the shortest path distance on HH whose edges are weighted according to d{\mathrm{d}}.

Given a weighted graph G=(V,E){G}=({V},{E}), and a set BV{{B}}\subseteq{V}, we denote by G|B{G}|_{{B}} the subgraph induced on B{{B}}. We also use the notation GB=G|VB{G}\setminus{{B}}={G}|_{{V}\setminus{{B}}}. A randomized graph G{G} is a probability distribution over the edge set E{E} for a given set of vertices V{V}.

An attack on a graph G=(V,E){G}=({V},{E}) is a set of vertices B{{B}} that fails and no longer can be used. An attack (on a randomized graph) is oblivious, if the set B{{B}} is picked stochastically independent of the edge set of the graph.

Definition 2.2 (Reliable spanner).

Let G=(P,E){G}=({P},{E}) be a tt-spanner for some t1t\geq 1 constructed by a (possibly) randomized algorithm. Given an attack B{{B}}, its damaged set B+{{B^{+}}} is a set of smallest possible size, such that for any pair of vertices p,qPB+{p},{q}\in{P}\setminus{{B^{+}}}, we have

dGB(p,q)td(p,q),{\mathrm{d}}_{{G}\setminus{{B}}}\left({{p},{q}}\right)\leq t\cdot{\mathrm{d}}\left({{p},{q}}\right), (2.1)

that is, distances are preserved (up to a factor of tt) for all pairs of points not contained in B+{{B^{+}}}. The quantity |B+B|\left|{{{B^{+}}}\setminus{{B}}}\right| is the loss of G{G} under the attack B{{B}}. The loss rate of G{G} is λ(G,B)=|B+B|/|B|{\lambda}\left({{G},{{B}}}\right)=\left|{{{B^{+}}}\setminus{{B}}}\right|/\left|{{{B}}}\right|. For ϑ(0,1){\vartheta}\in(0,1), the graph G{G} is ϑ{\vartheta}-reliable (in the deterministic or non-oblivious setting), if λ(G,B)ϑ{\lambda}\left({{G},{{B}}}\right)\leq{\vartheta} holds for any attack BP{{B}}\subseteq{P}. Furthermore, the graph G{G} is ϑ{\vartheta}-reliable in expectation (or in the oblivious model), if 𝔼[λ(G,B)]ϑ\mathop{\mathbb{E}}\!\left[{{\lambda}\left({{G},{{B}}}\right)}\right]\leq{\vartheta} holds for any oblivious attack BP{{B}}\subseteq{P}.

Remark 2.3.

The damaged set B+{{B^{+}}} is not necessarily unique, since there might be freedom in choosing the point to include in B+{{B^{+}}} for a pair that does not have a tt-path in GB{G}\setminus{{B}}.

2.3 Miscellaneous

For a graph G{G}, and two set of vertices Y,ZV(G)Y,Z\subseteq{V}\left({{G}}\right), let

ΓZ(Y)={xZ|xyE(G) and yY}{\Gamma}_{Z}\left({Y}\right)=\left\{x\in Z\;\middle|\;xy\in{E}\left({{G}}\right)\text{ and }y\in Y\right\}

denote the neighbors of YY in ZZ. The neighbors of YY in G{G} is denoted by Γ(Y)=ΓV(G)(Y){\Gamma}\left({Y}\right)={\Gamma}_{{V}\left({{G}}\right)}\left({Y}\right).

Definition 2.4.

For a collection of sets {\mathcal{F}}, and an element xx, let deg(x,)=|{X|xX}|\mathrm{deg}\left({x,{\mathcal{F}}}\right)=\left|{\left\{X\in{\mathcal{F}}\;\middle|\;x\in X\right\}}\right| denote the degree of xx in {\mathcal{F}}. The maximum degree of any element of {\mathcal{F}} is the depth of {\mathcal{F}}.

Notations

We use P+p=P{p}{P}+{p}={P}\cup\{{p}\} and Pp=P{p}{P}-{p}={P}\setminus\{{p}\}. Similarly, for a graph G{G}, and a vertex p{p}, let Gp{G}-{p} denote the graph resulting from removing p{p}.

3 Reliable spanners for uniform metric

Let P{P} be a set of nn points and let (P,d)\left({{P},{{{\mathrm{d}}}}}\right) be a metric space equipped with the uniform metric, that is, for all distinct pairs p,qP{p},{q}\in{P}, we have that d(p,q){\mathrm{d}}\left({{p},{q}}\right) is the same quantity (e.g., 11). Note that n1n-1 edges are enough to achieve a 22-spanner for the uniform metric by using the star graph.

3.1 A randomized construction for the oblivious case

Construction

Let ϑ(0,1){\vartheta}\in(0,1) be a fixed parameter. Set k=2ϑ1logϑ1+1k=2\left\lceil{{\vartheta}^{-1}\log{\vartheta}^{-1}}\right\rceil+1 and sample kk points from P{P} uniformly at random (with replacement). Let CP{C}\subseteq{P} be the resulting set of center points. For each point pC{p}\in{C}, build the star graph =p(P,{pq|qPp}){}_{{p}}=({P},\left\{{p}{q}\;\middle|\;{q}\in{P}-{p}\right\}), where p{p} is the center of the star. The constellation of C{C} is the graph =pCp{\Asterisk}=\bigcup_{{p}\in{C}}{}_{{p}}, which is the union of the star graphs induced by centers in C{C}.

Lemma 3.1.

The constellation , defined above, is a ϑ{\vartheta}-reliable 22-spanner in expectation. The number of its edges is 𝒪(nϑ1logϑ1){\mathcal{O}}(n{\vartheta}^{-1}\log{\vartheta}^{-1}).

Proof:

Let BP{{B}}\subseteq{P} be an oblivious attack, and let b=|B|b=\left|{{{B}}}\right|. If there is a point of C{C} that is not in B{{B}}, then there is a center point outside of the attack set, which provides 22-hop paths between all pairs of points in the residual graph, and therefore we choose B+=B{{B^{+}}}={{B}}. On the other hand, if CB{C}\subseteq{{B}}, then the residual graph contains only isolated vertices, and therefore we choose B+=P{{B^{+}}}={P}. If (1+ϑ)bn(1+{\vartheta})b\geq n then there is nothing to prove. Thus, since b/n<1/(1+ϑ)b/n<1/(1+{\vartheta}) and 1/(1+ϑ)1ϑ/21/(1+{\vartheta})\leq 1-{\vartheta}/2, we have

𝔼[λ(G,B)]\displaystyle\mathop{\mathbb{E}}\!\left[{{\lambda}\left({{G},{{B}}}\right)}\right] =0[CB]+n1bb[CB]nb(bn)k1(1+ϑ)k1exp(k12ϑ)ϑ.\displaystyle=0\mathop{{\mathbb{P}}}\left[{C}\nsubseteq{{B}}\right]+\frac{n-1-b}{b}\mathop{{\mathbb{P}}}\left[{C}\subseteq{{B}}\right]\leq\frac{n}{b}\left({\frac{b}{n}}\right)^{k}\leq\frac{1}{(1+{\vartheta})^{k-1}}\leq\exp\Bigl{(}-\frac{k-1}{2}{\vartheta}\Bigr{)}\leq{\vartheta}.

As for the number of edges, has at most k(n1)k(n-1) edges, since is the union of kk stars and each star has n1n-1 edges. Thus, by the choice of kk, the size of is 𝒪(nϑ1logϑ1){\mathcal{O}}(n{\vartheta}^{-1}\log{\vartheta}^{-1}).  

3.2 Lower bound for a deterministic construction

In the non-oblivious settings, the attacker knows the constructed graph G{G} when choosing the attack set B{{B}}.

Lemma 3.2.

Let G=(P,E){G}=(P,E) be a ϑ{\vartheta}-reliable tt-spanner on P{P} for the uniform metric, where ϑ(0,1){\vartheta}\in(0,1) and t1t\geq 1. Then, in the non-oblivious settings, G{G} must have Ω(n1+1/t)\Omega(n^{1+{1}/{t}}) edges.

Proof:

We assume that the distance between any pair of points of PP is one. Let the attack set B{{B}} be the set of all nodes of degree at least Δ\Delta, where Δ=n1/t/4\Delta=n^{1/t}/4. Assume, toward a contradiction, that |B|n/4\left|{{{B}}}\right|\leq n/4. By the reliability condition, there exists a set QPB{Q}\subseteq{P}\setminus{{B}} of nodes of the residual graph of size at least n(1+ϑ)|B|n/2n-(1+{\vartheta})\left|{{{B}}}\right|\geq n/2, that has tt-hop paths in GB{G}\setminus{{B}} between all pairs of vertices of Q{Q}. Let pQ{p}\in{Q} be an arbitrary vertex. Let 𝖻(p,t){\mathsf{b}}\left({{p},t}\right) be a ball of radius tt centered at p{p} in the shortest path metric in the residual graph GB{G}\setminus{{B}}. On the one hand, from the above, 𝖻(p,t)Q{\mathsf{b}}\left({{p},t}\right)\supseteq Q. On the other hand, the maximum degree is at most Δ\Delta, and therefore |𝖻(p,t)|Δtn/4\left|{{\mathsf{b}}\left({{p},t}\right)}\right|\leq\Delta^{t}\leq n/4. As such, we have n/4|𝖻(p,t)||Q|n/2n/4\geq\left|{{\mathsf{b}}\left({{p},t}\right)}\right|\geq\left|{{Q}}\right|\geq n/2. A contradiction.

Hence, |B|>n/4\left|{{{B}}}\right|>n/4. The claim follows since Δ|B|2|E|\Delta|{{B}}|\leq 2|{E}|, which implies |E|Δ|B|/2=Ω(n1+1/t).\left|{{E}}\right|\geq\Delta\left|{{{B}}}\right|/2=\Omega(n^{1+1/t}).  

Remark 3.3.

Erdős’ girth conjecture states that there exists a graph G{G} with nn vertices and Ω(n1+1/k)\Omega(n^{1+1/k}) edges, and girth at least 2k+12k+1, where the girth of G{G} is the length of the shortest cycle in G{G}. The argument in the proof of Lemma 3.2 is closely related to the standard argument for proving a tight counterpart – any graph with Ω(n1+1/k)\Omega(n^{1+1/k}) edges has girth at most 2k+12k+1.

3.3 Reliable spanners of the uniform metric for adaptive adversary

Here, we present a construction of reliable spanner that is close to being tight. The spanner is simply a high-degree expander whose properties are described in the following definition.

Definition 3.4.

Denote by λ(G)\lambda({G}) the second eigenvalue of the matrix M/d{M}/d, where M=Adj(G){M}=\mathrm{Adj}({G}) is the adjacency matrix of a dd-regular graph G{G}. A proper expander specifies a constant C>1C>1, and functions Cδ,cδ>0{\mathcalb{C}}_{\!\delta},{\mathcalb{c}}_{\!\delta}>0, such that for every δ(0,1/4)\delta\in(0,1/4) and even integers dCδd\geq{\mathcalb{C}}_{\!\delta}, nd2n\geq d^{2}, there exists an nn-vertex, dd-regular graph G=(V,E){G}=({V},{E}), such that:

  1. (P1)

    SV,|S|12n/(δd)|Γ(S)|>(1δ)n,\forall S\subseteq{V},\ |S|\geq{12n}/(\delta d)\;\implies\;|{\Gamma}\left({S}\right)|>(1-\delta)n,

  2. (P2)

    SV,|S|cδn/d|Γ(S)|(1δ)d|S|.\forall S\subseteq{V},\ |S|\leq{\mathcalb{c}}_{\!\delta}n/d\;\implies\;|{\Gamma}\left({S}\right)|\geq(1-\delta)d|S|.

  3. (P3)

    λ(G)C/d.\lambda({G})\leq{C}/{\sqrt{d}}.

For each one of the properties above, it is known that there exists an expander satisfying it: Property (P1) is essentially proved in [BHO19], Property (P2) is folklore, and Property (P3) appears in [DJPP13]. Since they hold almost surely for “random regular graphs”, they also hold simultaneously. However, we were unable to find in the literature proofs of almost sure existence in the same model of random regular graphs, and contiguity of the different random models does not necessarily hold in the high-degree regime (which is what we need here). Therefore, for completeness, Appendix A reprove (P1) and (P2) in the same random model in which (P3) was proved. We thus get the following:

Theorem 3.5.

The random graph constructed in Section A.1 is a proper expander (see Definition 3.4), asymptotically almost surely. Specifically, the probability the graph has the desired properties is 1nO(1)\geq 1-n^{-O(1)}.

With the appropriate choice of parameters, these expanders are reliable spanners for uniform metrics.

Theorem 3.6.

For every tt\in\mathbb{N}, θ(0,1)\theta\in(0,1), and n2n\in 2\mathbb{N}, such that neΩ(t)n\geq e^{\Omega(t)}, there exist:

  1. (i)

    ϑ{\vartheta}-reliable 2t2t-spanner with 𝒪(ϑ1n1+1/t){\mathcal{O}}{({\vartheta}^{-1}n^{1+1/t})} edges for nn-point uniform space, and

  2. (ii)

    ϑ{\vartheta}-reliable (2t1)(2t-1)-spanner with 𝒪(ϑ2n1+1/t){\mathcal{O}}{({\vartheta}^{-2}n^{1+1/t})} edges for nn-point uniform space.

The proof is somewhat cumbersome and is deferred to Section 6.

Applying the above theorem directly on non-uniform metric we obtain the following corollaries.

Corollary 3.7.

Let =(𝒳,d){\mathcal{M}}=(\mathcal{X},{\mathrm{d}}) be a metric space, and let P𝒳{P}\subseteq\mathcal{X} be a finite subset of size nn. Given parameters tt\in\mathbb{N}, and ϑ(0,1){\vartheta}\in(0,1), there exists a weighted graph G{G} on P{P}, such that:

  1. (A)

    The graph G{G} has |E(G)|=𝒪(ϑ2n1+1/t)\left|{{E}\left({{G}}\right)}\right|={\mathcal{O}}\left({{\vartheta}^{-2}\cdot n^{1+1/t}}\right) edges.

  2. (B)

    The graph G{G} is ϑ{\vartheta}-reliable. Namely, given any attack set B𝒳{{B}}\subset\mathcal{X}, there exists a subset QP{Q}\subseteq{P}, such that |Q||P|(1+ϑ)|BP|\left|{{Q}}\right|\geq\left|{{P}}\right|-(1+{\vartheta})\left|{{{B}}\cap{P}}\right|. Furthermore, for any two points p,qQ{p},{q}\in{Q}, we have

    d(p,q)dG|Q(p,q)(2t1)diam(P),{\mathrm{d}}_{{\mathcal{M}}}\left({{p},{q}}\right)\leq{\mathrm{d}}_{{G}|_{Q}}\left({{p},{q}}\right)\leq(2t-1)\cdot{\mathrm{diam}}_{{\mathcal{M}}}\left({{P}}\right),

    and the path realizing it has at most (2t1)(2t-1) hops.

In particular, G{G} has hop diameter at most 2t12t-1, and diameter at most (2t1)diam(P)(2t-1)\cdot{\mathrm{diam}}_{{\mathcal{M}}}\left({{P}}\right).

Corollary 3.8.

Let =(𝒳,d){\mathcal{M}}=(\mathcal{X},{\mathrm{d}}) be a metric space, and let P𝒳{P}\subseteq\mathcal{X} be a finite subset of size nn. Given parameters tt\in\mathbb{N}, and ϑ(0,1){\vartheta}\in(0,1), there exists a weighted graph G{G} on P{P}, such that:

  1. (A)

    The graph G{G} has |E(G)|=𝒪(ϑ1n1+1/t)\left|{{E}\left({{G}}\right)}\right|={\mathcal{O}}\left({{\vartheta}^{-1}\cdot n^{1+1/t}}\right) edges.

  2. (B)

    The graph G{G} is ϑ{\vartheta}-reliable. Namely, given any attack set B𝒳{{B}}\subset\mathcal{X}, there exists a subset QP{Q}\subseteq{P}, such that |Q||P|(1+ϑ)|BP|\left|{{Q}}\right|\geq\left|{{P}}\right|-(1+{\vartheta})\left|{{{B}}\cap{P}}\right|. Furthermore, for any two points p,qQ{p},{q}\in{Q}, we have

    d(p,q)dG|Q(p,q)2tdiam(P),{\mathrm{d}}_{{\mathcal{M}}}\left({{p},{q}}\right)\leq{\mathrm{d}}_{{G}|_{Q}}\left({{p},{q}}\right)\leq 2t\cdot{\mathrm{diam}}_{{\mathcal{M}}}\left({{P}}\right),

    and the path realizing it has at most 2t2t hops.

In particular, G{G} has hop diameter at most 2t2t, and diameter at most 2tdiam(P)2t\cdot{\mathrm{diam}}_{{\mathcal{M}}}\left({{P}}\right).

Remark 3.9.

Corollary 3.7 and Corollary 3.8 are quite weak as far as the guarantee on the length of the resulting path (i.e., they are not good spanners). A construction that provides a spanner guarantee is provided below in Lemma 5.4 below.

As aside, proper expanders are reliable spanners because their expansion property is robust, as testified by the following.

Theorem 3.10 (reliable vertex expander).

For every ϑ(0,1){\vartheta}\in(0,1) there exists a constant c=cϑ>0{\mathcalb{c}}_{\!}={\mathcalb{c}}_{\!{\vartheta}}>0, such that for any expansion constant hc2h\geq{\mathcalb{c}}_{\!}^{-2}, there exists a family of vertex expander graphs {G=(V,E)}n\{{G}=({V},{E})\}_{n} on nn vertices of degree at most 4h/ϑ4h/{\vartheta}, with the following resiliency property: For any BV{{B}}\subset{V}, there exists B+B{{B^{+}}}\supseteq{{B}}, |B+|(1+ϑ)|B||{{B^{+}}}|\leq(1+{\vartheta})|{{B}}| such that the graph GB+{G}\setminus{{B^{+}}} is a vertex expander in the sense that

  1. (i)

    diam(GB+)2loghn\mathrm{diam}\left({{G}\setminus{{B^{+}}}}\right)\leq 2\lceil\log_{h}n\rceil, and

  2. (ii)

    For any UVB+U\subset{V}\setminus{{B^{+}}} of size |U|cn/h|U|\leq{\mathcalb{c}}_{\!}n/h, we have |ΓGB+(U)|h|U||\Gamma_{{G}\setminus{{B^{+}}}}(U)|\geq h|U|.

