-
Higher-order Delsarte Dual LPs: Lifting, Constructions and Completeness
Authors:
Leonardo Nagami Coregliano,
Fernando Granha Jeronimo,
Chris Jones,
Nati Linial,
Elyassaf Loyfer
Abstract:
A central and longstanding open problem in coding theory is the rate-versus-distance trade-off for binary error-correcting codes. In a seminal work, Delsarte introduced a family of linear programs establishing relaxations on the size of optimum codes. To date, the state-of-the-art upper bounds for binary codes come from dual feasible solutions to these LPs. Still, these bounds are exponentially fa…
▽ More
A central and longstanding open problem in coding theory is the rate-versus-distance trade-off for binary error-correcting codes. In a seminal work, Delsarte introduced a family of linear programs establishing relaxations on the size of optimum codes. To date, the state-of-the-art upper bounds for binary codes come from dual feasible solutions to these LPs. Still, these bounds are exponentially far from the best-known existential constructions.
Recently, hierarchies of linear programs extending and strengthening Delsarte's original LPs were introduced for linear codes, which we refer to as higher-order Delsarte LPs. These new hierarchies were shown to provably converge to the actual value of optimum codes, namely, they are complete hierarchies. Therefore, understanding them and their dual formulations becomes a valuable line of investigation. Nonetheless, their higher-order structure poses challenges. In fact, analysis of all known convex programming hierarchies strengthening Delsarte's original LPs has turned out to be exceedingly difficult and essentially nothing is known, stalling progress in the area since the 1970s.
Our main result is an analysis of the higher-order Delsarte LPs via their dual formulation. Although quantitatively, our current analysis only matches the best-known upper bounds, it shows, for the first time, how to tame the complexity of analyzing a hierarchy strengthening Delsarte's original LPs. In doing so, we reach a better understanding of the structure of the hierarchy, which may serve as the foundation for further quantitative improvements. We provide two additional structural results for this hierarchy. First, we show how to \emph{explicitly} lift any feasible dual solution from level $k$ to a (suitable) larger level $\ell$ while retaining the objective value. Second, we give a novel proof of completeness using the dual formulation.
△ Less
Submitted 8 January, 2025;
originally announced January 2025.
-
The Rank-Ramsey Problem and the Log-Rank Conjecture
Authors:
Gal Beniamini,
Nati Linial,
Adi Shraibman
Abstract:
A graph is called Rank-Ramsey if (i) Its clique number is small, and (ii) The adjacency matrix of its complement has small rank. We initiate a systematic study of such graphs. Our main motivation is that their constructions, as well as proofs of their non-existence, are intimately related to the famous log-rank conjecture from the field of communication complexity. These investigations also open i…
▽ More
A graph is called Rank-Ramsey if (i) Its clique number is small, and (ii) The adjacency matrix of its complement has small rank. We initiate a systematic study of such graphs. Our main motivation is that their constructions, as well as proofs of their non-existence, are intimately related to the famous log-rank conjecture from the field of communication complexity. These investigations also open interesting new avenues in Ramsey theory.
We construct two families of Rank-Ramsey graphs exhibiting polynomial separation between order and complement rank. Graphs in the first family have bounded clique number (as low as $41$). These are subgraphs of certain strong products, whose building blocks are derived from triangle-free strongly-regular graphs. Graphs in the second family are obtained by applying Boolean functions to Erdős-Rényi graphs. Their clique number is logarithmic, but their complement rank is far smaller than in the first family, about $\mathcal{O}(n^{2/3})$. A key component of this construction is our matrix-theoretic view of lifts.
We also consider lower bounds on the Rank-Ramsey numbers, and determine them in the range where the complement rank is $5$ or less. We consider connections between said numbers and other graph parameters, and find that the two best known explicit constructions of triangle-free Ramsey graphs turn out to be far from Rank-Ramsey.
△ Less
Submitted 17 October, 2024; v1 submitted 12 May, 2024;
originally announced May 2024.
-
An Elementary Proof of the First LP Bound on the Rate of Binary Codes
Authors:
Nati Linial,
Elyassaf Loyfer
Abstract:
The asymptotic rate vs. distance problem is a long-standing fundamental problem in coding theory. The best upper bound to date was given in 1977 and has received since then numerous proofs and interpretations. Here we provide a new, elementary proof of this bound based on counting walks in the Hamming cube.