The proof is deferred to Section 6.3.4.

4 Covers for trees, bounded spread metrics, and planar graphs

Definition 4.1.

For a finite metric space =(P,𝖽){\mathcal{M}}=({P},{\mathsf{d}}), a tt-cover, is a family of subsets 𝒞={SiP|i=1,m}{\mathcal{C}}=\left\{{S}_{i}\subseteq{P}\;\middle|\;i=1,\ldots m\right\}, such that for any p,qP{p},{q}\in{P}, there exists an index ii, such that p,qSi{p},{q}\in{S}_{i}, and

diam(Si)/td(p,q)diam(Si).\mathrm{diam}\left({{S}_{i}}\right)/t\leq{\mathrm{d}}\left({{p},{q}}\right)\leq\mathrm{diam}\left({{S}_{i}}\right).

The size of a cover 𝒞{\mathcal{C}} is size(𝒞)=S𝒞|S|\mathrm{size}({\mathcal{C}})=\sum_{{S}\in{\mathcal{C}}}\left|{{S}}\right|. Recalling Definition 2.4, the degree in 𝒞{\mathcal{C}} of a point pP{p}\in{P} is the number of sets of 𝒞{\mathcal{C}} that contain it. The depth of 𝒞{\mathcal{C}} is the maximum degree of any element of P{P}, and is denoted by 𝖣(𝒞){\mathsf{D}}({\mathcal{C}}).

4.1 Lower bounds

Unfortunately, in the worst case, the depth and the size of any cover must depend on the spread of the metric.

Proposition 4.2.

For any parameter t>1t>1, any integer h>1h>1, Φ=th{\Phi}=t^{h}, and any nhn\geq h, there exists a metric =(P,d){\mathcal{M}}=({P},{\mathrm{d}}) of nn points, such that

  1. (I)

    Φ(P)=Φ{{\Phi}}\left({{P}}\right)={\Phi}, and

  2. (II)

    any tt-cover 𝒞{\mathcal{C}} of P{P} has size Ω(nlogtΦ)=Ω(nh)\Omega(n\log_{t}{\Phi})=\Omega(nh), average degree h/2\geq h/2, and depth hh.

Proof:

For simplicity of exposition, assume that hh divides nn. Let Pi{P}_{i} be a set of n/hn/h points, such that the distance between any pair of points of Pi{P}_{i} is i=(t+ε)i\ell_{i}=(t+\varepsilon)^{i}, for some fixed small ε>0\varepsilon>0, for i=1,,hi=1,\ldots,h. Furthermore, assume that the distance between any point of Pi{P}_{i} and any point of Pj{P}_{j}, for i<ji<j, is j\ell_{j}.

Let P=Pi{P}=\cup{P}_{i}, and observe that the distance function defined above is a metric (it is the union of uniform metrics of different resolutions). Now, consider any tt-cover 𝒞{\mathcal{C}} of P{P}. The rank of a cluster C𝒞{C}\in{\mathcal{C}}, is the highest jj, such that PjC{P}_{j}\cap{C}\neq\emptyset. We can assume that all the clusters of C{C} have at least two points, as otherwise they can be removed. For any index j[h]j\in\left[h\right], any point pPj{p}\in{P}_{j} and any point qi=1jPip{q}\in\cup_{i=1}^{j}{P}_{i}-{p}, by definition, there exists a cluster C𝒞{C}\in{\mathcal{C}}, such that p,qC{p},{q}\in{C}, and diam(C)tj\mathrm{diam}\left({{C}}\right)\leq t\cdot\ell_{j}, since d(p,q)=j{\mathrm{d}}\left({{p},{q}}\right)=\ell_{j}. It follows that C{C} can not contain any point of i=j+1hPi\cup_{i=j+1}^{h}{P}_{i}. Namely, the rank of C{C} is jj.

If there are two clusters of rank jj in 𝒞{\mathcal{C}}, then we can merge them, since merging does not increase their diameter, and such an operation does not increase the size of the cover, and the degrees of its elements. As such, in the end of this process, the cover 𝒞{\mathcal{C}} has hh clusters, and for any j[h]j\in\left[h\right], there is a cluster Cj𝒞{C}_{j}\in{\mathcal{C}} that is of rank jj, and contains (exactly) all the elements of i=1jPi\cup_{i=1}^{j}{P}_{i}. Namely, 𝒞={C1,,Ch}{\mathcal{C}}=\{{C}_{1},\ldots,{C}_{h}\}, where Cj=i=1jPi{C}_{j}=\cup_{i=1}^{j}{P}_{i}. It is easy to verify that this cover has the desired properties.  

Proposition 4.3.

For any t2,3,t\in{2,3,\ldots}, and any sufficiently large nn there exists an nn-point metric space for which any tt-cover must be of size at least Ω(n1+1/2t)\Omega(n^{1+{1}/{2t}}).

Proof:

Let g=2t+2g=2t+2. By a standard probabilistic argument, for any sufficiently large nn\in\mathbb{N} there exists a simple graph G=(V,E)G=(V,E) on nn vertices and m=Ω(n1+1g2)m=\Omega(n^{1+\frac{1}{g-2}}) edges whose girth is at least gg. (This is not the best known bound, but it is sufficient for our purposes.) Consider GG as a metric space with the shortest (unweighted) path metric. Let 𝒞{\mathcal{C}} be a tt-cover for GG, and let 𝒞={S𝒞|diam(S)t}{\mathcal{C}}^{\prime}=\left\{S\in{\mathcal{C}}\;\middle|\;\mathrm{diam}\left({S}\right)\leq t\right\}. For SVS\subset V let E(S)={uvE|u,vS}E(S)=\left\{uv\in E\;\middle|\;u,v\in S\right\}.

We claim that for every S𝒞S\in{\mathcal{C}}^{\prime}, E(S)E(S) is a forest, and hence |E(S)|<|S||E(S)|<|S|. Indeed, Suppose E(S)E(S) contains a cycle (v0,v1,,vh,v0)(v_{0},v_{1},\ldots,v_{h},v_{0}) such that vivi+1,v0vhE(S)Ev_{i}v_{i+1},v_{0}v_{h}\in E(S)\subseteq E. Denote di=d(v0,vi)d_{i}={\mathrm{d}}\left({v_{0},v_{i}}\right). Since {vi,vi+1}E\{v_{i},v_{i+1}\}\in E, we have di+1di{1,0,1}d_{i+1}-d_{i}\in\{-1,0,1\}. Let jj be the smallest ii such that di+1did_{i+1}\leq d_{i}. Hence, dj=jd_{j}=j. Let P=(vj+1,u1,uk,v0)P=(v_{j+1},u_{1},\ldots u_{k},v_{0}) be a shortest path in GG between vj+1v_{j+1} and v0v_{0}. PP’s length is at most dj=jd_{j}=j. Thus, the sequence v0,v1,vj,vj+1,u1,,uk,v0v_{0},v_{1},\ldots v_{j},v_{j+1},u_{1},\ldots,u_{k},v_{0} is closed path of length at most 2j+12j+1, and hence contains a cycle of length at most 2j+12j+1. By the girth condition, 2j+12t+22j+1\geq 2t+2, and hence dj=j>td_{j}=j>t. But since v0,vjSv_{0},v_{j}\in S, this means that diam(S)dj>t\mathrm{diam}\left({S}\right)\geq d_{j}>t which contradicts the definition of 𝒞{\mathcal{C}}^{\prime}.

For every edge pqEpq\in E, d(p,q)=1{\mathrm{d}}\left({p,q}\right)=1, and by the tt-covering condition there exists S𝒞S\in\mathcal{C}^{\prime} such that p,qSp,q\in S. Therefore pqE(S)pq\in E(S). Hence ES𝒞E(S)E\subset\bigcup_{S\in{\mathcal{C}}^{\prime}}E(S), and therefore

cn1+12t|E|S𝒞|E(S)|<S𝒞|S|=size(𝒞)size(𝒞).cn^{1+\tfrac{1}{2t}}\leq|E|\leq\sum_{S\in\mathcal{C}^{\prime}}|E(S)|<\sum_{S\in{\mathcal{C}}^{\prime}}|S|=\mathrm{size}({\mathcal{C}}^{\prime})\leq\mathrm{size}({\mathcal{C}}).

 

4.2 Cover for ultrametrics

Definition 4.4.

A hierarchically well-separated tree (HST) is a metric space defined on the leaves of a rooted tree T{T}. To each vertex uTu\in{T} there is an associated label Δu0\Delta_{u}\geq 0. This label is zero for all the leaves of T{T}, and it is a positive number for all the interior nodes. The labels satisfy for every non-root vertex vTv\in{T}, ΔvΔp¯(v)/k\Delta_{v}\leq\Delta_{\overline{\mathrm{p}}\left({v}\right)}/k, where p¯(v)\overline{\mathrm{p}}\left({v}\right) is the parent of vv in T{T}. The distance between two leaves x,yTx,y\in{T} is defined as Δlca(x,y)\Delta_{\mathop{\mathrm{lca}}(x,y)}, where lca(x,y)\mathop{\mathrm{lca}}(x,y) is the least common ancestor of xx and yy in T{T}. An HST T{T} is a kk-HST if for every non-root vertex vTv\in{T}, ΔvΔp¯(v)/k\Delta_{v}\leq\Delta_{\overline{\mathrm{p}}\left({v}\right)}/k.

HSTs are one of the simplest non-trivial metrics, and surprisingly, general metrics can be embedded randomly into HSTs with expected distortion of 𝒪(logn){\mathcal{O}}(\log n) [Bar98, FRT04].

Definition 4.5.

A metric =(P,d){\mathcal{M}}=({P},{\mathrm{d}}) is an ultrametric, if for any x,y,zPx,y,z\in{P}, we have that d(x,z)max(d(x,y),d(y,z)){\mathrm{d}}\left({x,z}\right)\leq\max({\mathrm{d}}\left({x,y}\right),{\mathrm{d}}\left({y,z}\right)). Notice, that this is a stronger version of the triangle inequality, which states that d(x,z)d(x,y)+d(y,z){\mathrm{d}}\left({x,z}\right)\leq{\mathrm{d}}\left({x,y}\right)+{\mathrm{d}}\left({y,z}\right).

The following is folklore, and it also easy to verify (see, e.g., [BLMN05, Lemma 3.5].

Lemma 4.6.

For k1k\geq 1, every finite ultrametric can be kk-approximated by a kk-HST.

Lemma 4.7.

For k>1k>1, every kk-HST with spread Φ{\Phi} has 11-cover of depth at most logkΦ\log_{k}{\Phi}.

Proof:

Let TT be the tree of the HST. With every vertex uTu\in T we associate a cluster CuC_{u} the leaves of the subtree rooted at uu. The covers is defined as 𝒞={Cu|uT}{\mathcal{C}}=\left\{C_{u}\;\middle|\;u\in T\right\}. The properties of the cover are immediate.  

Corollary 4.8.

Let =(P,d){\mathcal{M}}=({P},{\mathrm{d}}) be an ultrametric over nn points with spread Φ{\Phi}. For any ε(0,1)\varepsilon\in(0,1), one can compute a (1+ε)(1+\varepsilon)-cover of {\mathcal{M}} of depth 𝒪(ε1logΦ){\mathcal{O}}(\varepsilon^{-1}\log{\Phi}).

4.3 Cover for general finite metrics

We need the following result.

Lemma 4.9 ([MN07]).

Let (P,d)({P},{\mathrm{d}}) be an nn-point metric space and k1k\geq 1. Then there exists a distribution over decreasing sequences of subsets P=P0P1Ps={P}={P}_{0}\supsetneq{P}_{1}\supsetneq\cdots\supsetneq{P}_{s}=\emptyset (ss itself is a random variable), such that for all μ>1/k\mu>-1/k, we have 𝔼[j=1s1|Pj|μ]max(k1+μk,1)nμ+1/k,\mathop{\mathbb{E}}\!\left[{\sum_{j=1}^{s-1}|{P}_{j}|^{\mu}}\right]\leq\max\left({\frac{k}{1+\mu k},1}\right)\cdot n^{\mu+1/k}, and such that for each j[s]j\in\left[s\right] there exists an ultrametric ρj\rho_{j} on Pj1{P}_{j-1} satisfying for every p,qP{p},{q}\in{P}, that ρj(p,q)d(p,q)\rho_{j}({p},{q})\geq{\mathrm{d}}\left({{p},{q}}\right), and if pPj1{p}\in{P}_{j-1} and qPj1Pj{q}\in{P}_{j-1}\setminus{P}_{j} then ρj(p,q)𝒪(k)d(p,q)\rho_{j}({p},{q})\leq{\mathcal{O}}(k)\cdot{\mathrm{d}}\left({{p},{q}}\right).

By computing a cover (using Corollary 4.8) for each ultrametric generated by the above lemma, we get the following.

Lemma 4.10.

For an nn-point metric space =(P,𝖽){\mathcal{M}}=({P},{\mathsf{d}}) with spread Φ{\Phi}, and a parameter k>1k>1, one can compute, in polynomial time, an 𝒪(k){\mathcal{O}}(k)-cover of {\mathcal{M}} of size 𝒪(n1+1/klogΦ){\mathcal{O}}(n^{1+1/k}\log{\Phi}) and depth 𝒪(kn1/klogΦ){\mathcal{O}}(kn^{1/k}\log{\Phi}).

Proof:

Using Lemma 4.9, for the parameter kk, compute the sequence P=P0P1Ps={P}={P}_{0}\supsetneq{P}_{1}\supsetneq\cdots\supsetneq{P}_{s}=\emptyset, and the associated ultrametrics ρ1,,ρs1\rho_{1},\ldots,\rho_{s-1}. We build a 11-HST Ti{T}_{i} for ρi\rho_{i}, when restricted to the set Pi1{P}_{i-1}, for i=1,,s1i=1,\ldots,s-1. For every HST in this collection, we compute a 22-cover using Corollary 4.8. Let 𝒞{\mathcal{C}} be the union of all these covers.

Since for every HST Ti{T}_{i} the resulting cover has size 𝒪(|Pi|logΦ){\mathcal{O}}(\left|{{P}_{i}}\right|\log{\Phi}), then by Lemma 4.9 (applied with μ=1\mu=1), we have

𝔼[.i|Pi|logΦ]=𝒪(n1+1/klogΦ).\mathop{\mathbb{E}}\!\left[{\Bigl{.}\smash{\sum_{i}}\left|{{P}_{i}}\right|\log{\Phi}}\right]={\mathcal{O}}(n^{1+1/k}\log{\Phi}).

As for the quality of the cover, let p,q{p},{q} be any two points in P{P}, and assume (without loss of generality) that pPi1{p}\in{P}_{i-1} and qPi1Pi{q}\in{P}_{i-1}\setminus{P}_{i} for some i[s]i\in[s]. We have that ρi(p,q)=𝒪(k)d(p,q)\rho_{i}({p},{q})={\mathcal{O}}(k)\cdot{\mathrm{d}}\left({{p},{q}}\right), and since we computed a 22-cover for this tree, there is a cluster in the computed cover that contains both points and its diameter is at most twice the distance between those points.

The maximum depth, follows by using Lemma 4.9 with μ=0\mu=0. This implies that a point of P{P} participates in s=𝒪(kn1/k)s={\mathcal{O}}(kn^{1/k}) HSTs, and each cover generated by such an HST might a point an element at most 𝒪(logΦ){\mathcal{O}}(\log{\Phi}) times.

The bounds on the size and depth hold in expectation, and one can repeat the construction if they exceed the desired size by (say) a factor of eight. In expectation, after a constant number of iterations, the algorithm would compute with high probability a cover with the desired bounds.  

4.4 Covers for trees

Using a tree separator, and applying it recursively, implies the following construction of covers for trees.

Lemma 4.11.

For a weighted tree metric T=(P,𝖽){T}=({P},{\mathsf{d}}), with spread Φ{\Phi}, and a parameter ε(0,1)\varepsilon\in(0,1), one can compute in polynomial time a (2+ε)(2+\varepsilon)-cover of T{T} of depth 𝒪(ε1logΦlogn){\mathcal{O}}(\varepsilon^{-1}\log{\Phi}\log n), and size 𝒪(nε1logΦlogn){\mathcal{O}}(n\varepsilon^{-1}\log{\Phi}\log n), where n=|P|n=\left|{{P}}\right|.

Proof:

The proof is by induction on nn. When n=1n=1 the trivial cover is sufficient. Assume next that n1n\geq 1 and the minimum distance in T{T} is one. Find a separator node 𝗌{\mathsf{s}} in T{T} such that there is no connected component in T{T} larger than n/2n/2 after removing 𝗌{\mathsf{s}}. For i{0,1,2,,m=log1+ε/2Φ}i\in\{0,1,2,\dots,m=\left\lceil{\smash{\log_{1+\varepsilon/2}{\Phi}}}\right\rceil\}, define the sets P(𝗌,i)={pP|d(𝗌,p)(1+ε/2)i}.{P}({\mathsf{s}},i)=\left\{{p}\in{P}\;\middle|\;{\mathrm{d}}\left({{\mathsf{s}},{p}}\right)\leq(1+\varepsilon/2)^{i}\right\}. By the inductive hypothesis, for each connected component QQ of T𝗌{T}-{\mathsf{s}} there is a (2+ε)(2+\varepsilon)-cover 𝒞Q{\mathcal{C}}_{Q} of QQ of depth 𝒪(ε1logΦlog(n/2)){\mathcal{O}}(\varepsilon^{-1}\log{\Phi}\log(n/2)) and size 𝒪(nε1logΦlog(n/2)){\mathcal{O}}(n\varepsilon^{-1}\log{\Phi}\log(n/2)). The cover for PP is

𝒞P={P(𝗌,i)|i=0,1,2,,m}Q𝒞Q.{\mathcal{C}}_{P}=\left\{{P}({\mathsf{s}},i)\;\middle|\;i=0,1,2,\dots,m\right\}\cup{\bigcup\nolimits_{Q}}{\mathcal{C}}_{Q}.

Since every element of P{P} participates in at most mm sets in each level of the recursion, the bound on the depth and size is immediate.

As for the quality of the cover, by the inductive hypothesis, we need only to check pairs of points p,qP{p},{q}\in{P}, that are separated by 𝗌{\mathsf{s}}. Assume d(𝗌,p)d(𝗌,q){\mathrm{d}}\left({{\mathsf{s}},{p}}\right)\geq{\mathrm{d}}\left({{\mathsf{s}},{q}}\right), and let jj be the minimum index, such that d(𝗌,p)(1+ε/2)j{\mathrm{d}}\left({{\mathsf{s}},{p}}\right)\leq(1+\varepsilon/2)^{j}. We have that p,qP(𝗌,j){p},{q}\in{P}({\mathsf{s}},j), and

d(p,q)d(𝗌,p)(1+ε/2)j1anddiam(P(𝗌,j))2(1+ε/2)j.{\mathrm{d}}\left({{p},{q}}\right)\geq{\mathrm{d}}\left({{\mathsf{s}},{p}}\right)\geq(1+\varepsilon/2)^{j-1}\qquad\text{and}\qquad\mathrm{diam}\left({{P}({\mathsf{s}},j)}\right)\leq 2(1+\varepsilon/2)^{j}.

As such, we have diam(P(𝗌,j))d(p,q)2(1+ε/2)j(1+ε/2)j1=2+ε.\displaystyle\frac{\mathrm{diam}\left({{P}({\mathsf{s}},j)}\right)}{{\mathrm{d}}\left({{p},{q}}\right)}\leq\frac{2(1+\varepsilon/2)^{j}}{(1+\varepsilon/2)^{j-1}}=2+\varepsilon.  

4.5 Covers for planar graphs

The next lemma can be traced back to the work of Lipton and Tarjan [LT79].

Lemma 4.12.

Let H=(P,E){H}=({P},{E}) be a planar triangulated graph with non-negative edge weights. There is a partition of P{P} to three sets X,Y,Z{X},{Y},{Z}, such that

  1. (i)

    |X|(2/3)n\left|{{X}}\right|\leq(2/3)n and |Y|(2/3)n\left|{{Y}}\right|\leq(2/3)n,

  2. (ii)

    there is no edge between X{X} and Y{Y}, and

  3. (iii)

    Z{Z} is composed of two interior disjoint shortest paths that share one of their endpoints, and an edge connecting their other two endpoints.

Definition 4.13.

For a metric space (𝒳,d)(\mathcal{X},{\mathrm{d}}) and a parameter rr, an rr-net N{N} is a maximal set of points in 𝒳\mathcal{X} satisfying:

  1. (i)

    For any two net points p,qN{p},{q}\in{N}, pqp\neq q, we have d(p,q)>r{\mathrm{d}}\left({{p},{q}}\right)>r.

  2. (ii)

    For any p𝒳{p}\in\mathcal{X}, 𝖽(p,N)=minqNd(p,q)r{\mathsf{d}}\left({{p},{N}}\right)=\min_{{q}\in{N}}{\mathrm{d}}\left({{p},{q}}\right)\leq r.

A net can be computed by repeatedly picking the point furthest away from the current net N{N}, and adding it to the net if this distance is larger than rr, and stopping otherwise. We denote a net computed by this algorithm by net(𝒳,r){\mathrm{net}}\left({\mathcal{X},r}\right).

The following lemma testifies that if we restrict the net to lay along a shortest path in the graph, locally the cover it induces has depth as if the graph was one dimensional.

Lemma 4.14.

Let G{G} be a weighted graph, and let d{\mathrm{d}} be the shortest path metric it induces. Let π\pi be a shortest path in G{G} and let N=net(π,r)π{N}={\mathrm{net}}\left({\pi,r}\right)\subseteq\pi be a net computed for some distance r>0r>0. For some R>0R>0, consider the set of balls ={𝖻(p,R)|pN}{\mathcal{B}}=\left\{{\mathsf{b}}\left({{p},R}\right)\;\middle|\;{p}\in{N}\right\}. For any point qV(G){q}\in{V}\left({{G}}\right), we have that the degree of q{q} in {\mathcal{B}} is at most 2R/r+12R/r+1.

Proof:

Let p1,,pk{p}_{1},\ldots,{p}_{k} be the points of N𝖻(q,R){N}\cap{\mathsf{b}}\left({{q},R}\right) sorted by their order along π\pi – these are the only points that their balls in {\mathcal{B}} would contain q{q}. By the definition of the net, d(pi,pi+1)>r{\mathrm{d}}\left({{p}_{i},{p}_{i+1}}\right)>r for all ii. Since π\pi is a shortest path, we also have that

2Rdiam(𝖻(q,R))d(p1,pk)=i=1k1d(pi,pi+1)>(k1)r.2R\geq\mathrm{diam}\left({{\mathsf{b}}\left({{q},R}\right)}\right)\geq{\mathrm{d}}\left({{p}_{1},{p}_{k}}\right)=\sum_{i=1}^{k-1}{\mathrm{d}}\left({{p}_{i},{p}_{i+1}}\right)>(k-1)r.

 

Construction

Let ε(0,1)\varepsilon\in(0,1) be an input parameter, and let G{G} be a weighted planar graph. We assume that G{G} is triangulated, as otherwise it can be triangulated (we also assume that we have its planar embedding). Any new edge pq{p}{q} is assigned as weight the distance between its endpoints in the original graph. This can be done in linear time. As usual, we assume that the minimum distance in G{G} is one, and its spread is Φ{\Phi}.

Let Z{Z} be the cycle separator given by Lemma 4.12 made out of two shortest paths π1\pi_{1} and π2\pi_{2}. Let p1,p2,p3{p}_{1},{p}_{2},{p}_{3} be the endpoints of these two paths.

For i=0,,m=log1+ε/8Φi=0,\ldots,m=\left\lceil{\smash{\log_{1+\varepsilon/8}{\Phi}}}\right\rceil, let Ni=net(π1,εri/8)net(π2,εri/8){p1,p2,p3}{N}_{i}={\mathrm{net}}\left({\pi_{1},\varepsilon r_{i}/8}\right)\cup{\mathrm{net}}\left({\pi_{2},\varepsilon r_{i}/8}\right)\cup\{{p}_{1},{p}_{2},{p}_{3}\}, where ri=(1+ε/8)ir_{i}=(1+\varepsilon/8)^{i}. The associated set of balls is

i={𝖻(p,(1+ε/8)ri)|pNi}.{\mathcal{B}}_{i}=\left\{{\mathsf{b}}\left({{p},(1+\varepsilon/8)r_{i}}\right)\;\middle|\;{p}\in{N}_{i}\right\}.

The resulting set of balls is (Z)=ii{\mathcal{B}}({Z})=\bigcup_{i}{\mathcal{B}}_{i}. We add the sets of (Z){\mathcal{B}}({Z}) to the cover, and continue recursively on the connected components of GZ{G}-{Z}. Let 𝒞{\mathcal{C}} denote the resulting cover.

Analysis
Lemma 4.15.

For any two vertices p,qV(G){p},{q}\in{V}\left({{G}}\right), there exists a cluster C𝒞{C}\in{\mathcal{C}}, such that p,qC{p},{q}\in{C}, and diam(C)(2+ε)dG(p,q).\mathrm{diam}\left({{C}}\right)\leq(2+\varepsilon){\mathrm{d}}_{G}\left({{p},{q}}\right). That is, 𝒞{\mathcal{C}} is a (2+ε)(2+\varepsilon)-cover of G{G}.

Proof:

Assume, for the simplicity of exposition, that p{p} and q{q} get separated in the top level of the recursion (otherwise, apply the argument to the inductive step in which they get separated). The shortest path between p{p} and q{q} must intersect the separator Z{Z}, say at a vertex vv. Assume that dG(p,v)dG(v,q){\mathrm{d}}_{G}\left({{p},v}\right)\geq{\mathrm{d}}_{G}\left({v,{q}}\right) and that rj1dG(p,v)rj=(1+ε/8)jr_{j-1}\leq{\mathrm{d}}_{G}\left({{p},v}\right)\leq r_{j}=(1+\varepsilon/8)^{j}. There is a point uNju\in{N}_{j} within distance εrj/8\varepsilon r_{j}/8 from vv. As such,

dG(p,u)dG(p,v)+dG(v,u)(1+ε/8)j+(ε/8)(1+ε/8)j(1+ε/8)j+1,{\mathrm{d}}_{G}\left({{p},u}\right)\leq{\mathrm{d}}_{G}\left({{p},v}\right)+{\mathrm{d}}_{G}\left({v,u}\right)\leq(1+\varepsilon/8)^{j}+(\varepsilon/8)(1+\varepsilon/8)^{j}\leq(1+\varepsilon/8)^{j+1},

which implies that p𝖻=𝖻(u,(1+ε/8)rj){p}\in{\mathsf{b}}={\mathsf{b}}\left({u,(1+\varepsilon/8)r_{j}}\right). A similar argument shows that q{q} is also in 𝖻{\mathsf{b}}. Furthermore, we have that d(p,q)d(p,v)rj1.{\mathrm{d}}\left({{p},{q}}\right)\geq{\mathrm{d}}\left({{p},v}\right)\geq r_{j-1}. Note that 𝖻j𝒞{\mathsf{b}}\in{\mathcal{B}}_{j}\subseteq{\mathcal{C}}. We have that

diam(𝖻)d(p,q)2(1+ε/8)rjrj1=2(1+ε/8)22+ε.\frac{\mathrm{diam}\left({{\mathsf{b}}}\right)}{{\mathrm{d}}\left({{p},{q}}\right)}\leq\frac{2(1+\varepsilon/8)r_{j}}{r_{j-1}}=2(1+\varepsilon/8)^{2}\leq 2+\varepsilon.

 

Lemma 4.16.

The depth of 𝒞{\mathcal{C}} is 𝒪(ε2lognlogΦ){\mathcal{O}}(\varepsilon^{-2}\log n\log{\Phi}).

Proof:

Fix a vertex p{p}. By Lemma 4.14, for each ii, at most

3+2(2(1+ε/8)riεri/8+1)=𝒪(1/ε).3+2\left({\frac{2(1+\varepsilon/8)r_{i}}{\varepsilon r_{i}/8}+1}\right)={\mathcal{O}}(1/\varepsilon).

balls of i{\mathcal{B}}_{i} contains p{p}. The number of such sets is 𝒪(log1+ε/8Φ)=𝒪(ε1logΦ){\mathcal{O}}(\log_{1+\varepsilon/8}{\Phi})={\mathcal{O}}(\varepsilon^{-1}\log{\Phi}). The vertex p{p} get sent down to at most one recursive subproblem, and the recursion depth is 𝒪(logn){\mathcal{O}}(\log n). It follows that the depth of any point (and the degree of p{p} specifically) is at most 𝒪(ε2logΦlogn){\mathcal{O}}(\varepsilon^{-2}\log{\Phi}\log n).  

Theorem 4.17.

Let G{G} be a weighted planar graph over nn vertices with spread Φ{\Phi}. Then, given a parameter ε(0,1)\varepsilon\in(0,1), one can construct a (2+ε)(2+\varepsilon)-cover of G{G} with depth 𝒪(ε2lognlogΦ){\mathcal{O}}(\varepsilon^{-2}\log n\log{\Phi}) in polynomial time.

Remark 4.18.

It is possible to generalize Theorem 4.17 to the shortest path metric on families of graphs excluding a fixed minor. Specifically, by [KLMN05, Lemma 3.3], there exists O(s2)O(s^{2})-cover of depth O(3slogΦ)O(3^{s}\log{\Phi}) for every metric space supported on a graph excluding KsK_{s} minor and spread Φ{\Phi}. It may be possible to improve the approximation parameter to O(s)O(s) using [AGG+19]. This approach does not have a logn\log n factor in the depth parameter, but it can not provide a (2+ε)(2+\varepsilon)-approximation as in Theorem 4.17. As communicated to us by an anonymous referee, the approach used here to prove Theorem 4.17 can also be extended to any family of graphs excluding fixed minor and obtain (2+ε)(2+\varepsilon)-cover using the shortest paths separators of [AG06]. We have not pursued those avenues.

5 From covers to reliable spanners

5.1 The oblivious construction

Lemma 5.1.

Let =(P,d){\mathcal{M}}=({P},{\mathrm{d}}) be a finite metric space, and suppose there exists a ξ{\xi}-cover 𝒞{\mathcal{C}} of {\mathcal{M}} of size ss and depth 𝖣{\mathsf{D}}. Then, there exists an oblivious ϑ{\vartheta}-reliable 22-hop 2ξ2{\xi}-spanner for {\mathcal{M}}, of size 𝒪(s𝖣ϑlog𝖣ϑ).{\mathcal{O}}\bigl{(}s\frac{{\mathsf{D}}}{{\vartheta}}\log\frac{{\mathsf{D}}}{{\vartheta}}\bigr{)}.

Proof:

For each cluster C𝒞{C}\in{\mathcal{C}}, let C be a random constellation graph on C{C} as defined in Section 3.1, with reliability parameter ψ=ϑ/𝖣{\psi}={\vartheta}/{\mathsf{D}}. The resulting spanner is the union C𝒞C\bigcup_{{C}\in{\mathcal{C}}}{}_{C}. The number of edges in the resulting graph is at most

C𝒞𝒪(|C|ψ1logψ1)=𝒪(sψ1logψ1)=𝒪(s𝖣ϑlog𝖣ϑ).\sum_{C\in{\mathcal{C}}}{\mathcal{O}}(\left|{C}\right|{\psi}^{-1}\log{\psi}^{-1})={\mathcal{O}}(s{\psi}^{-1}\log{\psi}^{-1})={\mathcal{O}}\Bigl{(}s\frac{{\mathsf{D}}}{{\vartheta}}\log\frac{{\mathsf{D}}}{{\vartheta}}\Bigr{)}.

Fix an attack set BP{{B}}\subset{P}. A cluster C𝒞{C}\in{\mathcal{C}} is failed if |C|(1+ψ)|CB|\left|{C}\right|\leq(1+{\psi})\left|{C\cap{{B}}}\right|. Denote the set failed clusters by {\mathcal{F}}. For C𝒞C\in{\mathcal{C}}\setminus{\mathcal{F}}, let dmg(C,B)\mathrm{dmg}\left({C,{{B}}}\right) be the (random) set of damaged points in CC as defined in the proof of Lemma 3.1, i.e., dmg(C,B)=C\mathrm{dmg}\left({C,{{B}}}\right)=C if B{{B}} contains all the constellation’s centers, and dmg(C,B)=B\mathrm{dmg}\left({C,{{B}}}\right)={{B}} otherwise. by Lemma 3.1, 𝔼[|dmg(C,B)|](1+ψ)|BC|.\mathop{\mathbb{E}}\!\left[{|\mathrm{dmg}\left({C,{{B}}}\right)|}\right]\leq(1+{\psi})\left|{{{B}}\cap C}\right|. The damaged set is defined as

B+=(CC)(C𝒞dmg(C,B)).{{B^{+}}}=\Bigl{(}\bigcup_{C\in{\mathcal{F}}}C\Bigr{)}\cup\Bigl{(}\bigcup_{C\in{\mathcal{C}}\setminus{\mathcal{F}}}\mathrm{dmg}\left({C,{{B}}}\right)\Bigr{)}.

We next bound expected size of the loss B+B{{B^{+}}}\setminus{{B}}:

𝔼[|B+B|]\displaystyle\mathop{\mathbb{E}}\!\left[{\left|{{{B^{+}}}\setminus{{B}}}\right|}\right] 𝔼[.C|CB|]+C𝒞𝔼[|dmg(C,B)B|.]\displaystyle\leq\mathop{\mathbb{E}}\!\left[{\Bigl{.}\smash{\sum_{C\in{\mathcal{F}}}}|C\setminus{{B}}|}\right]+\sum_{C\in{\mathcal{C}}\setminus{\mathcal{F}}}\mathop{\mathbb{E}}\!\left[{\left|{\mathrm{dmg}\left({C,{{B}}}\right)\setminus{{B}}}\right|\Bigr{.}\!\,}\right]
ψC|BC|+ψC𝒞|BC|=ψpBC𝒞1pCψ𝖣|B|=ϑ|B|.\displaystyle\leq{\psi}\sum_{C\in{\mathcal{F}}}|{{B}}\cap C|+{\psi}\sum_{C\in{\mathcal{C}}\setminus{\mathcal{F}}}|{{B}}\cap C|={\psi}\sum_{{p}\in{{B}}}\sum_{C\in{\mathcal{C}}}1_{{p}\in C}\leq{\psi}{\mathsf{D}}|B|={\vartheta}|B|.

Finally, for any two points p,qPB+{p},{q}\in{P}\setminus{{B^{+}}}, there exists a non-failed cluster C𝒞C\in{\mathcal{C}} that contains both points, such that diam(C)ξd(p,q)\mathrm{diam}\left({C}\right)\leq{\xi}{\mathrm{d}}\left({{p},{q}}\right). As such, the two hops in the resulting graph are going to be of length at most 2diam(C)2ξd(p,q)2\mathrm{diam}\left({C}\right)\leq 2{\xi}{\mathrm{d}}\left({{p},{q}}\right).  

5.2 The deterministic construction

Lemma 5.2.

Let =(P,d){\mathcal{M}}=({P},{\mathrm{d}}) be a finite metric space over nn points, and let 𝒞{\mathcal{C}} be a ξ{\xi}-cover of it of depth 𝖣{\mathsf{D}} and size ss. Then, for any integer t1t\geq 1, there exists:

  1. (A)

    A ϑ{\vartheta}-reliable (2t1)(2t-1)-hop (2t1)ξ(2t-1){\xi}-spanner for {\mathcal{M}}, of size 𝒪(ϑ2𝖣2sn1/t){\mathcal{O}}({\vartheta}^{-2}{\mathsf{D}}^{2}sn^{1/t}).

  2. (B)

    A ϑ{\vartheta}-reliable 2t2t-hop 2tξ2t{\xi}-spanner for {\mathcal{M}}, of size 𝒪(ϑ1𝖣sn1/t){\mathcal{O}}({\vartheta}^{-1}{\mathsf{D}}sn^{1/t}).

Proof:

(A) For a cluster C𝒞{C}\in{\mathcal{C}}, let G(C){G}({C}) be a ψ{\psi}-reliable spanner constructed using Corollary 3.7 on C{C}, with the reliability parameter ψ=ϑ/𝖣{\psi}={\vartheta}/{\mathsf{D}}. Let G=(P,E){G}=({P},E) be a graph whose edge set EE is the union of the edge sets of G(C){G}({C}) for C𝒞{C}\in{\mathcal{C}}.