The asymptotic rate vs. distance problem is a long-standing fundamental problem in coding theory. The best upper bound to date was given in 1977 and has received since then numerous proofs and interpretations. Here we provide a new, elementary proof of this bound based on counting walks in the Hamming cube.
△ Less
Submitted 29 March, 2023;
originally announced March 2023.
-
Linear Programming Hierarchies in Coding Theory: Dual Solutions
Authors:
Elyassaf Loyfer,
Nati Linial
Abstract:
The rate vs. distance problem is a long-standing open problem in coding theory. Recent papers have suggested a new way to tackle this problem by appealing to a new hierarchy of linear programs. If one can find good dual solutions to these LPs, this would result in improved upper bounds for the rate vs. distance problem of linear codes. In this work, we develop the first dual feasible solutions to…
▽ More
The rate vs. distance problem is a long-standing open problem in coding theory. Recent papers have suggested a new way to tackle this problem by appealing to a new hierarchy of linear programs. If one can find good dual solutions to these LPs, this would result in improved upper bounds for the rate vs. distance problem of linear codes. In this work, we develop the first dual feasible solutions to the LPs in this hierarchy. These match the best-known bound for a wide range of parameters. Our hope is that this is a first step towards better solutions, and improved upper bounds for the rate vs. distance problem of linear codes.
△ Less
Submitted 23 November, 2022;
originally announced November 2022.
-
New LP-based Upper Bounds in the Rate-vs.-Distance Problem for Linear Codes
Authors:
Elyassaf Loyfer,
Nati Linial
Abstract:
We develop a new family of linear programs, that yield upper bounds on the rate of binary linear codes of a given distance. Our bounds apply {\em only to linear codes.} Delsarte's LP is the weakest member of this family and our LP yields increasingly tighter upper bounds on the rate as its control parameter increases. Numerical experiments show significant improvement compared to Delsarte. These c…
▽ More
We develop a new family of linear programs, that yield upper bounds on the rate of binary linear codes of a given distance. Our bounds apply {\em only to linear codes.} Delsarte's LP is the weakest member of this family and our LP yields increasingly tighter upper bounds on the rate as its control parameter increases. Numerical experiments show significant improvement compared to Delsarte. These convincing numerical results, and the large variety of tools available for asymptotic analysis, give us hope that our work will lead to new and improved asymptotic upper bounds on the possible rate of linear codes.
A concurrent work by Coregliano, Jeronimo, and Jones offers a closely related family of linear programs which converges to the true bound. Here we provide a new proof of convergence for the same LPs.
△ Less
Submitted 15 November, 2022; v1 submitted 18 June, 2022;
originally announced June 2022.
-
Bounds on Unique-Neighbor Codes
Authors:
Nati Linial,
Edan Orzech
Abstract:
Recall that a binary linear code of length $n$ is a linear subspace $\mathcal{C} = \{x\in\mathbb{F}_2^n\mid Ax=0\}$. Here the parity check matrix $A$ is a binary $m\times n$ matrix of rank $m$. We say that $\mathcal{C}$ has rate $R=1-\frac mn$. Its distance, denoted $δn$ is the smallest Hamming weight of a non-zero vector in $\mathcal{C}$. The rate vs.\ distance problem for binary linear codes is…
▽ More
Recall that a binary linear code of length $n$ is a linear subspace $\mathcal{C} = \{x\in\mathbb{F}_2^n\mid Ax=0\}$. Here the parity check matrix $A$ is a binary $m\times n$ matrix of rank $m$. We say that $\mathcal{C}$ has rate $R=1-\frac mn$. Its distance, denoted $δn$ is the smallest Hamming weight of a non-zero vector in $\mathcal{C}$. The rate vs.\ distance problem for binary linear codes is a fundamental open problem in coding theory, and a fascinating question in discrete mathematics. It concerns the function $R_L(δ)$, the largest possible rate $R$ for given $0\leδ\le1$ and arbitrarily large length $n$. Here we investigate a variation of this fundamental question that we describe next.