Let nin_{i} be the size of the iith cluster, for i{1,,m=|𝒞|}i\in\{1,\ldots,m=\left|{{\mathcal{C}}}\right|\}. The number of edges in G{G} is bounded by 𝒪(ψ2N)=𝒪(𝖣2ϑ2N){\mathcal{O}}({\psi}^{-2}N)={\mathcal{O}}({\mathsf{D}}^{2}{\vartheta}^{-2}N), where

N=i=1mni1+1/t=i=1mni1/tni(maxini1/t)i=1mnin1/ts,N=\sum_{i=1}^{m}n_{i}^{1+1/t}=\sum_{i=1}^{m}n_{i}^{1/t}n_{i}\leq(\max_{i}n_{i}^{1/t})\sum_{i=1}^{m}n_{i}\leq n^{1/t}s,

since ini=s\sum_{i}n_{i}=s, and maxinin\max_{i}n_{i}\leq n. We conclude that the total number of edges in G{G} is 𝒪(ϑ2𝖣2sn1/t){\mathcal{O}}({\vartheta}^{-2}{\mathsf{D}}^{2}sn^{1/t}).

Let BP{{B}}\subset{P} be an attack set. The damage set B+B{{B^{+}}}\supseteq{{B}} is constructed as in the proof of Lemma 5.1. Thus, as argued there, |B+|(1+ϑ)|B|\left|{{{B^{+}}}}\right|\leq(1+{\vartheta})\left|{{{B}}}\right|. For every two points p,qPB+{p},{q}\in{P}\setminus{{B^{+}}}, there exists a cluster C𝒞{C}\in{\mathcal{C}}, such that p,qC{p},{q}\in{C}, and

diam(C)/ξd(p,q)diam(C).\mathrm{diam}\left({{C}}\right)/{\xi}\leq{\mathrm{d}}\left({{p},{q}}\right)\leq\mathrm{diam}\left({{C}}\right).

Furthermore, by Corollary 3.7, we have

dGB(p,q)dG(C)B(p,q)(2t1)diam(C)(2t1)ξd(p,q).{\mathrm{d}}_{{G}\setminus{{B}}}\left({{p},{q}}\right)\leq{\mathrm{d}}_{{G}({C})\setminus{{B}}}\left({{p},{q}}\right)\leq(2t-1)\mathrm{diam}\left({{C}}\right)\leq(2t-1){\xi}\cdot{\mathrm{d}}\left({{p},{q}}\right).

and shortest path in G(C)B{G}({C})\setminus{{B}} realizing dG(C)B(p,q){\mathrm{d}}_{{G}({C})\setminus{{B}}}\left({{p},{q}}\right) has at most 2t12t-1 hops.

(B) Similar to the above, except for using Corollary 3.8 instead of Corollary 3.7.  

5.3 Applications

5.3.1 General metrics

Lemma 5.3.

Let =(P,d){\mathcal{M}}=({P},{\mathrm{d}}) be an nn-point metric space of spread at most Φ{\Phi}, and let ϑ(0,1){\vartheta}\in(0,1) and kk\in\mathbb{N} be parameters. Then, one can build an oblivious ϑ{\vartheta}-reliable 𝒪(k){\mathcal{O}}(k)-spanner for {\mathcal{M}} with

𝒪(ϑ1kn1+1/klog2Φlogkn1/klogΦϑ){\mathcal{O}}\Bigl{(}{\vartheta}^{-1}kn^{1+1/k}\log^{2}{\Phi}\log\frac{kn^{1/k}\log{\Phi}}{{\vartheta}}\Bigr{)}

edges. In particular, for k=lognk=\log n, we obtain a ϑ{\vartheta}-reliable 𝒪(logn){\mathcal{O}}(\log n)-spanner for {\mathcal{M}} with

𝒪(ϑ1nlognlog2Φ(loglogn+loglogΦ+logϑ1){\mathcal{O}}\bigl{(}{\vartheta}^{-1}n\log n\log^{2}{\Phi}(\log\log n+\log\log{\Phi}+\log{\vartheta}^{-1}\bigr{)}

edges.

Proof:

By Lemma 4.10, {\mathcal{M}} has a ξ=𝒪(k){\xi}={\mathcal{O}}(k)-cover of size 𝒪(n1+1/2klogΦ){\mathcal{O}}(n^{1+1/2k}\log{\Phi}) and depth 𝖣=𝒪(kn1/2klogΦ){\mathsf{D}}={\mathcal{O}}(kn^{1/2k}\log{\Phi}). Plugging this into Lemma 5.1 yields an oblivious ϑ{\vartheta}-reliable 𝒪(k){\mathcal{O}}(k)-spanner with 𝒪(ϑ1kn1+1/klog2Φlogkn1/klogΦϑ){\mathcal{O}}\bigl{(}{\vartheta}^{-1}kn^{1+1/k}\log^{2}{\Phi}\log\frac{kn^{1/k}\log{\Phi}}{{\vartheta}}\bigr{)} edges.  

Lemma 5.4.

Let =(P,d){\mathcal{M}}=({P},{\mathrm{d}}) be a finite metric over nn points of spread Φ{\Phi}, and let ϑ(0,1){\vartheta}\in(0,1) and k,tk,t\in\mathbb{N} be parameters. Then, one can build a ϑ{\vartheta}-reliable 𝒪(kt){\mathcal{O}}(kt)-spanner for {\mathcal{M}} with 𝒪(ϑ1kn1+1/k+1/tlog2Φ){\mathcal{O}}({\vartheta}^{-1}kn^{1+1/k+1/t}\log^{2}{\Phi}) edges. In particular, when t=lognt=\log n, we obtain a ϑ{\vartheta}-reliable 𝒪(klogn){\mathcal{O}}(k\log n)-spanner for {\mathcal{M}} with

𝒪(ϑ1kn1+1/(2k)log2Φ){\mathcal{O}}({\vartheta}^{-1}kn^{1+1/(2k)}\log^{2}{\Phi})

edges, and when t=kt=k we obtain ϑ{\vartheta}-reliable 𝒪(t2){\mathcal{O}}(t^{2})-spanner for {\mathcal{M}} with 𝒪(ϑ1tn1+1/tlog2Φ){\mathcal{O}}({\vartheta}^{-1}tn^{1+1/t}\log^{2}{\Phi}) edges.

Proof:

By Lemma 4.10, {\mathcal{M}} has a 𝒪(k){\mathcal{O}}(k)-cover of size s=𝒪(n1+1/2klogΦ)s={\mathcal{O}}(n^{1+1/2k}\log{\Phi}) and depth 𝖣=𝒪(kn1/2klogΦ){\mathsf{D}}={\mathcal{O}}(kn^{1/2k}\log{\Phi}). Plugging this into Lemma 5.2 (B), yields a ϑ{\vartheta}-reliable 𝒪(kt){\mathcal{O}}(kt)-spanner with the number of edges bounded by

𝒪(ϑ1𝖣sn1/t)=𝒪(ϑ1kn1+1/k+1/tlog2Φ).{\mathcal{O}}({\vartheta}^{-1}{\mathsf{D}}sn^{1/t})={\mathcal{O}}({\vartheta}^{-1}kn^{1+1/k+1/t}\log^{2}{\Phi}).

 

5.3.2 Ultrametrics

Lemma 5.5.

Let =(P,d){\mathcal{M}}=({P},{\mathrm{d}}) be an ultrametric over nn points with spread Φ{\Phi}, and let ϑ,ε(0,1){\vartheta},\varepsilon\in(0,1) be parameters. Then, one can build an oblivious ϑ{\vartheta}-reliable (2+ε)(2+\varepsilon)-spanner for {\mathcal{M}} with

𝒪(ϑ1ε2nlog2ΦloglogΦϑε){\mathcal{O}}\bigl{(}{\vartheta}^{-1}\varepsilon^{-2}n\log^{2}{\Phi}\log\frac{\log{\Phi}}{{\vartheta}\varepsilon}\bigr{)}

edges.

Proof:

By Corollary 4.8, one can build a (1+ε/2)(1+\varepsilon/2)-cover of {\mathcal{M}} of depth 𝖣=𝒪(ε1logΦ){\mathsf{D}}={\mathcal{O}}(\varepsilon^{-1}\log{\Phi}) and size 𝒪(n𝖣){\mathcal{O}}(n{\mathsf{D}}). Plugging this into Lemma 5.1 yields an oblivious ϑ{\vartheta}-reliable 22-hop (2+ε)(2+\varepsilon)-spanner for {\mathcal{M}}, of size

𝒪(n𝖣2ϑlog𝖣ϑ)=𝒪(ϑ1ε2nlog2ΦloglogΦϑε).{\mathcal{O}}\Bigl{(}\frac{n{\mathsf{D}}^{2}}{{\vartheta}}\log\frac{{\mathsf{D}}}{{\vartheta}}\Bigr{)}={\mathcal{O}}\Bigl{(}{\vartheta}^{-1}\varepsilon^{-2}n\log^{2}{\Phi}\log\frac{\log{\Phi}}{{\vartheta}\varepsilon}\Bigr{)}.

 

Lemma 5.6.

Let =(P,d){\mathcal{M}}=({P},{\mathrm{d}}) be an ultrametric over nn points with spread Φ{\Phi}, and let ϑ,ε(0,1){\vartheta},\varepsilon\in(0,1), and t𝐍t\in\mathbf{N} be parameters. Then, one can build a ϑ{\vartheta}-reliable ((2+ε)t1)((2+\varepsilon)t-1)-spanner for {\mathcal{M}} of size 𝒪(ϑ2ε3tn1+1/tlog3Φ).{\mathcal{O}}({\vartheta}^{-2}\varepsilon^{-3}t\cdot n^{1+1/t}\log^{3}{\Phi}).

Proof:

By Corollary 4.8, one can build a ξ=(1+ε/2){\xi}=(1+\varepsilon/2)-cover of {\mathcal{M}} of depth 𝖣=𝒪(ε1logΦ){\mathsf{D}}={\mathcal{O}}(\varepsilon^{-1}\log{\Phi}) and size 𝒪(n𝖣){\mathcal{O}}(n{\mathsf{D}}). Plugging this into Lemma 5.2 (A) yields a deterministic ϑ{\vartheta}-reliable (2t1)ξ(2t-1){\xi}-spanner for {\mathcal{M}}, of size

𝒪(ϑ2t𝖣3n1+1/t)=𝒪(ϑ2ε3tn1+1/tlog3Φ).{\mathcal{O}}({\vartheta}^{-2}t{\mathsf{D}}^{3}n^{1+1/t})={\mathcal{O}}({\vartheta}^{-2}\varepsilon^{-3}t\cdot n^{1+1/t}\log^{3}{\Phi}).

 

5.3.3 Tree metrics

Lemma 5.7.

Let =(P,d){\mathcal{M}}=({P},{\mathrm{d}}) be a tree metric over nn points with spread Φ{\Phi}, and let ϑ,ε(0,1){\vartheta},\varepsilon\in(0,1) be parameters. Then, one can build an oblivious ϑ{\vartheta}-reliable (3+ε)(3+\varepsilon)-spanner for {\mathcal{M}} with

𝒪(ϑ1ε2npolylog(n,Φ)){\mathcal{O}}\bigl{(}{\vartheta}^{-1}\varepsilon^{-2}n\mathop{\mathrm{polylog}}(n,{\Phi})\bigr{)}

edges, where polylog(n,Φ)=log2nlog2ΦloglogΦlognϑε\mathop{\mathrm{polylog}}(n,{\Phi})=\log^{2}n\log^{2}{\Phi}\log\frac{\log{\Phi}\log n}{{\vartheta}\varepsilon}.

Proof:

By Lemma 4.11, one can build a (2+ε/2)(2+\varepsilon/2)-cover of T{T} of depth 𝖣=𝒪(ε1logΦlogn){\mathsf{D}}={\mathcal{O}}(\varepsilon^{-1}\log{\Phi}\log n) and size 𝒪(n𝖣){\mathcal{O}}(n{\mathsf{D}}). Plugging this into Lemma 5.1 yields an oblivious ϑ{\vartheta}-reliable 22-hop (4+ε)(4+\varepsilon)-spanner for {\mathcal{M}}, of size

𝒪(n𝖣2ϑlog𝖣ϑ)=𝒪(ϑ1ε2nlog2nlog2ΦloglogΦlognϑε).{\mathcal{O}}\Bigl{(}\frac{n{\mathsf{D}}^{2}}{{\vartheta}}\log\frac{{\mathsf{D}}}{{\vartheta}}\Bigr{)}={\mathcal{O}}\Bigl{(}{\vartheta}^{-1}\varepsilon^{-2}n\log^{2}n\log^{2}{\Phi}\log\frac{\log{\Phi}\log n}{{\vartheta}\varepsilon}\Bigr{)}.

To get the improved bound on the dilation, let p,qP{p},{q}\in{P} be two points and let C𝒞{C}\in{\mathcal{C}} be the cluster such that p,qC{p},{q}\in{C} and diam(C)(2+ε)d(p,q)\mathrm{diam}\left({{C}}\right)\leq(2+\varepsilon){\mathrm{d}}\left({{p},{q}}\right). Assume that |C|>2|C|>2 (otherwise C={p,q}C=\{{p},{q}\} and there is nothing to prove). By construction, for the separator node 𝗌C{\mathsf{s}}\in{C}, we have d(𝗌,z)diam(C)/2{\mathrm{d}}\left({{\mathsf{s}},{z}}\right)\leq\mathrm{diam}\left({{C}}\right)/2 for all zC{z}\in{C}. Observe, that the (shortest) path between p{p} and q{q} in T{T} passes through 𝗌{\mathsf{s}}. Thus, using the triangle inequality, the length of a 22-hop path between p{p} and q{q} via zC{z}\in{C} can be bounded by

d(p,z)+d(z,q)d(p,𝗌)+2d(𝗌,z)+d(𝗌,q)d(p,q)+2diam(C)/2(3+ε)d(p,q).{\mathrm{d}}\left({{p},{z}}\right)+{\mathrm{d}}\left({{z},{q}}\right)\leq{\mathrm{d}}\left({{p},{\mathsf{s}}}\right)+2{\mathrm{d}}\left({{\mathsf{s}},{z}}\right)+{\mathrm{d}}\left({{\mathsf{s}},{q}}\right)\leq{\mathrm{d}}\left({{p},{q}}\right)+2\mathrm{diam}\left({{C}}\right)/2\leq(3+\varepsilon){\mathrm{d}}\left({{p},{q}}\right).

 

Lemma 5.8.

Let =(P,d){\mathcal{M}}=({P},{\mathrm{d}}) be a tree metric over nn points with spread Φ{\Phi}, and let ϑ,ε(0,1){\vartheta},\varepsilon\in(0,1), and t𝐍t\in\mathbf{N} be parameters. Then, one can build a ϑ{\vartheta}-reliable ((4+ε)t3)((4+\varepsilon)t-3)-spanner for {\mathcal{M}} of size 𝒪(ϑ2ε3n1+1/tlog3nlog3Φ).{\mathcal{O}}({\vartheta}^{-2}\varepsilon^{-3}\cdot n^{1+1/t}\log^{3}n\log^{3}{\Phi}).

Proof:

By Lemma 4.11, one can build a ξ=(2+ε/2){\xi}=(2+\varepsilon/2)-cover of T{T} of depth 𝖣=𝒪(ε1logΦlogn){\mathsf{D}}={\mathcal{O}}(\varepsilon^{-1}\log{\Phi}\log n) and size 𝒪(n𝖣){\mathcal{O}}(n{\mathsf{D}}). Plugging this into Lemma 5.2 (A) yields a deterministic ϑ{\vartheta}-reliable (2t1)ξ(2t-1){\xi}-spanner for {\mathcal{M}}, of size

𝒪(ϑ2𝖣3n1+1/t)=𝒪(ϑ2ε3n1+1/tlog3nlog3Φ).{\mathcal{O}}({\vartheta}^{-2}{\mathsf{D}}^{3}n^{1+1/t})={\mathcal{O}}({\vartheta}^{-2}\varepsilon^{-3}\cdot n^{1+1/t}\log^{3}n\log^{3}{\Phi}).

To get the improved bound on the dilation, consider a (2t1)(2t-1)-hop path p=p0,p1,,p2t1=q{p}={p}_{0},{p}_{1},\dots,{p}_{2t-1}={q}, such that piC{p}_{i}\in{C}, for all ii, for some cluster C{C}. We have

i=02t2d(pi,pi+1)=d(p,p1)+d(p2t2,q)+i=12t3d(pi,pi+1)d(p,𝗌)+d(𝗌,p1)+d(p2t2,𝗌)+d(𝗌,q)+(2t3)diam(C)d(p,q)+(2t2)diam(C)(1+ξ(2t2))d(p,q)\begin{split}\sum_{i=0}^{2t-2}{\mathrm{d}}\left({{p}_{i},{p}_{i+1}}\right)&={\mathrm{d}}\left({{p},{p}_{1}}\right)+{\mathrm{d}}\left({{p}_{2t-2},{q}}\right)+\sum_{i=1}^{2t-3}{\mathrm{d}}\left({{p}_{i},{p}_{i+1}}\right)\\ &\leq{\mathrm{d}}\left({{p},{\mathsf{s}}}\right)+{\mathrm{d}}\left({{\mathsf{s}},{p}_{1}}\right)+{\mathrm{d}}\left({{p}_{2t-2},{\mathsf{s}}}\right)+{\mathrm{d}}\left({{\mathsf{s}},{q}}\right)+(2t-3)\mathrm{diam}\left({{C}}\right)\\ &\leq{\mathrm{d}}\left({{p},{q}}\right)+(2t-2)\mathrm{diam}\left({{C}}\right)\leq(1+{\xi}(2t-2))\cdot{\mathrm{d}}\left({{p},{q}}\right)\end{split}

for the length of the path. Thus, by using 2+ε/22+\varepsilon/2 for the cover quality, we get

1+(2t2)(2+ε/2)=(4+ε)t3ε(4+ε)t3.1+(2t-2)(2+\varepsilon/2)=(4+\varepsilon)t-3-\varepsilon\leq(4+\varepsilon)t-3.

 

5.3.4 Planar graphs

Lemma 5.9.

Let G{G} be a weighted planar graph with nn vertices and spread Φ{\Phi}. Furthermore, let ϑ,ε(0,1){\vartheta},\varepsilon\in(0,1) be parameters. Then, one can build an oblivious ϑ{\vartheta}-reliable (3+ε)(3+\varepsilon)-spanner for G{G} with 𝒪(ϑ1ε4npolylog(n,Φ)){\mathcal{O}}\bigl{(}{\vartheta}^{-1}\varepsilon^{-4}n\mathop{\mathrm{polylog}}(n,{\Phi})\bigr{)} edges, where polylog(n,Φ)=log2nlog2ΦloglogΦlognϑε\mathop{\mathrm{polylog}}(n,{\Phi})=\log^{2}n\log^{2}{\Phi}\log\frac{\log{\Phi}\log n}{{\vartheta}\varepsilon}.

Proof:

By Theorem 4.17, one can build a (2+ε/2)(2+\varepsilon/2)-cover of G{G} of depth 𝖣=𝒪(ε2logΦlogn){\mathsf{D}}={\mathcal{O}}(\varepsilon^{-2}\log{\Phi}\log n) and size 𝒪(n𝖣){\mathcal{O}}(n{\mathsf{D}}). Plugging this into Lemma 5.1 yields an oblivious ϑ{\vartheta}-reliable 22-hop (4+ε)(4+\varepsilon)-spanner for G{G}, of size 𝒪(n𝖣2ϑlog𝖣ϑ)=𝒪(ϑ1ε4nlog2nlog2ΦloglogΦlognϑε).{\mathcal{O}}\Bigl{(}\frac{n{\mathsf{D}}^{2}}{{\vartheta}}\log\frac{{\mathsf{D}}}{{\vartheta}}\Bigr{)}={\mathcal{O}}\Bigl{(}{\vartheta}^{-1}\varepsilon^{-4}n\log^{2}n\log^{2}{\Phi}\log\frac{\log{\Phi}\log n}{{\vartheta}\varepsilon}\Bigr{)}.

We next show the improved bound on the dilation. Using the notation from Lemma 4.15, for a pair of points p,qP{p},{q}\in{P}, there is a cluster C𝒞{C}\in{\mathcal{C}}, such that p,qC{p},{q}\in{C} and diam(C)(2+ε)d(p,q)\mathrm{diam}\left({{C}}\right)\leq(2+\varepsilon){\mathrm{d}}\left({{p},{q}}\right). Let vv be the point where the cycle separator and the shortest path between p{p} and q{q} intersect. Notice, that for the center point uCu\in{C} we have d(u,z)diam(C)/2{\mathrm{d}}\left({u,{z}}\right)\leq\mathrm{diam}\left({{C}}\right)/2, for all zC{z}\in{C}. Furthermore, using the notation from the proof of Lemma 4.15, for some jj, we have

d(u,v)d(p,q)εrj/8rj1=ε8(1+ε8)ε4.\frac{{\mathrm{d}}\left({u,v}\right)}{{\mathrm{d}}\left({{p},{q}}\right)}\leq\frac{\varepsilon r_{j}/8}{r_{j-1}}=\frac{\varepsilon}{8}\left(1+\frac{\varepsilon}{8}\right)\leq\frac{\varepsilon}{4}.

Thus, the length of a 22-hop path between p{p} and q{q} via zC{z}\in{C} can be bounded by

d(p,z)+d(z,q)\displaystyle{\mathrm{d}}\left({{p},{z}}\right)+{\mathrm{d}}\left({{z},{q}}\right) d(p,v)+2d(v,u)+2d(u,z)+d(v,q)d(p,q)+2ε4d(p,q)+2diam(C)/2\displaystyle\leq{\mathrm{d}}\left({{p},v}\right)+2{\mathrm{d}}\left({v,u}\right)+2{\mathrm{d}}\left({u,{z}}\right)+{\mathrm{d}}\left({v,{q}}\right)\leq{\mathrm{d}}\left({{p},{q}}\right)+2\frac{\varepsilon}{4}{\mathrm{d}}\left({{p},{q}}\right)+2\mathrm{diam}\left({{C}}\right)/2
(1+ε2+2+ε)d(p,q)(3+2ε)d(p,q).\displaystyle\leq\left(1+\frac{\varepsilon}{2}+2+\varepsilon\right){\mathrm{d}}\left({{p},{q}}\right)\leq(3+2\varepsilon){\mathrm{d}}\left({{p},{q}}\right).

 

Lemma 5.10.

Let G{G} be a weighted planar graph with nn vertices and spread Φ{\Phi}. Furthermore, let ϑ,ε(0,1){\vartheta},\varepsilon\in(0,1) and t𝐍t\in\mathbf{N} be parameters. Then, one can build a deterministic ϑ{\vartheta}-reliable ((4+ε)t3)((4+\varepsilon)t-3)-spanner for G{G} of size 𝒪(ϑ2ε6n1+1/tlog3nlog3Φ).{\mathcal{O}}({\vartheta}^{-2}\varepsilon^{-6}\cdot n^{1+1/t}\log^{3}n\log^{3}{\Phi}).

Proof:

Let ξ=2+ε/2{\xi}=2+\varepsilon/2. By Theorem 4.17, one can build a ξ{\xi}-cover of G{G} with depth 𝖣=𝒪(ε2logΦlogn){\mathsf{D}}={\mathcal{O}}(\varepsilon^{-2}\log{\Phi}\log n) and size 𝒪(n𝖣){\mathcal{O}}(n{\mathsf{D}}). Plugging this into Lemma 5.2 (A) yields a deterministic ϑ{\vartheta}-reliable (2t1)ξ(2t-1){\xi}-spanner for {\mathcal{M}}, of size 𝒪(ϑ2𝖣3n1+1/t)=𝒪(ϑ2ε6n1+1/tlog3nlog3Φ).{\mathcal{O}}({\vartheta}^{-2}{\mathsf{D}}^{3}n^{1+1/t})={\mathcal{O}}({\vartheta}^{-2}\varepsilon^{-6}\cdot n^{1+1/t}\log^{3}n\log^{3}{\Phi}).

The improved dilation follows by using the same argument as in the proof of Lemma 5.9 and Lemma 5.8 

6 Proper expanders as reliable spanners for uniform metric

The purpose of this section is to prove that with the appropriate parameters, proper expanders (as defined in Definition 3.4) and edge-union of proper expanders constitute good reliable spanners for uniform metrics. That is, they satisfy the requirements of Theorem 3.6.

6.1 Preliminaries

In the following, (n,d)(n,d)-graph denotes a dd-regular, nn-vertex graph. Recall that λ(G)\lambda(G) denotes the normalized eigenvalue of G{G} – that is, the second largest in absolute value (see Definition 3.4).

For a given graph G=(V,E){G}=({V},{E}) and S,HVS,H\subseteq{V}, denote by ΓH(S)={vH|uS,uvE}\Gamma_{H}(S)=\left\{v\in H\;\middle|\;\exists u\in S,\;uv\in E\right\} the neighbors of SS in HH. For S,TVS,T\subset{V}, denote E(S,T)={uvE|uS,vT}{E}\left({S,T}\right)=\left\{uv\in E\;\middle|\;u\in S,\;v\in T\right\}.

The following is well known result on expanders, attributed to Alon and Chung [AC88] in [HLW06, Sec. 2.4].

Lemma 6.1 (Expander mixing lemma).

Let G=(V,E){G}=({V},{E}) be an (n,d)(n,d) graph. Then for every S,TVS,T\subset{V},

||E(S,T)|d|S||T|n|λ(G)d|S||T|.\left|{|{E}\left({S,T}\right)|-\frac{d|S|\,|T|}{n}}\right|\leq\lambda(G)d\sqrt{|S|\,|T|}. (6.1)

We also need the following lemma which is a minor variant of known constructions.

Lemma 6.2 ([BHO19]).

Let L,RL,R be two disjoint sets, with a total of n2n\in 2\mathbb{N} elements, and let ξ(0,1){\xi}\in(0,1) be a parameter. There exists a bipartite graph G=(LR,E){G}=(L\cup R,E) with 𝒪(n/ξ2){\mathcal{O}}(n/{\xi}^{2}) edges, such that

  1. (I)

    for any subset XLX\subseteq L, with |X|ξ|L|\left|{X}\right|\geq{\xi}|L|, we have that |Γ(X)|>(1ξ)|R|\left|{{\Gamma}\left({X}\right)}\right|>(1-{\xi})|R|, and

  2. (II)

    for any subset YRY\subseteq R, with |Y|ξ|R|\left|{Y}\right|\geq{\xi}|R|, we have that |Γ(Y)|>(1ξ)|L|\left|{{\Gamma}\left({Y}\right)}\right|>(1-{\xi})|L|.

Remark 6.3.

The randomized construction of Lemma 6.2 succeeds with probability 11/nO(1)1-1/n^{O(1)}. Since we use the construction below on sets that are polynomially large (i.e., n1/tn^{1/t}), one can use Lemma 6.2 constructively in this case (potentially losing an additional logt\log t factor). This also applies to the other expander constructions used in this paper. But while the randomized construction works with high probability, verifying it seems computationally intractable.

6.2 Construction of reliable spanners from proper expanders

Theorem 3.6 states the existence two different spanners and accordingly, we present two different graphs, 𝒢n,ϑ,2t1{\mathcal{G}}_{n,{\vartheta},2t-1} and 𝒢n,ϑ,2t{\mathcal{G}}_{n,{\vartheta},2t}.

We begin with 𝒢n,ϑ,2t{\mathcal{G}}_{n,{\vartheta},2t}. Recall the definition of proper expander (Definition 3.4) with parameter δ\delta. Fix n,tn,t\in\mathbb{N}, n>tn>t, and δ(0,1/4)\delta\in(0,1/4) such that

n1/t(cδδ)1.n^{1/t}\geq({\mathcalb{c}}_{\!\delta}\delta)^{-1}. (6.2)

The graph 𝒢n,ϑ,2t{\mathcal{G}}_{n,{\vartheta},2t} is defined to be an (n,d)(n,d)-graph that is a proper expander with δ=ϑ\delta={\vartheta}, and

d=max{2ϑ1n1/t,36C2ϑ3cϑ2}d=\left\lceil{\max\{2{\vartheta}^{-1}n^{1/t},36C^{2}{\vartheta}^{-3}{\mathcalb{c}}_{\!{\vartheta}}^{-2}\}}\right\rceil (6.3)

To define 𝒢n,ϑ,2t1{\mathcal{G}}_{n,{\vartheta},2t-1} we follow an idea we used slightly inferior construction in a preliminary version of this paper, see [HMO21, Section 3.3.1]: Let n=n11/tn^{\prime}=n^{1-1/t}. Partition the space to n1/tn^{1/t} subsets, A1,,An1/tA_{1},\ldots,A_{n^{1/t}}, each of size nn^{\prime}, and let t=t1t^{\prime}=t-1. For every AiA_{i} construct a copy of the graph 𝒢n,ϑ,2t{\mathcal{G}}_{n^{\prime},{\vartheta},2t^{\prime}} with AiA_{i} being the vertices. The degree in those graphs is d=𝒪(ϑ1n1/t)=𝒪(ϑ1n1/t)d^{\prime}={\mathcal{O}}({\vartheta}^{-1}{n^{\prime}}^{1/t^{\prime}})={\mathcal{O}}({\vartheta}^{-1}n^{1/t}).

In addition, for every iji\neq j connect AiA_{i} with AjA_{j} with a bipartite expander (AiAj,Eij)(A_{i}\cup A_{j},{E}_{ij}) according to Lemma 6.2, with ξ=ϑ{\xi}={\vartheta}. This increases the degree by 𝒪(ϑ2n1/t){\mathcal{O}}({\vartheta}^{-2}n^{1/t}). Thus, the total degree of 𝒢n,ϑ,2t1{\mathcal{G}}_{n,{\vartheta},2t-1} is 𝒪(ϑ2n1/t){\mathcal{O}}({\vartheta}^{-2}n^{1/t}).

6.3 Analysis of 𝒢n,ϑ,2t{\mathcal{G}}_{n,{\vartheta},2t}

For the rest of Section 6.3, let G=𝒢n,ϑ,2t{G}={\mathcal{G}}_{n,{\vartheta},2t}, and λ=λ(G)\lambda=\lambda({G}).

6.3.1 The shadow of a bad set

Let BV{{B}}\subset{V} an arbitrary subset, such that (1+5ϑ)|B|<n(1+5{\vartheta})|{{B}}|<n (otherwise, we can choose B+=V{{B^{+}}}={V} and there is nothing to prove). Choose

ε=(1+ϑ)(|B|/n+λ/ϑ)(P3)(1+ϑ)(|B|/n+C/dϑ)Eq.(6.3)(1+ϑ)|B|/n+cϑϑ,\varepsilon=(1+{\vartheta})(|{{B}}|/n+\lambda/\sqrt{{\vartheta}})\stackrel{{\scriptstyle\hyperref@@ii[item:p:3]{\ref*{item:p:3}{}}}}{{\leq}}(1+{\vartheta})(|{{B}}|/n+C/\sqrt{d{\vartheta}})\stackrel{{\scriptstyle\hyperref@@ii[equation:d]{Eq.~{}(\ref*{equation:d}{)}}}}{{\leq}}(1+{\vartheta})|{{B}}|/n+{\mathcalb{c}}_{\!{\vartheta}}{\vartheta}, (6.4)

so (recalling that δ=ϑ\delta={\vartheta}, and cδ=cϑ{\mathcalb{c}}_{\!\delta}={\mathcalb{c}}_{\!{\vartheta}})

1δε1ϑ|B|/nϑ|B|/ncϑϑ1ϑ(14ϑ)ϑcϑϑϑ.1-\delta-\varepsilon\geq 1-{\vartheta}-|{{B}}|/n-{\vartheta}|{{B}}|/n-{\mathcalb{c}}_{\!{\vartheta}}{\vartheta}\geq 1-{\vartheta}-(1-4{\vartheta})-{\vartheta}-{\mathcalb{c}}_{\!{\vartheta}}{\vartheta}\geq{\vartheta}. (6.5)

Define the “shadows of B{{B}}” as follows. Let S0={S}_{0}=\emptyset, and for i>0i>0 let

Si=Si1{uV(BSi1)||E(u,BSi1)|εd.}.{S}_{i}={S}_{i-1}\cup\left\{u\in{V}\setminus({{B}}\cup{S}_{i-1})\;\middle|\;|{E}\left({u,{{B}}\cup{S}_{i-1}}\right)|\geq\varepsilon d\bigr{.}\right\}. (6.6)

These are all the “bad” vertices that have a lot of neighbors inside the (growing) bad set BSi1{{B}}\cup{S}_{i-1}. Lastly, set S=iSi{S}=\cup_{i}{S}_{i} the limit of Si{S}_{i}, and B+=BS{{B^{+}}}={{B}}\cup{S}.

6.3.2 Bounding the size of the shadow

By the construction above of the damaged set B+{{B^{+}}}, for every uVB+u\in{V}\setminus{{B^{+}}},

|E(u,B+)|<εd.|{E}\left({u,{{B^{+}}}}\right)|<\varepsilon d. (6.7)
Claim 6.4.

|S|ϑ|B||{S}|\leq{\vartheta}|{{B}}|.

Proof:

The argument we use is similar to the one used in the proof of [BLMN05, Lemma 5.3]. Let Δi=SiSi1\Delta_{i}={S}_{i}\setminus{S}_{i-1}. We have that

εd|Si|\displaystyle\varepsilon d|{S}_{i}| εdj=1i|Δj|Eq.(6.6)j=1i|E(Δj,BSj1)|=|E(Si,BSi1)|\displaystyle\quad\leq\quad\varepsilon d\sum_{j=1}^{i}|\Delta_{j}|\stackrel{{\scriptstyle\hyperref@@ii[equation:S:i]{Eq.~{}(\ref*{equation:S:i}{)}}}}{{\leq}}\sum_{j=1}^{i}|{E}\left({\Delta_{j},{{B}}\cup{S}_{j-1}}\right)|=|{E}\left({{S}_{i},{{B}}\cup{S}_{i-1}}\right)|
Eq.(6.1)d(|B|+|Si1|)|Si|n+λd(|B|+|Si1|)|Si|.\displaystyle\stackrel{{\scriptstyle\hyperref@@ii[equation:mixing]{Eq.~{}(\ref*{equation:mixing}{)}}}}{{\leq}}\frac{d(|{{B}}|+|{S}_{i-1}|)|{S}_{i}|}{n}+\lambda d\sqrt{(|{{B}}|+|{S}_{i-1}|)|{S}_{i}|}.

This implies that |Si|(ε|B|+|Si1|n)λ(|B|+|Si1|)|Si|.|{S}_{i}|\left({\varepsilon-\frac{|{{B}}|+|{S}_{i-1}|}{n}}\right)\leq\lambda\sqrt{(|{{B}}|+|{S}_{i-1}|)|{S}_{i}|}. Squaring, and dividing by |Si||{S}_{i}|, we have

|Si|(ε|B|+|Si1|n)2λ2(|B|+|Si1|)λ2(|B|+|Si|),|{S}_{i}|\left(\varepsilon-\frac{|{{B}}|+|{S}_{i-1}|}{n}\right)^{2}\leq\lambda^{2}(|{{B}}|+|{S}_{i-1}|)\leq\lambda^{2}(|{{B}}|+|{S}_{i}|),

which implies, for ρi=(ε|B|+|Si1|n)2\rho_{i}=\left(\varepsilon-\frac{|{{B}}|+|{S}_{i-1}|}{n}\right)^{2}, that

|Si|λ2|B|ρiλ2.|{S}_{i}|\leq\frac{\lambda^{2}|{{B}}|}{\rho_{i}-\lambda^{2}}.

We claim that |Si|ϑ|B||{S}_{i}|\leq{\vartheta}|{{B}}| for every ii. Indeed, otherwise let ii the smallest index such that |Si|>ϑ|B||{S}_{i}|>{\vartheta}|{{B}}|. So |Si1|ϑ|B|<|Si||{S}_{i-1}|\leq{\vartheta}|{{B}}|<|{S}_{i}|. By definition, see Eq. (6.4), we have that ε=(1+ϑ)(|B|/n+λ/ϑ)\varepsilon=(1+{\vartheta})(|{{B}}|/n+\lambda/\sqrt{{\vartheta}}) and as such

ρi=(ε|B|+|Si1|n)2((1+ϑ)(|B|n+λϑ)(1+ϑ)|B|n)21+ϑϑλ2\rho_{i}=\left({\varepsilon-\frac{|{{B}}|+|{S}_{i-1}|}{n}}\right)^{2}\geq\left({(1+{\vartheta})\left({\frac{|{{B}}|}{n}+\frac{\lambda}{\sqrt{{\vartheta}}}}\right)-\frac{(1+{\vartheta})|{{B}}|}{n}}\right)^{2}\geq\frac{1+{\vartheta}}{{\vartheta}}\lambda^{2}

We thus have that

ϑ|B|<|Si|λ2|B|ρiλ2λ2|B|1+ϑϑλ2λ2=ϑ|B|,{\vartheta}|{{B}}|<|{S}_{i}|\leq\frac{\lambda^{2}|{{B}}|}{\rho_{i}-\lambda^{2}}\leq\frac{\lambda^{2}|{{B}}|}{\frac{1+{\vartheta}}{{\vartheta}}\lambda^{2}-\lambda^{2}}={\vartheta}|{{B}}|,

which is a contradiction.  