Clearly, $\mathcal{C}$ has distance $δn$, if and only if for every $0<n'<δn$, every $m\times n'$ submatrix of $A$ has a row of odd weight. Motivated by several problems from coding theory, we say that $A$ has the unique-neighbor property with parameter $δn$, if every such submatrix has a row of weight $1$. Let $R_U(δ)$ be the largest possible asymptotic rate of linear codes with a parity check matrix that has this stronger property. Clearly, $R_U(\cdot),R_L(\cdot)$ are non-increasing functions, and $R_U(δ)\le R_L(δ)$ for all $δ$. Also, $R_U(0)=R_L(0)=1$, and $R_U(1)=R_L(1)=0$, so let $0\leδ_U \leδ_L\le1$ be the smallest values of $δ$ at which $R_U$ resp.\ $R_L$ vanish. It is well known that $δ_L=\frac12$ and we conjecture that $δ_U$ is strictly smaller than $\frac12$, i.e., the rate of linear codes with the unique-neighbor property is more strictly bounded. While the conjecture remains open, we prove here several results supporting it.
The reader is not assumed to have any specific background in coding theory, but we occasionally point out some relevant facts from that area.
△ Less
Submitted 3 April, 2025; v1 submitted 19 March, 2022;
originally announced March 2022.
-
Hyperpaths
Authors:
Amir Dahari,
Nati Linial
Abstract:
Hypertrees are high-dimensional counterparts of graph theoretic trees. They have attracted a great deal of attention by various investigators. Here we introduce and study Hyperpaths -- a particular class of hypertrees which are high dimensional analogs of paths in graph theory. A $d$-dimensional hyperpath is a $d$-dimensional hypertree in which every $(d-1)$-dimensional face is contained in at mos…
▽ More
Hypertrees are high-dimensional counterparts of graph theoretic trees. They have attracted a great deal of attention by various investigators. Here we introduce and study Hyperpaths -- a particular class of hypertrees which are high dimensional analogs of paths in graph theory. A $d$-dimensional hyperpath is a $d$-dimensional hypertree in which every $(d-1)$-dimensional face is contained in at most $(d+1)$ faces of dimension $d$. We introduce a possibly infinite family of hyperpaths for every dimension, and investigate its properties in greater depth for dimension $d=2$.
△ Less
Submitted 19 November, 2020;
originally announced November 2020.
-
A King in every two consecutive tournaments
Authors:
Yehuda Afek,
Eli Gafni,
Nati Linial
Abstract:
We think of a tournament $T=([n], E)$ as a communication network where in each round of communication processor $P_i$ sends its information to $P_j$, for every directed edge $ij \in E(T)$. By Landau's theorem (1953) there is a King in $T$, i.e., a processor whose initial input reaches every other processor in two rounds or less. Namely, a processor $P_ν$ such that after two rounds of communication…
▽ More
We think of a tournament $T=([n], E)$ as a communication network where in each round of communication processor $P_i$ sends its information to $P_j$, for every directed edge $ij \in E(T)$. By Landau's theorem (1953) there is a King in $T$, i.e., a processor whose initial input reaches every other processor in two rounds or less. Namely, a processor $P_ν$ such that after two rounds of communication along $T$'s edges, the initial information of $P_ν$ reaches all other processors. Here we consider a more general scenario where an adversary selects an arbitrary series of tournaments $T_1, T_2,\ldots$, so that in each round $s=1, 2, \ldots$, communication is governed by the corresponding tournament $T_s$. We prove that for every series of tournaments that the adversary selects, it is still true that after two rounds of communication, the initial input of at least one processor reaches everyone. Concretely, we show that for every two tournaments $T_1, T_2$ there is a vertex in $[n]$ that can reach all vertices via (i) A step in $T_1$, or (ii) A step in $T_2$ or (iii) A step in $T_1$ followed by a step in $T_2$. }
△ Less
Submitted 21 October, 2019;
originally announced October 2019.
-
On the weight distribution of random binary linear codes
Authors:
Nati Linial,
Jonathan Mosheiff
Abstract:
We investigate the weight distribution of random binary linear codes. For $0<λ<1$ and $n\to\infty$ pick uniformly at random $λn$ vectors in $\mathbb{F}_2^n$ and let $C \le \mathbb{F}_2^n$ be the orthogonal complement of their span. Given $0<γ<1/2$ with $0< λ< h(γ)$ let $X$ be the random variable that counts the number of words in $C$ of Hamming weight $γn$. In this paper we determine the asymptoti…
▽ More
We investigate the weight distribution of random binary linear codes. For $0<λ<1$ and $n\to\infty$ pick uniformly at random $λn$ vectors in $\mathbb{F}_2^n$ and let $C \le \mathbb{F}_2^n$ be the orthogonal complement of their span. Given $0<γ<1/2$ with $0< λ< h(γ)$ let $X$ be the random variable that counts the number of words in $C$ of Hamming weight $γn$. In this paper we determine the asymptotics of the moments of $X$ of all orders $o(\frac{n}{\log n})$.