6.3.3 The expansion happens outside the bad set

Claim 6.5.

Let UVB+U\subseteq{V}\setminus{{B^{+}}}. If |U|cϑϑn11/t|U|\leq{\mathcalb{c}}_{\!{\vartheta}}{\vartheta}n^{1-1/t} then |ΓVB+(U)|n1/t|U||\Gamma_{{V}\setminus{{B^{+}}}}(U)|\geq n^{1/t}|U|.

Proof:

We have

|U|cϑϑn11/t=cϑnn1/t/ϑcϑnd/2cϑnd,|U|\leq{\mathcalb{c}}_{\!{\vartheta}}{\vartheta}n^{1-1/t}=\frac{{\mathcalb{c}}_{\!{\vartheta}}n}{n^{1/t}/{\vartheta}}\leq\frac{{\mathcalb{c}}_{\!{\vartheta}}n}{d/2}\leq\frac{{\mathcalb{c}}_{\!{\vartheta}}n}{d},

since d2n1/t/ϑd\geq 2n^{1/t}/{\vartheta} by Eq. (6.3). As such, by the expansion property (P2), |ΓV(U)|(1δ)d|U||\Gamma_{V}(U)|\geq(1-\delta)d|U| (as δ=ϑ\delta={\vartheta}). Furthermore

|ΓB+(U)||E(U,B+)|Eq.(6.7)εd|U|,|\Gamma_{{{B^{+}}}}(U)|\leq|{E}\left({U,{{B^{+}}}}\right)|\stackrel{{\scriptstyle\hyperref@@ii[equation:E:u:EBSet]{Eq.~{}(\ref*{equation:E:u:EBSet}{)}}}}{{\leq}}\varepsilon d|U|,

so

|ΓVB+(U)||ΓV(U)||ΓB+(U)|(1δε)d|U|Eq.(6.5)ϑd|U|Eq.(6.3)n1/t|U|.|\Gamma_{{V}\setminus{{B^{+}}}}(U)|\geq|\Gamma_{V}(U)|-|\Gamma_{{{B^{+}}}}(U)|\geq(1-\delta-\varepsilon)d|U|\stackrel{{\scriptstyle\hyperref@@ii[equation:1-delta-epsilon]{Eq.~{}(\ref*{equation:1-delta-epsilon}{)}}}}{{\geq}}{\vartheta}d|U|\stackrel{{\scriptstyle\hyperref@@ii[equation:d]{Eq.~{}(\ref*{equation:d}{)}}}}{{\geq}}n^{1/t}|U|.

 

Let

𝔹GB+(u,i)={vVB+|𝖽GB+(u,v)i},{\mathbb{B}}_{{G}\setminus{{B^{+}}}}(u,i)=\left\{v\in{V}\setminus{{B^{+}}}\;\middle|\;\mathsf{d}_{{G}\setminus{{B^{+}}}}(u,v)\leq i\right\},

the ball of radius ii around uu in the shortest path metric of the graph GB+{G}\setminus{{B^{+}}}. Define U0={u}U_{0}=\{u\}, and Ui=ΓVB+(Ui1)U_{i}=\Gamma_{{V}\setminus{{B^{+}}}}(U_{i-1}). Observe that 𝔹GB+(u,i)=j=0iUj{\mathbb{B}}_{{G}\setminus{{B^{+}}}}(u,i)=\bigcup_{j=0}^{i}U_{j}.

Claim 6.6.

|𝔹GB+(u,t1)|n11/t|{\mathbb{B}}_{{G}\setminus{{B^{+}}}}(u,t-1)|\geq n^{1-1/t}.

Proof:

Assume the contrary. Then |Ut1||𝔹GB+(u,t1)|<n11/t|U_{t-1}|\leq|{\mathbb{B}}_{{G}\setminus{{B^{+}}}}(u,t-1)|<n^{1-1/t}, which implies that there exists j<t1j<t-1 such that |Uj+1|<|Uj|n1/t|U_{j+1}|<|U_{j}|n^{1/t}. From Claim 6.5, this means that |Uj|>cϑϑn11/t|U_{j}|>{\mathcalb{c}}_{\!{\vartheta}}{\vartheta}n^{1-1/t}. So take KUjK\subset U_{j} such that |K|=cϑϑn11/t|K|={\mathcalb{c}}_{\!{\vartheta}}{\vartheta}n^{1-1/t}. Then again by Claim 6.5,

|𝔹GB+(u,t1)||ΓVB+(K)|cϑϑn11/tn1/tEq.(6.2)n11/t.|{\mathbb{B}}_{{G}\setminus{{B^{+}}}}(u,t-1)|\geq|\Gamma_{V\setminus{{B^{+}}}}(K)|\geq{\mathcalb{c}}_{\!{\vartheta}}{\vartheta}n^{1-1/t}n^{1/t}\stackrel{{\scriptstyle\hyperref@@ii[equation:n^1/t>theta]{Eq.~{}(\ref*{equation:n^1/t>theta}{)}}}}{{\geq}}n^{1-1/t}.

 

We summarize the relevant properties of 𝒢n,ϑ,2t{\mathcal{G}}_{n,{\vartheta},2t}.

Lemma 6.7.

The graph 𝒢n,ϑ,2t{\mathcal{G}}_{n,{\vartheta},2t} has the following properties. For any BV{{B}}\subset{V}, there exists a set B+{{B^{+}}}, BB+V{{B}}\subseteq{{B^{+}}}\subset{V}, |B+|(1+5ϑ)|B||{{B^{+}}}|\leq(1+5{\vartheta})|{{B}}|, such that for any uVB+u\in{V}\setminus{{B^{+}}}, we have

|𝔹GB+(u,t1)|\displaystyle|{\mathbb{B}}_{{G}\setminus{{B^{+}}}}(u,t-1)| n11/t,\displaystyle\geq n^{1-1/t}, (6.8)
ΓV(𝔹GB+(u,t1))\displaystyle\Gamma_{V}({\mathbb{B}}_{{G}\setminus{{B^{+}}}}(u,t-1)) (1ϑ)n,\displaystyle\geq(1-{\vartheta})n, (6.9)
|𝔹GB+(u,t)|\displaystyle|{\mathbb{B}}_{{G}\setminus{{B^{+}}}}(u,t)| ϑn.\displaystyle\geq{\vartheta}n. (6.10)

Proof:

Recall that when (1+5ϑ)|B|n(1+5{\vartheta})|B|\geq n, we can take B+=V{{B^{+}}}=V and there is nothing to prove. So assume from now on that (1+5ϑ)|B|<n(1+5{\vartheta})|B|<n.

When t=1t=1, 𝒢n,ϑ,2t{\mathcal{G}}_{n,{\vartheta},2t} is the complete graph, and we take B+=B{{B^{+}}}={{B}}. Eq. (6.8) is equiavlent to 111\geq 1; Eq. (6.9) follows since ΓV({u})=V\Gamma_{V}(\{u\})=V in complete graph; and Eq. (6.10) follows simply because |B|n/(1+5ϑ)(1ϑ)n|B|\leq n/(1+5{\vartheta})\leq(1-{\vartheta})n.

Assume next that t2t\geq 2. Eq. (6.8) is just a repetition of Claim 6.6. To prove Eq. (6.9). Observe that

|𝔹GB+(u,t1)|Eq.(6.8)n11/tEq.(6.3)2n/(ϑd).|{\mathbb{B}}_{{G}\setminus{{B^{+}}}}(u,t-1)|\stackrel{{\scriptstyle\hyperref@@ii[equation:n^(1-1/t)]{Eq.~{}(\ref*{equation:n^(1-1/t)}{)}}}}{{\geq}}n^{1-1/t}\stackrel{{\scriptstyle\hyperref@@ii[equation:d]{Eq.~{}(\ref*{equation:d}{)}}}}{{\geq}}2n/({\vartheta}d).

We conclude using Property (P1) of proper expanders that

ΓV(𝔹GB+(u,t1))(1ϑ)n.\Gamma_{V}({\mathbb{B}}_{{G}\setminus{{B^{+}}}}(u,t-1))\geq(1-{\vartheta})n.

By Claim 6.4,

|B+|(1+ϑ)|B|(1+ϑ)|n1+5ϑ(13ϑ)n.|{{B^{+}}}|\leq(1+{\vartheta})|{{B}}|\leq(1+{\vartheta})|\frac{n}{1+5{\vartheta}}\leq(1-3{\vartheta})n.

So,

|𝔹GB+(u,t)||ΓVB+(𝔹GB+(u,t1))||ΓV(Ut1)||B+|ϑn.|{\mathbb{B}}_{{G}\setminus{{B^{+}}}}(u,t)|\geq|\Gamma_{V\setminus{{B^{+}}}}({\mathbb{B}}_{{G}\setminus{{B^{+}}}}(u,t-1))|\geq|\Gamma_{V}(U_{t-1})|-|{{B^{+}}}|\geq{\vartheta}n.

This complete the proof of Lemma 6.7.  

Proposition 6.8.

The graph 𝒢n,ϑ,2t{\mathcal{G}}_{n,{\vartheta},2t} is a ϑ{\vartheta}-reliable 2t2t-spanner for nn-point uniform space with 𝒪(ϑ1n1+1/t){\mathcal{O}}({\vartheta}^{-1}n^{1+1/t}) edges.

Proof:

Fix u,vVB+u,v\in{V}\setminus{{B^{+}}}. By Eq. (6.9),

min{|ΓV(𝔹GB+(u,t1))|,|ΓV(𝔹GB+(v,t1))|}(1ϑ)n.\min\Bigl{\{}|\Gamma_{V}({\mathbb{B}}_{{G}\setminus{{B^{+}}}}(u,t-1))|,\,|\Gamma_{V}({\mathbb{B}}_{{G}\setminus{{B^{+}}}}(v,t-1))|\Bigr{\}}\geq(1-{\vartheta})n.

So,

|𝔹GB+(u,t)𝔹GB+(v,t)|\displaystyle|{\mathbb{B}}_{{G}\setminus{{B^{+}}}}(u,t)\cap{\mathbb{B}}_{{G}\setminus{{B^{+}}}}(v,t)| |ΓV(𝔹GB+(u,t1))ΓV(𝔹GB+(v,t1))||B+|\displaystyle\geq|\Gamma_{V}({\mathbb{B}}_{{G}\setminus{{B^{+}}}}(u,t-1))\cap\Gamma_{V}({\mathbb{B}}_{{G}\setminus{{B^{+}}}}(v,t-1))|-|{{B^{+}}}|
(12ϑ)n(15ϑ)n>0.\displaystyle\geq(1-2{\vartheta})n-(1-5{\vartheta})n>0.

We conclude that 𝔹GB+(u,t)𝔹GB+(v,t){\mathbb{B}}_{{G}\setminus{{B^{+}}}}(u,t)\cap{\mathbb{B}}_{{G}\setminus{{B^{+}}}}(v,t)\neq\emptyset, which means that there is a path of length at most 2t2t in GB+G\setminus{{B^{+}}} between uu and vv.  

6.3.4 Proof of Theorem 3.10

Restatement of Theorem 3.10. For every ϑ(0,1){\vartheta}\in(0,1) there exists a constant c=cϑ>0{\mathcalb{c}}_{\!}={\mathcalb{c}}_{\!{\vartheta}}>0, such that for any expansion constant hc2h\geq{\mathcalb{c}}_{\!}^{-2}, there exists a family of vertex expander graphs {G=(V,E)}n\{{G}=({V},{E})\}_{n} on nn vertices of degree at most 4h/ϑ4h/{\vartheta}, with the following resiliency property: For any BV{{B}}\subset{V}, there exists B+B{{B^{+}}}\supseteq{{B}}, |B+|(1+ϑ)|B||{{B^{+}}}|\leq(1+{\vartheta})|{{B}}| such that the graph GB+{G}\setminus{{B^{+}}} is a vertex expander in the sense that

  1. (i)

    diam(GB+)2loghn\mathrm{diam}\left({{G}\setminus{{B^{+}}}}\right)\leq 2\lceil\log_{h}n\rceil, and

  2. (ii)

    For any UVB+U\subset{V}\setminus{{B^{+}}} of size |U|cn/h|U|\leq{\mathcalb{c}}_{\!}n/h, we have |ΓGB+(U)|h|U||\Gamma_{{G}\setminus{{B^{+}}}}(U)|\geq h|U|.

Proof:

The graphs are simply 𝒢n,ϑ,2t{\mathcal{G}}_{n,{\vartheta},2t} for t=loghnt=\log_{h}n. The claims follow immediately from Claim 6.5 and Proposition 6.8.  

6.4 Analysis of 𝒢n,ϑ,2t1{\mathcal{G}}_{n,{\vartheta},2t-1}

Proposition 6.9.

The graph 𝒢n,ϑ,2t1{\mathcal{G}}_{n,{\vartheta},2t-1} is ϑ{\vartheta}-reliable (2t1)(2t-1)-spanner for nn-point uniform space with 𝒪(ϑ2n1+1/t){\mathcal{O}}({\vartheta}^{-2}n^{1+1/t}) edges.

Proof:

Recall the definition of 𝒢n,ϑ,2t1{\mathcal{G}}_{n,{\vartheta},2t-1} from Section 6.2. Let t=t1t^{\prime}=t-1, and n=n11/tn^{\prime}=n^{1-1/t}. Given an attack set B{{B}}, construct Bi+B^{+}_{i} from AiBA_{i}\cap{{B}} according to Lemma 6.7, and define B+=iBi+{{B^{+}}}=\cup_{i}B^{+}_{i}. Fix u,vVB+u,v\in{V}\setminus{{B^{+}}}. If u,vAiu,v\in A_{i} then by Proposition 6.8, there is a 2t=2t22t^{\prime}=2t-2 path between them.

If uAiu\in A_{i}, vAjv\in A_{j}, iji\neq j. Then by Eq. (6.10), |𝔹AiB+(u,t)|ϑn|{\mathbb{B}}_{A_{i}\setminus{{B^{+}}}}(u,t^{\prime})|\geq{\vartheta}n^{\prime}, and |𝔹AjB+(v,t)|ϑn|{\mathbb{B}}_{A_{j}\setminus{{B^{+}}}}(v,t^{\prime})|\geq{\vartheta}n^{\prime}. By Lemma 6.2, there is an edge in EijE_{ij} between 𝔹AiB+(u,t){\mathbb{B}}_{A_{i}\setminus{{B^{+}}}}(u,t^{\prime}) and 𝔹AjB+(v,t){\mathbb{B}}_{A_{j}\setminus{{B^{+}}}}(v,t^{\prime}). and hence a path of length 2t+1=2t12t^{\prime}+1=2t-1 in GB+G\setminus{{B^{+}}}.  

7 Concluding remarks and open problems

Subsequent work

Recently Filtser and Le [FL21] improved some of the results here. They obtained bounds that do not depend on the spread of the metrics in some cases, for the oblivious adversary case. They also obtained reliable spanners (in the oblivious adversary model) for trees with tight stretch of 22 and for planar graphs with tight stretch of 2+ε2+\varepsilon.

Tradeoffs in deterministic constructions for general spaces

Classical spanners are known to have an approximation–size trade-off for general nn-point metrics: To achieve Θ(t)\Theta(t) approximation it is sufficient and necessary to have n1+1/tn^{1+1/t} edges in the worst case. In contrast, for reliable spanners we were only able to show an upper bound on the trade-off, with no asymptotically matching lower bound: To achieve 𝒪(t2){\mathcal{O}}(t^{2}) approximation it is sufficient to have 𝒪~(n1+1/t)\widetilde{{\mathcal{O}}}(n^{1+1/t}) edges. Classically, the uniform metric is 22-approximated by a star graph with only n1n-1 edges. In contrast, we have shown here reliable spanners for uniform metric have approximation–size trade similar to the classical spanner for general metrics. The connection between the two problems is quite intriguing, and is worthy of further research.

The dependence of the size on the spread

The size of spanners constructed in this paper depends on the spread of the metric space. This is because of the reduction to uniform spaces via covers, in which the dependence on the spread is unavoidable in general. However, in some setting this dependence is avoidable. For example [BHO19, BHO20] achieves this for fixed dimensional Euclidean spaces, and [FL21] achieves it for doubling spaces, and general finite spaces in the oblivious adversary model. Getting spread-free bounds for the non-oblivious adversary is an interesting problem for further research.

Explicit constructions

To the best of our knowledge, there is no known polynomial time deterministic algorithm for constructing expanders with Property (P1) or Property (P2). Getting such a construction is an interesting open problem.

Acknowledgments

The authors thank Kevin Buchin for useful discussions and references, and the anonymous referees for their suggestions.

References

  • [AC88] N. Alon and F. R. K. Chung. Explicit construction of linear sized tolerant networks. In Proceedings of the First Japan Conference on Graph Theory and Applications (Hakone, 1986), volume 72, pages 15–19, 1988.
  • [AG06] Ittai Abraham and Cyril Gavoille. Object location using path separators. In Eric Ruppert and Dahlia Malkhi, editors, Proc. 25th ACM Sympos. Prin. Dist. Comput., pages 188–197, New York, NY, USA, 2006. ACM.
  • [AGG+19] Ittai Abraham, Cyril Gavoille, Anupam Gupta, Ofer Neiman, and Kunal Talwar. Cops, robbers, and threatening skeletons: padded decomposition for minor-free graphs. SIAM J. Comput., 48(3):1120–1145, 2019.
  • [AGMW10] Ittai Abraham, Cyril Gavoille, Dahlia Malkhi, and Udi Wieder. Strong-diameter decompositions of minor free graphs. Theory Comput. Syst., 47(4):837–855, 2010.
  • [AP90] Baruch Awerbuch and David Peleg. Sparse partitions. In Proc. 31st Annu. IEEE Sympos. Found. Comput. Sci. (FOCS), pages 503–513, USA, 1990. IEEE Computer Society.
  • [Bar98] Y. Bartal. On approximating arbitrary metrics by tree metrics. In Proc. 30th Annu. ACM Sympos. Theory Comput. (STOC), pages 161–168, New York, NY, USA, 1998. ACM.
  • [BCDM18] P. Bose, P. Carmi, V. Dujmovic, and P. Morin. Near-optimal O(k)-robust geometric spanners. CoRR, abs/1812.09913, 2018.
  • [BDMS13] P. Bose, V. Dujmović, P. Morin, and M. Smid. Robust geometric spanners. SIAM Journal on Computing, 42(4):1720–1736, 2013.
  • [BDPW18] Greg Bodwin, Michael Dinitz, Merav Parter, and Virginia Vassilevska Williams. Optimal Vertex Fault Tolerant Spanners (for fixed stretch), pages 1884–1900. SIAM, 3600 University City Science Center, Philadelphia, PA, USA, 2018.
  • [BHO19] Kevin Buchin, Sariel Har-Peled, and Dániel Oláh. A spanner for the day after. In Gill Barequet and Yusu Wang, editors, Proc. 35th Int. Annu. Sympos. Comput. Geom. (SoCG), volume 129 of LIPIcs, pages 19:1–19:15, internet, 2019. Schloss Dagstuhl - Leibniz-Zentrum fuer Informatik.
  • [BHO20] Kevin Buchin, Sariel Har-Peled, and Dániel Oláh. Sometimes reliable spanners of almost linear size. In Fabrizio Grandoni, Grzegorz Herman, and Peter Sanders, editors, Proc. 29th Annu. Euro. Sympos. Alg. (ESA), volume 173 of LIPIcs, pages 27:1–27:15, internet, 2020. Schloss Dagstuhl - Leibniz-Zentrum für Informatik.
  • [BLMN05] Yair Bartal, Nathan Linial, Manor Mendel, and Assaf Naor. On metric Ramsey-type phenomena. Ann. of Math. (2), 162(2):643–709, 2005.
  • [BLT14] Costas Busch, Ryan LaFortune, and Srikanta Tirthapura. Sparse covers for planar graphs and graphs that exclude a fixed minor. Algorithmica, 69(3):658–684, 2014.
  • [CLNS15] T.-H. H. Chan, M. Li, L. Ning, and S. Solomon. New doubling spanners: Better and simpler. SIAM Journal on Computing, 44(1):37–53, 2015.
  • [DJPP13] Ioana Dumitriu, Tobias Johnson, Soumik Pal, and Elliot Paquette. Functional limit theorems for random regular graphs. Probab. Theory Related Fields, 156(3-4):921–975, 2013.
  • [FKS89] Joel Friedman, Jeff Kahn, and Endre Szemerédi. On the second eigenvalue in random regular graphs. In David S. Johnson, editor, Proceedings of the 21st Annual ACM Symposium on Theory of Computing, May 14-17, 1989, Seattle, Washigton, USA, pages 587–598, New York, NY, USA, 1989. ACM.
  • [FL21] Arnold Filtser and Hung Le. Reliable spanners: Locality-sensitive orderings strike back. Arxiv CoRR, abs/2101.07428, 2021.
  • [Fri03] Joel Friedman. A proof of alon’s second eigenvalue conjecture. In Lawrence L. Larmore and Michel X. Goemans, editors, Proc. 35th Annu. ACM Sympos. Theory Comput. (STOC), pages 720–724, New York, NY, USA, 2003. ACM.
  • [FRT04] J. Fakcharoenphol, S. Rao, and K. Talwar. A tight bound on approximating arbitrary metrics by tree metrics. J. Comput. Sys. Sci., 69(3):485–497, 2004.
  • [HLW06] Shlomo Hoory, Nathan Linial, and Avi Wigderson. Expander graphs and their applications. Bull. Amer. Math. Soc. (N.S.), 43(4):439–561, 2006.
  • [HMO21] Sariel Har-Peled, Manor Mendel, and Dániel Oláh. Reliable spanners for metric spaces. In Kevin Buchin and Éric Colin de Verdière, editors, Proc. 37th Int. Annu. Sympos. Comput. Geom. (SoCG), volume 189 of LIPIcs, pages 43:1–43:13, internet, 2021. Schloss Dagstuhl - Leibniz-Zentrum für Informatik.
  • [KLMN05] R. Krauthgamer, J. R. Lee, M. Mendel, and A. Naor. Measured descent: A new embedding method for finite metric spaces. Geom. funct. anal. (GAFA), 15(4):839–858, 2005.
  • [LNS98] Christos Levcopoulos, Giri Narasimhan, and Michiel H. M. Smid. Efficient algorithms for constructing fault-tolerant geometric spanners. In Jeffrey Scott Vitter, editor, Proc. 30th Annu. ACM Sympos. Theory Comput. (STOC), pages 186–195, New York, NY, USA, 1998. ACM.
  • [LNS02] C. Levcopoulos, G. Narasimhan, and M. Smid. Improved algorithms for constructing fault-tolerant spanners. Algorithmica, 32(1):144–156, 2002.
  • [LT79] R. J. Lipton and R. E. Tarjan. A separator theorem for planar graphs. SIAM Journal on Applied Mathematics, 36(2):177–189, 1979.
  • [Luk99] T. Lukovszki. New results of fault tolerant geometric spanners. In Proc. 6th Int. Workshop Alg. and Data Struct., pages 193–204, Berlin, Heidelberg, 1999. Springer-Verlag.
  • [MN07] M. Mendel and A. Naor. Ramsey partitions and proximity data structures. J. Euro. Math. Soc., 009(2):253–275, 2007.
  • [NS07] G. Narasimhan and M. Smid. Geometric spanner networks. Cambridge University Press, New York, NY, USA, 2007.
  • [Pel00] D. Peleg. Distributed Computing: A Locality-Sensitive Approach. SIAM, 3600 University City Science Center, Philadelphia, PA, USA, 2000.
  • [PS89] D. Peleg and A. Schäffer. Graph spanners. J. Graph Theory, 13:99–116, 1989.
  • [Pud15] Doron Puder. Expansion of random graphs: new proofs, new results. Invent. Math., 201(3):845–908, 2015.
  • [Sol14] S. Solomon. From hierarchical partitions to hierarchical covers: Optimal fault-tolerant spanners for doubling metrics. In Proceedings of the 46th ACM Symposium on Theory of Computing, STOC ’14, page 363–372, New York, NY, USA, 2014. ACM.
  • [Wor99] N. C. Wormald. Models of random regular graphs. In Surveys in combinatorics, 1999 (Canterbury), volume 267 of London Math. Soc. Lecture Note Ser., pages 239–298. Cambridge Univ. Press, Cambridge, UK, 1999.

Appendix A Random regular graphs

In Theorem 3.5 above we prove that random construction using a union of random permutations yields a proper expander, see Definition 3.4. The properties are well known and are held by random regular graphs asymptotically almost surely. However, the literature on random regular graphs is more concerned with the setting of constant dd and nn tends to infinity, especially regarding Property (P3)). Here we need a slightly different range, where dd is sufficiently large, and nd2n\geq d^{2}. For completeness, we gather here proofs and references that prove Theorem 3.5.

A.1 Construction

We use the permutation model 𝒢n,d{\mathcal{G}_{n,d}} for constructing random regular graph (see [Wor99] for other models). Assuming dd is even, sample d/2d/2 independent and identically distributed random permutations π1,,πd/2Sn\pi_{1},\ldots,\pi_{d/2}\in S_{n}, where SnS_{n} is the set of all permutations of [n]={1,,n}\left[n\right]=\{1,\ldots,n\}. The resulting graph G=(V,E){G}=({V},{E}), has V=[n]{V}=\left[n\right] and

E={{i,πj(i)}|i[n] and j[d/2].}.{E}=\left\{\{i,\pi_{j}(i)\}\;\middle|\;i\in\left[n\right]\text{ and }j\in\left[d/2\right]\bigr{.}\right\}.

A.2 Analysis

A.2.1 Property (P3)

All the proofs that random dd-regular graphs have second eigenvalue at most C/dC/\sqrt{d} (that we are aware of) are non-trivial. It was first proved in [FKS89], and by now there are many proofs of this. Notably, Friedman [Fri03] showed that random regular graphs are “almost Ramanujan”, i.e., λ2d1+εd\lambda\leq\frac{2\sqrt{d-1}+\varepsilon}{d} with probability 1on(1)1-o_{n}(1), see also [Pud15] for a recent and simpler proof. Unfortunately, most of those papers are interested in the settings where the degree dd is constant and only the number of vertices nn\to\infty, which is not suitable for our needs. However, the argument of Kahn and Szemerédi [FKS89] does work when dd is allowed to tend to infinity (together with nn). Specifically, Dumitiriu et al. [DJPP13] used it to prove the following.

Theorem A.1 ([DJPP13, Theorem 24]).

Fix C=41000C=41000. There exists K>0K>0 such that for any even dd and n>max{K,d}n>\max\{K,d\}, we have [λ(G)C/d]12/n2,\displaystyle\mathop{{\mathbb{P}}}\left[\lambda({G})\leq C/\sqrt{d}\,\right]\geq 1-2/n^{2}, where []\mathop{{\mathbb{P}}}\left[\right] is the probability in the permutation model 𝒢n,d{\mathcal{G}_{n,d}}.

This implies (P3) holds a.a.s. in the permutation model. Properties (P1) and (P2) are much easier to prove and are considered folklore. They are usually proved in different, more convenient, models of random regular graphs. However, contrary to the case when the degree is fixed, in the high-degree regime the different random models are not necessarily contiguous, i.e., asymptotically almost surely properties are not necessarily equivalent among the different models (for more on this, see [Wor99]). Therefore, for completeness, we next provide proofs that Properties (P1) and (P2) hold asymptotically almost surely in the permutation model.

A.2.2 Property (P1)

Lemma A.2.

For integers s,t,n0s,t,n\geq 0, such that stns\leq t\leq n, we have (ts)/(ns)(tn)s.\binom{t}{s}/\binom{n}{s}\leq\left({\frac{t}{n}}\right)^{s}.

Similarly, if stns\leq t\leq n and n3sn\geq 3s, we have (ts)/(n2s)(tn)s(2sns)s..\binom{t}{s}/\binom{n}{2s}\leq\left({\frac{t}{n}}\right)^{s}\bigl{(}{\frac{2s}{n-s}}\bigr{)}^{s}.\Bigr{.}

Proof:

Observe that for kmmk\leq m^{\prime}\leq m, we have

(mk)(mk)=mmm1m1mk+1mk+1(mm)k,\frac{\binom{m^{\prime}}{k}}{\binom{m}{k}}=\frac{m^{\prime}}{m}\cdot\frac{m^{\prime}-1}{m-1}\cdots\frac{m^{\prime}-k+1}{m-k+1}\leq\left({\frac{m^{\prime}}{m}}\right)^{k},

as (mi)/(mi)m/m(m^{\prime}-i)/(m-i)\leq m^{\prime}/m. As such, using the (easy to verify) identity (n2s)=(ns)(nss)/(2ss),\binom{n}{2s}=\binom{n}{s}\binom{n-s}{s}/\binom{2s}{s}, we have

(ts)(n2s)=(ts)(2ss)(ns)(nss)(tn)s(2sns)s.\frac{\binom{t}{s}}{\binom{n}{2s}}=\frac{\binom{t}{s}\binom{2s}{s}}{\binom{n}{s}\binom{n-s}{s}}\leq\left({\frac{t}{n}}\right)^{s}\left({\frac{2s}{n-s}}\right)^{s}.

 

Lemma A.3.

Fix sets S,TVS,T\subset{V}, with s=|S|s=|S| and t=|T|t=|T|, such that |S||T||S|\leq|T|. Then [.Γ(S)T](tn)sd/2.\mathop{{\mathbb{P}}}\left[\bigl{.}\Gamma(S)\subseteq T\right]\leq\left({\frac{t}{n}}\right)^{sd/2}.

Proof:

Fix TTT^{\prime}\subseteq T of cardinality |T|=s|T^{\prime}|=s. Then [π(S)=T]=1/(ns).\mathop{{\mathbb{P}}}\left[\pi(S)=T^{\prime}\right]=1/\binom{n}{s}. By the union bound, and Lemma A.2, we have

[.π(S)T]=T(Ts)[π(S)=T.]=(ts)(ns)(tn)s.\mathop{{\mathbb{P}}}\left[\bigl{.}\pi(S)\subseteq T\right]=\sum_{{T^{\prime}\in\binom{T}{s}}}\mathop{{\mathbb{P}}}\left[\pi(S)=T^{\prime}\bigr{.}\right]=\frac{\binom{t}{s}}{\binom{n}{s}}\leq\left(\frac{t}{n}\right)^{s}.

Let π1,,πd/2\pi_{1},\ldots,\pi_{d/2} be the permutations used in the construction. We have that

[.Γ(S)T][k=1d/2πk(S)T]=([π(S)T].)d/2(tn)sd/2.\mathop{{\mathbb{P}}}\left[\bigl{.}\Gamma(S)\subseteq T\right]\leq{\mathbb{P}}\Bigl{[}\,\bigwedge_{k=1}^{d/2}{\pi_{k}}(S)\subseteq T\Bigr{]}=\left({\mathop{{\mathbb{P}}}\left[{\pi}(S)\subseteq T\right]\Bigr{.}}\right)^{d/2}\leq\left(\frac{t}{n}\right)^{sd/2}.

 

Lemma A.4.

Fix ε,δ(0,1/2)\varepsilon,\delta\in(0,1/2). If nd12/(εδ)\sqrt{n}\geq d\geq 12/(\varepsilon\delta) is an integer. Then, asymptotically almost surely over the random (n,d)(n,d)-graph G=(V,E)G=(V,E), for any SVS\subset{V} with |S|εn|S|\geq\varepsilon n, we have |Γ(S)|(1δ)n|\Gamma(S)|\geq(1-\delta)n. That is, property (P1) of Definition 3.4 holds.

Proof:

Using the union bound on the bound established in Lemma A.3 we conclude that the probability that there exists a subset SS of size s=εns=\varepsilon n, such that |Γ(S)|(1δ)n=t|\Gamma(S)|\leq(1-\delta)n=t, is at most

(ns)(nt)(tn)sd/2\displaystyle\binom{n}{s}\binom{n}{t}\left(\frac{t}{n}\right)^{sd/2} (ens)s2n(tn)sd/2(eε)εn2n(1δ)εnd/2(eε)εn2nexp(δεnd/2)\displaystyle\leq\left({\frac{en}{s}}\right)^{s}2^{n}\left(\frac{t}{n}\right)^{sd/2}\leq\left({\frac{e}{\varepsilon}}\right)^{\varepsilon n}2^{n}(1-\delta)^{\varepsilon nd/2}\leq\left({\frac{e}{\varepsilon}}\right)^{\varepsilon n}2^{n}\exp\left({-\delta\varepsilon nd/2}\right)
2nexp(εn+εnln1ε6n)exp(2n+n/e6n)exp(3n),\displaystyle\leq 2^{n}\exp\left({\varepsilon n+\varepsilon n\ln\frac{1}{\varepsilon}-6n}\right)\leq\exp\left({2n+n/e-6n}\right)\leq\exp(-3n),

since maxx>0xln(1/x)=1/e\max_{x>0}x\ln(1/x)=1/e, as easy calculation shows111Indeed, for f(x)=xln(1/x)f(x)=x\ln(1/x), we have f(x)=1+lnx1f^{\prime}(x)=-1+\ln x^{-1}. Setting f(x)=0f^{\prime}(x)=0, implies 1=lnx11=\ln x^{-1}. Namely, the maximum of ff is achieved at x=1/ex=1/e, where f(1/e)=1/ef(1/e)=1/e..  

A.2.3 Property (P2)

The argument above used only the forward edges associated with the d/2d/2 random permutations. Since there are only d/2d/2 such edges associated with every vertex, that argument can not prove vertex-expansion close to dd as is stated in (P2). For this we have to also use the backward edges associated with the permutation as well. Since the backward edges and the forward edges associated with the same random permutation are not stochastically independent, more care is needed to prove this property, as we shall now see.

Claim A.5.

Fix ε,η(0,1)\varepsilon,\eta\in(0,1) such that 0<ε<13d<3d<η<10<\varepsilon<\frac{1}{3d}<\frac{3}{d}<\eta<1. Then

[.SV,|S|εn,|E(S,S)|>ηd|S|]=O(n0.4).\mathop{{\mathbb{P}}}\left[\Bigl{.}\exists S\subset{V},|S|\leq\varepsilon n,|{E}\left({S,S}\right)|>\eta d|S|\right]=O(n^{-0.4}).

Proof:

Fix SVS\subset{V}, of cardinality sεns\leq\varepsilon n. Observe that

E(S,S)={{u,πi(u)}|uS,πi(u)S, and i[d/2].}.E(S,S)=\left\{\{u,\pi_{i}(u)\}\;\middle|\;u\in S,\pi_{i}(u)\in S,\text{ and }i\in\left[d/2\right]\bigr{.}\right\}.

Fix RS×[d/2]R\subseteq S\times\left[d/2\right]. For any (u,i)R(u,i)\in R, let u,i{\mathcal{E}}_{u,i} be the event that πi(u)S\pi_{i}(u)\in S. The events induced by RR, that is {u,i|(u,i)R}\left\{{\mathcal{E}}_{u,i}\;\middle|\;(u,i)\in R\right\} are not independent, but they are non-positively correlated. Indeed, for a set RS×[d/2]R\subset S\times\left[d/2\right] consider the event R=(z,j)Rz,j{\mathcal{E}}_{R}=\bigwedge_{(z,j)\in R}{\mathcal{E}}_{z,j}. For (u,i)R(u,i)\notin R we have,

[πi(u)S|R][πi(u)S.]=sn.\mathop{{\mathbb{P}}}\left[\pi_{i}(u)\in S\;\big{|}\;{\mathcal{E}}_{R}\right]\leq\mathop{{\mathbb{P}}}\left[\pi_{i}(u)\in S\bigr{.}\right]=\frac{s}{n}.

Indeed, observe that all the sub-events in R{\mathcal{E}}_{R} involving jij\neq i, are irrelevant (that is, stochastically independent of the events x,i{\mathcal{E}}_{x,i}). As such, let X={x|(x,i)R}X=\left\{x\;\middle|\;(x,i)\in R\right\}. Imagine now picking the permutation πi\pi_{i} by first picking the locations of the elements of XX, and then choosing the location of uu. Clearly, if the event R{\mathcal{E}}_{R} happened, then there are only s|X|s-|X| empty slots in SS, and n|X|n-|X| slots available overall. As such, we have that

[πi(u)S|R]s|X|n|X|sn=[πi(u)S.].\mathop{{\mathbb{P}}}\!\left[\pi_{i}(u)\in S\;\middle|\;{\mathcal{E}}_{R}\right]\leq\frac{s-|X|}{n-|X|}\leq\frac{s}{n}=\mathop{{\mathbb{P}}}\left[\pi_{i}(u)\in S\bigr{.}\right].