△ Less
Submitted 21 June, 2018;
originally announced June 2018.
-
On The Communication Complexity of High-Dimensional Permutations
Authors:
Nati Linial,
and Toniann Pitassi,
Adi Shraibman
Abstract:
We study the multiparty communication complexity of high dimensional permutations, in the Number On the Forehead (NOF) model. This model is due to Chandra, Furst and Lipton (CFL) who also gave a nontrivial protocol for the Exactly-n problem where three players receive integer inputs and need to decide if their inputs sum to a given integer $n$. There is a considerable body of literature dealing wi…
▽ More
We study the multiparty communication complexity of high dimensional permutations, in the Number On the Forehead (NOF) model. This model is due to Chandra, Furst and Lipton (CFL) who also gave a nontrivial protocol for the Exactly-n problem where three players receive integer inputs and need to decide if their inputs sum to a given integer $n$. There is a considerable body of literature dealing with the same problem, where $(\mathbb{N},+)$ is replaced by some other abelian group. Our work can be viewed as a far-reaching extension of this line of work.
We show that the known lower bounds for that group-theoretic problem apply to all high dimensional permutations. We introduce new proof techniques that appeal to recent advances in Additive Combinatorics and Ramsey theory. We reveal new and unexpected connections between the NOF communication complexity of high dimensional permutations and a variety of well known and thoroughly studied problems in combinatorics.
Previous protocols for Exactly-n all rely on the construction of large sets of integers without a 3-term arithmetic progression. No direct algorithmic protocol was previously known for the problem, and we provide the first such algorithm. This suggests new ways to significantly improve the CFL protocol.
Many new open questions are presented throughout.
△ Less
Submitted 27 November, 2018; v1 submitted 7 June, 2017;
originally announced June 2017.
-
On the Rigidity of Sparse Random Graphs
Authors:
Nati Linial,
Jonathan Mosheiff
Abstract:
A graph with a trivial automorphism group is said to be rigid. Wright proved that for $\frac{\log n}{n}+ω(\frac 1n)\leq p\leq \frac 12$ a random graph $G\in G(n,p)$ is rigid whp. It is not hard to see that this lower bound is sharp and for $p<\frac{(1-ε)\log n}{n}$ with positive probability $\text{aut}(G)$ is nontrivial. We show that in the sparser case…
▽ More
A graph with a trivial automorphism group is said to be rigid. Wright proved that for $\frac{\log n}{n}+ω(\frac 1n)\leq p\leq \frac 12$ a random graph $G\in G(n,p)$ is rigid whp. It is not hard to see that this lower bound is sharp and for $p<\frac{(1-ε)\log n}{n}$ with positive probability $\text{aut}(G)$ is nontrivial. We show that in the sparser case $ω(\frac 1 n)\leq p\leq \frac{\log n}{n}+ω(\frac 1n)$, it holds whp that $G$'s $2$-core is rigid. We conclude that for all $p$, a graph in $G(n,p)$ is reconstrutible whp. In addition this yields for $ω(\frac 1n)\leq p\leq \frac 12$ a canonical labeling algorithm that almost surely runs in polynomial time with $o(1)$ error rate. This extends the range for which such an algorithm is currently known.
△ Less
Submitted 5 May, 2015;
originally announced May 2015.
-
Market Share Indicates Quality
Authors:
Amir Ban,
Nati Linial
Abstract:
Market share and quality, or customer satisfaction, go together. Yet inferring one from the other appears difficult. Indeed, such an inference would need detailed information about customer behavior, and might be clouded by modes of behavior such as herding (following popularity) or elitism, where customers avoid popular products. We investigate a fixed-price model where customers are informed abo…
▽ More
Market share and quality, or customer satisfaction, go together. Yet inferring one from the other appears difficult. Indeed, such an inference would need detailed information about customer behavior, and might be clouded by modes of behavior such as herding (following popularity) or elitism, where customers avoid popular products. We investigate a fixed-price model where customers are informed about their history with products and about market share data. We find that it is in fact correct to make a Bayesian inference that the product with the higher market share has the better quality under few and unrestrictive assumptions on customer behavior.