Now, order the elements of RR in arbitrary order, and let R(i)R(i) be the subset with the first ii elements in that order. We have

[R.]=i=1|R|[R(i)|R(i1)](s/n)|R|.\mathop{{\mathbb{P}}}\left[{\mathcal{E}}_{R}\bigl{.}\right]=\prod_{i=1}^{|R|}\mathop{{\mathbb{P}}}\!\left[{\mathcal{E}}_{R(i)}\;\middle|\;{\mathcal{E}}_{R(i-1)}\right]\leq(s/n)^{|R|}.

Therefore,

[.|E(S,S)|>ηds]\displaystyle\mathop{{\mathbb{P}}}\left[\bigl{.}|E(S,S)|>\eta ds\right] r=ηds+1ds/2R(S×[d/2]r)[R.]r=ηdsds/2(ds/2r)(sn)rr=ηdsds/2(eds/2rsn)r\displaystyle\leq\sum_{r=\eta ds+1}^{ds/2}\sum_{R\in\binom{S\times\left[d/2\right]}{r}}\mathop{{\mathbb{P}}}\left[{\mathcal{E}}_{R}\bigr{.}\right]\leq\sum_{r=\eta ds}^{ds/2}\binom{ds/2}{r}\cdot\left({\frac{s}{n}}\right)^{r}\leq\sum_{r=\eta ds}^{ds/2}\left({\frac{eds/2}{r}\cdot\frac{s}{n}}\right)^{r}
r=ηdsds/2(eds22rn)r(esηn)ηds,\displaystyle\leq\sum_{r=\eta ds}^{ds/2}\left({\frac{eds^{2}}{2rn}}\right)^{r}\leq\left(\frac{es}{\eta n}\right)^{\eta ds},

since the summations behaves like a geometric series, using eds22rneds22ηdsn=es2ηn12.\frac{eds^{2}}{2rn}\leq\frac{eds^{2}}{2\eta dsn}=\frac{es}{2\eta n}\leq\frac{1}{2}. Therefore

Pr[SV,|S|εn,|E(S,S)|>ηds]s=1εn(ns)(esηn)ηdss=1εn(e2(sn)ηd1ηηd)s=O(n0.4).\displaystyle\Pr\bigl{[}\exists S\subset{V},|S|\leq\varepsilon n,|E(S,S)|>\eta ds\bigr{]}\leq\sum_{s=1}^{\varepsilon n}\binom{n}{s}\left(\frac{es}{\eta n}\right)^{\eta ds}\leq\sum_{s=1}^{\varepsilon n}\left(e^{2}\left(\frac{s}{n}\right)^{\eta d-1}\eta^{-\eta d}\right)^{s}=O(n^{-0.4}).

The last inequality is derived by partitioning the sum on ss up-to 2logn2\log n – this part is bounded by O(log3/2nn).O\bigl{(}{\frac{\log^{3/2}n}{\sqrt{n}}}\bigr{)}. The second part of the summation from 2logn2\log n to εn\varepsilon n be bounded by O((ε/η)2logn)O\bigl{(}(\varepsilon/\eta)^{2\log n}\bigr{)}.  

Lemma A.6.

For every δ(0,1/4)\delta\in(0,1/4), there exists cδ>0c_{\delta}>0, such that for any sufficiently large integer de32/δd\geq e^{32/\delta}, and any nd2n\geq d^{2}, asymptotically (in nn) almost surely, a random (n,d)(n,d) graph G=(V,E)𝒢n,dG=(V,E)\sim{\mathcal{G}_{n,d}} have the following vertex expansion property: For any SVS\subset{V} with |S|e2.2/δn/d|S|\leq e^{-2.2/\delta}n/d, we have |Γ(S)|(1δ)d|S||\Gamma(S)|\geq(1-\delta)d|S|.

That is, Property (P2) of Definition 3.4 holds.

Proof:

Fix η=δ/2\eta=\delta/2. Define \mathcal{B} to be the event, that for all subsets SVS\subset{V} of cardinality at most εn\varepsilon n, we have that

|E(S,S)|ηd|S|.|E(S,S)|\leq\eta d|S|.

By Claim A.5, Pr[]1O(n0.4)\Pr[\mathcal{B}]\geq 1-O(n^{-0.4}). So, fix SVS\subset{V} of cardinality sεns\leq\varepsilon n. The following stochastic process samples a random bijection π\pi on VV uniformly.

  1. (i)

    Order the vertices of SS arbitrarily v1,,vsv_{1},\ldots,v_{s}.

  2. (ii)

    For i=1,,si=1,\ldots,s, pick π(vi)\pi(v_{i}) uniformly at random from V{π(v1),,π(vi1)}V\setminus\{\pi(v_{1}),\ldots,\pi(v_{i-1})\}.

  3. (iii)

    Let U=Sπ(S)U=S\cap\pi(S). Reorder the elements of SS as u1,,usu_{1},\ldots,u_{s}, such that U={u+1,,us}U=\{u_{\ell+1},\ldots,u_{s}\}.

  4. (iv)

    For i=1,,i=1,\ldots,\ell, pick π1(ui)\pi^{-1}(u_{i}) uniformly at random from V(S{π1(u1),,π1(ui1)})V\setminus(S\cup\{\pi^{-1}(u_{1}),\ldots,\pi^{-1}(u_{i-1})\}).

  5. (v)

    Denote by π\pi^{\prime} the partial injection constructed so far.

  6. (vi)

    Note that, for i=+1,,si=\ell+1,\ldots,s, the element π1(ui)\pi^{-1}(u_{i}) is already fixed, and it is in SS, as such we do nothing.

  7. (vii)

    Complete π\pi^{\prime} to a full permutation by sampling a random bijection (V(Sπ(S)))(V(Sπ(S)))(V\setminus(S\cup\pi^{\prime}(S)))\mapsto(V\setminus(S\cup\pi^{\prime}(S)))

One can sample random regular graph G{G} from 𝒢n,d{\mathcal{G}_{n,d}} by applying the above independently d/2d/2 times, and constructing d/2d/2 random permutations π1,πd/2\pi_{1},\ldots\pi_{d/2}.

We next do for a fixed SVS\subset{V} of cardinality sεns\leq\varepsilon n the following thought experiment. In the construction of each of the d/2d/2 permutation, we replace step (vi) above with the following step generating a new partial injection τ\tau into VV, as follows:

  1.   (vi)

    For i=+1,,si=\ell+1,\ldots,s, pick τ1(ui)\tau^{-1}(u_{i}) randomly and uniformly from

    V(S{π1(u1),,π1(u),τ1(u+1),,τ1(ui1)}).{V}\setminus(S\cup\{\pi^{-1}(u_{1}),\ldots,\pi^{-1}(u_{\ell}),\tau^{-1}(u_{\ell+1}),\ldots,\tau^{-1}(u_{i-1})\}).

Denote by τi=τiS\tau_{i}=\tau_{i}^{S} that random partial injection. Denote

FS={.{u,τi1(u)}|i[d/2],uSπi(S)}F_{S}=\left\{\bigr{.}\smash{\{u,{\tau^{-1}_{i}(u)}\}}\;\middle|\;i\in\left[d/2\right],u\in S\cap\pi_{i}(S)\right\}

the set of auxiliary edges associated with the above process for SS. Denote GS=(V,EFS){G_{S}}=(V,E\cup F_{S}) – it is the result of redirecting all the edges of E(S,S)E(S,S) to be outside the S×SS\times S biclique. The graph GS{G_{S}} is randomly generated, but the content of FSF_{S} is not measurable in 𝒢n,d{\mathcal{G}_{n,d}} – meaning that the underline probability spaces is more refined than 𝒢n,d{\mathcal{G}_{n,d}}. We denote by 𝒢n,d~{\widetilde{\mathcal{G}_{n,d}}} a probability space from which one can sample all of GS{G_{S}}. I.e., it contains d/2d/2 random permutations π1,,πd/2\pi_{1},\ldots,\pi_{d/2} that constitute 𝒢n,d{\mathcal{G}_{n,d}} and all the relevant partial injections τiS\tau^{S}_{i}.

Note that |FS|=|E(S,S)||F_{S}|=|E(S,S)|, and therefore depends only on 𝒢n,d{\mathcal{G}_{n,d}}. Furthermore, If G{G}\in\mathcal{B}, then for every SVS\subset{V}, with cardinality at most εn\varepsilon n, we have that |FS|ηd|S||F_{S}|\leq\eta d|S|. Hence, conditioned on \mathcal{B}, for any such SS,

|ΓG(S)||ΓGS(S)|ηd|S|.|\Gamma_{G}(S)|\geq|\Gamma_{{G_{S}}}(S)|-\eta d|S|. (A.1)

We next bound for a fixed SVS\subset{V} of cardinality sεns\leq\varepsilon n, the probability that there exists a subset TT of cardinality tt for which ΓG(S)T\Gamma_{G}(S)\subseteq T. Begin by fixing STVS\subseteq T\subseteq{V} of cardinality |T|=t|T|=t. We claim that

[.π(S)π(S)1τ1(Sπ(S))T](tn)2s.\mathop{{\mathbb{P}}}\left[\Bigl{.}\pi(S)\cup\pi{}^{-1}(S)\cup\tau^{-1}(S\cap\pi(S))\subseteq T\right]\leq\left({\frac{t}{n}}\right)^{2s}. (A.2)

Indeed, fix TTT^{\prime}\subset T of cardinality ss. Then

[.π(S)=T](ns)1.\mathop{{\mathbb{P}}}\left[\Bigl{.}\pi(S)=T^{\prime}\right]\leq\binom{n}{s}^{-1}.

Fix T′′VST^{\prime\prime}\subset{V}\setminus S of cardinality ss. Recall that we denoted by ππ\pi^{\prime}\subset\pi the partial injection sampled at the end of step (v) in the sampling process described above. Observe that conditioned on π(S)=T\pi(S)=T^{\prime} the partial injection π1τ1:SVS\pi^{\prime-1}\cup\tau^{-1}:S\to{V}\setminus S is also uniformly random, and therefore

[(π1τ1)(S)=T′′|π(S)=T](nss)1.\mathop{{\mathbb{P}}}\!\left[(\pi^{\prime}{}^{-1}\cup\tau^{-1})(S)=T^{\prime\prime}\;\middle|\;\pi(S)=T^{\prime}\right]\leq\binom{n-s}{s}^{-1}.

Hence,

[π(S)π1(S)τ1(Sπ(S))T]\displaystyle\mathop{{\mathbb{P}}}\left[\pi(S)\cup\pi^{-1}(S)\cup\tau^{-1}(S\cap\pi(S))\subseteq T\right]
=[.TT,|T|=sT′′TT,|T′′|=sπ(S)=T(π1τ1)(S)=T′′]\displaystyle=\mathop{{\mathbb{P}}}\left[\Bigl{.}\smash{\bigvee\nolimits_{{T^{\prime}\subset T,\,|T^{\prime}|=s}}\;\bigvee\nolimits_{{T^{\prime\prime}\subset T\setminus T^{\prime},\,|T^{\prime\prime}|=s}}}\pi(S)=T^{\prime}\wedge(\pi^{\prime}{}^{-1}\cup\tau^{-1})(S)=T^{\prime\prime}\right]
TT,|T|=sT′′TT,|T′′|=sPr[π(S)=T]Pr[(π1τ1)(S)=T′′|π(S)=T]\displaystyle\leq\sum\nolimits_{{T^{\prime}\subset T,\,|T^{\prime}|=s}}\;\sum\nolimits_{T^{\prime\prime}\subset T\setminus T^{\prime},\,|T^{\prime\prime}|=s}\Pr\bigl{[}\pi(S)=T^{\prime}\bigr{]}\cdot\Pr\bigl{[}(\pi^{\prime}{}^{-1}\cup\tau^{-1})(S)=T^{\prime\prime}\,\big{|}\,\pi(S)=T^{\prime}\bigr{]}
=(ts)(tss)(ns)(nss)(tn)s(tsns)s(tn)2s,\displaystyle=\frac{\binom{t}{s}\binom{t-s}{s}}{\binom{n}{s}\binom{n-s}{s}}\leq\left(\frac{t}{n}\right)^{s}\left(\frac{t-s}{n-s}\right)^{s}\leq\left(\frac{t}{n}\right)^{2s},

which complete the proof of Eq. (A.2).

The probability on {GS}SV𝒢n,d~\{{G_{S}}\}_{S\subset{V}}\sim{\widetilde{\mathcal{G}_{n,d}}} that there exists SVS\subset{V} of cardinality at most εn\varepsilon n for which |ΓGS(S)|t|\Gamma_{{G_{S}}}(S)|\leq t, where t=(1η)dst=(1-\eta)ds is at most the probability that there exists T~VS{\widetilde{T}}\subset{V}\setminus S of cardinality tt for which ΓGS(S)T~S\Gamma_{{G_{S}}}(S)\subseteq{\widetilde{T}}\cup S. By the union bound and Eq. (A.2) this is at most

s=1εn(ns)(nst)((t+sn)2d)d/2.\sum_{s=1}^{\varepsilon n}\binom{n}{s}\binom{n-s}{t}\left({\left({\frac{t+s}{n}}\right)^{2d}}\right)^{d/2}.

We bound the summand above by

(ns)(nst)(t+sn)sd\displaystyle\binom{n}{s}\binom{n-s}{t}\left(\frac{t+s}{n}\right)^{sd} (ns)(nt)(t+sn)sd\displaystyle\leq\binom{n}{s}\binom{n}{t}\left(\frac{t+s}{n}\right)^{sd}
(ens)s(en(1η)ds)(1η)ds((1η)ds+sn)ds\displaystyle\leq\left(\frac{en}{s}\right)^{s}\left(\frac{en}{(1-\eta)ds}\right)^{(1-\eta)ds}\left(\frac{(1-\eta)ds+s}{n}\right)^{ds}
(e(1η)d+1(1η)d((1η)d+1(1η)d)d((1η)dsn)ηd1)s\displaystyle\leq\left(e^{(1-\eta)d+1}(1-\eta)d\left(\frac{(1-\eta)d+1}{(1-\eta)d}\right)^{d}\left(\frac{(1-\eta)ds}{n}\right)^{\eta d-1}\right)^{s}
(edde1/(1η)(dsn)ηd1)s\displaystyle\leq\left(e^{d}{}de^{1/(1-\eta)}\left(\frac{ds}{n}\right)^{\eta d-1}\right)^{s}
(4edd(dsn)ηd1)s.\displaystyle\leq\left(4e^{d}{}d\left(\frac{ds}{n}\right)^{\eta d-1}\right)^{s}.

To bound the above we let ε=e1.1/η/d\varepsilon=e^{-1.1/\eta}/d, and we do a case analysis according to the value of ss.

For 3log2nse1.1/ηn/d3\log_{2}n\leq s\leq e^{-1.1/\eta}n/d, we have

(4edd(dsn)ηd1)s(4edde1.1de1.1/η)s2sn2.\displaystyle\left(4e^{d}{}d\left(\frac{ds}{n}\right)^{\eta d-1}\right)^{s}\leq\left(4e^{d}{}de^{-1.1d}e^{1.1/\eta}\right)^{s}\leq 2^{-s}\leq n^{-2}.

For 1s3log2n1\leq s\leq 3\log_{2}n, we have

(4edd(dsn)ηd1)s\displaystyle\left(4e^{d}{}d\left(\frac{ds}{n}\right)^{\eta d-1}\right)^{s} (4edd(3log2nn)ηd1)s\displaystyle\leq\left(4e^{d}{}d\left(\frac{3\log_{2}n}{\sqrt{n}}\right)^{\eta d-1}\right)^{s}
(4d(e1/η3log2nn)ηd1)s(d(1n6/16)ηd1)sn2.\displaystyle\leq\left(4d\left(\frac{e^{1/\eta}3\log_{2}n}{\sqrt{n}}\right)^{\eta d-1}\right)^{s}\leq\left(d\left(\frac{1}{{n}^{6/16}}\right)^{\eta d-1}\right)^{s}\leq n^{-2}.

Summing the above over s=1,,e1.1/ηn/ds=1,\ldots,e^{-1.1/\eta}n/d, we have

{GS}SV𝒢n,d~[SV,|S|e1.1/ηn/d|ΓGS(S)|(1η)d|S|]1O(n1).{\mathbb{P}}_{\{{G_{S}}\}_{S\subset{V}}\sim{\widetilde{\mathcal{G}_{n,d}}}}\bigl{[}\forall S\subset{V},|S|\leq e^{-1.1/\eta}n/d\implies|\Gamma_{{G_{S}}}(S)|\geq(1-\eta)d|S|\bigr{]}\geq 1-O(n^{-1}).

So, conditioned on \mathcal{B}, by Eq. (A.1), SV,|S|e1.1/ηn/d\forall S\subset{V},|S|\leq e^{-1.1/\eta}n/d we have

|ΓG(S)||ΓGS(S)|ηd|S|.|\Gamma_{G}(S)|\geq|\Gamma_{{G_{S}}}(S)|-\eta d|S|.

Since Pr𝒢n,d[]1O(n0.4)\Pr_{{\mathcal{G}_{n,d}}}[\mathcal{B}]\geq 1-O(n^{-0.4}), we conclude

G𝒢n,d[SV,|S|e1.1/ηn/d|ΓG(S)|(12η)d|S|]1O(n0.4).{\mathbb{P}}_{{G}\sim{\mathcal{G}_{n,d}}}\Bigl{[}{\forall S\subset{V},|S|\leq e^{-1.1/\eta}n/d\implies|\Gamma_{G}(S)|\geq(1-2\eta)d|S|}\Bigr{]}\geq 1-O(n^{-0.4}).

 

A.3 Summary

Restatement of Theorem 3.5. The random graph constructed in Section A.1 is a proper expander (see Definition 3.4), asymptotically almost surely. Specifically, the probability the graph has the desired properties is 1nO(1)\geq 1-n^{-O(1)}.

Proof:

Property (P1) is proved in Lemma A.4. Property (P2) is proved in Lemma A.6. Property (P3) follows from Theorem A.1.