△ Less
Submitted 14 July, 2014;
originally announced July 2014.
-
From average case complexity to improper learning complexity
Authors:
Amit Daniely,
Nati Linial,
Shai Shalev-Shwartz
Abstract:
The basic problem in the PAC model of computational learning theory is to determine which hypothesis classes are efficiently learnable. There is presently a dearth of results showing hardness of learning problems. Moreover, the existing lower bounds fall short of the best known algorithms.
The biggest challenge in proving complexity results is to establish hardness of {\em improper learning} (a.…
▽ More
The basic problem in the PAC model of computational learning theory is to determine which hypothesis classes are efficiently learnable. There is presently a dearth of results showing hardness of learning problems. Moreover, the existing lower bounds fall short of the best known algorithms.
The biggest challenge in proving complexity results is to establish hardness of {\em improper learning} (a.k.a. representation independent learning).The difficulty in proving lower bounds for improper learning is that the standard reductions from $\mathbf{NP}$-hard problems do not seem to apply in this context. There is essentially only one known approach to proving lower bounds on improper learning. It was initiated in (Kearns and Valiant 89) and relies on cryptographic assumptions.
We introduce a new technique for proving hardness of improper learning, based on reductions from problems that are hard on average. We put forward a (fairly strong) generalization of Feige's assumption (Feige 02) about the complexity of refuting random constraint satisfaction problems. Combining this assumption with our new technique yields far reaching implications. In particular,
1. Learning $\mathrm{DNF}$'s is hard.
2. Agnostically learning halfspaces with a constant approximation ratio is hard.
3. Learning an intersection of $ω(1)$ halfspaces is hard.
△ Less
Submitted 9 March, 2014; v1 submitted 10 November, 2013;
originally announced November 2013.
-
More data speeds up training time in learning halfspaces over sparse vectors
Authors:
Amit Daniely,
Nati Linial,
Shai Shalev Shwartz
Abstract:
The increased availability of data in recent years has led several authors to ask whether it is possible to use data as a {\em computational} resource. That is, if more data is available, beyond the sample complexity limit, is it possible to use the extra examples to speed up the computation time required to perform the learning task?
We give the first positive answer to this question for a {\em…
▽ More
The increased availability of data in recent years has led several authors to ask whether it is possible to use data as a {\em computational} resource. That is, if more data is available, beyond the sample complexity limit, is it possible to use the extra examples to speed up the computation time required to perform the learning task?
We give the first positive answer to this question for a {\em natural supervised learning problem} --- we consider agnostic PAC learning of halfspaces over $3$-sparse vectors in $\{-1,1,0\}^n$. This class is inefficiently learnable using $O\left(n/ε^2\right)$ examples. Our main contribution is a novel, non-cryptographic, methodology for establishing computational-statistical gaps, which allows us to show that, under a widely believed assumption that refuting random $\mathrm{3CNF}$ formulas is hard, it is impossible to efficiently learn this class using only $O\left(n/ε^2\right)$ examples. We further show that under stronger hardness assumptions, even $O\left(n^{1.499}/ε^2\right)$ examples do not suffice. On the other hand, we show a new algorithm that learns this class efficiently using $\tildeΩ\left(n^2/ε^2\right)$ examples. This formally establishes the tradeoff between sample and computational complexity for a natural supervised learning problem.
△ Less
Submitted 10 November, 2013;
originally announced November 2013.
-
The complexity of learning halfspaces using generalized linear methods
Authors:
Amit Daniely,
Nati Linial,
Shai Shalev-Shwartz
Abstract:
Many popular learning algorithms (E.g. Regression, Fourier-Transform based algorithms, Kernel SVM and Kernel ridge regression) operate by reducing the problem to a convex optimization problem over a vector space of functions. These methods offer the currently best approach to several central problems such as learning half spaces and learning DNF's. In addition they are widely used in numerous appl…
▽ More
Many popular learning algorithms (E.g. Regression, Fourier-Transform based algorithms, Kernel SVM and Kernel ridge regression) operate by reducing the problem to a convex optimization problem over a vector space of functions. These methods offer the currently best approach to several central problems such as learning half spaces and learning DNF's. In addition they are widely used in numerous application domains. Despite their importance, there are still very few proof techniques to show limits on the power of these algorithms.
We study the performance of this approach in the problem of (agnostically and improperly) learning halfspaces with margin $γ$. Let $\mathcal{D}$ be a distribution over labeled examples. The $γ$-margin error of a hyperplane $h$ is the probability of an example to fall on the wrong side of $h$ or at a distance $\leγ$ from it. The $γ$-margin error of the best $h$ is denoted $\mathrm{Err}_γ(\mathcal{D})$. An $α(γ)$-approximation algorithm receives $γ,ε$ as input and, using i.i.d. samples of $\mathcal{D}$, outputs a classifier with error rate $\le α(γ)\mathrm{Err}_γ(\mathcal{D}) + ε$. Such an algorithm is efficient if it uses $\mathrm{poly}(\frac{1}γ,\frac{1}ε)$ samples and runs in time polynomial in the sample size.
The best approximation ratio achievable by an efficient algorithm is $O\left(\frac{1/γ}{\sqrt{\log(1/γ)}}\right)$ and is achieved using an algorithm from the above class. Our main result shows that the approximation ratio of every efficient algorithm from this family must be $\ge Ω\left(\frac{1/γ}{\mathrm{poly}\left(\log\left(1/γ\right)\right)}\right)$, essentially matching the best known upper bound.
△ Less
Submitted 10 May, 2014; v1 submitted 3 November, 2012;
originally announced November 2012.
-
Musical chairs
Authors:
Yehuda Afek,
Yakov Babichenko,
Uriel Feige,
Eli Gafni,
Nati Linial,
Benny Sudakov
Abstract:
In the {\em Musical Chairs} game $MC(n,m)$ a team of $n$ players plays against an adversarial {\em scheduler}. The scheduler wins if the game proceeds indefinitely, while termination after a finite number of rounds is declared a win of the team. At each round of the game each player {\em occupies} one of the $m$ available {\em chairs}. Termination (and a win of the team) is declared as soon as eac…
▽ More
In the {\em Musical Chairs} game $MC(n,m)$ a team of $n$ players plays against an adversarial {\em scheduler}. The scheduler wins if the game proceeds indefinitely, while termination after a finite number of rounds is declared a win of the team. At each round of the game each player {\em occupies} one of the $m$ available {\em chairs}. Termination (and a win of the team) is declared as soon as each player occupies a unique chair. Two players that simultaneously occupy the same chair are said to be {\em in conflict}. In other words, termination (and a win for the team) is reached as soon as there are no conflicts. The only means of communication throughout the game is this: At every round of the game, the scheduler selects an arbitrary nonempty set of players who are currently in conflict, and notifies each of them separately that it must move. A player who is thus notified changes its chair according to its deterministic program. As we show, for $m\ge 2n-1$ chairs the team has a winning strategy. Moreover, using topological arguments we show that this bound is tight. For $m\leq 2n-2$ the scheduler has a strategy that is guaranteed to make the game continue indefinitely and thus win. We also have some results on additional interesting questions. For example, if $m \ge 2n-1$ (so that the team can win), how quickly can they achieve victory?
△ Less
Submitted 3 August, 2012;
originally announced August 2012.
-
On the practically interesting instances of MAXCUT
Authors:
Yonatan Bilu,
Amit Daniely,
Nati Linial,
Michael Saks
Abstract:
The complexity of a computational problem is traditionally quantified based on the hardness of its worst case. This approach has many advantages and has led to a deep and beautiful theory. However, from the practical perspective, this leaves much to be desired. In application areas, practically interesting instances very often occupy just a tiny part of an algorithm's space of instances, and the v…
▽ More
The complexity of a computational problem is traditionally quantified based on the hardness of its worst case. This approach has many advantages and has led to a deep and beautiful theory. However, from the practical perspective, this leaves much to be desired. In application areas, practically interesting instances very often occupy just a tiny part of an algorithm's space of instances, and the vast majority of instances are simply irrelevant. Addressing these issues is a major challenge for theoretical computer science which may make theory more relevant to the practice of computer science.
Following Bilu and Linial, we apply this perspective to MAXCUT, viewed as a clustering problem. Using a variety of techniques, we investigate practically interesting instances of this problem. Specifically, we show how to solve in polynomial time distinguished, metric, expanding and dense instances of MAXCUT under mild stability assumptions. In particular, $(1+ε)$-stability (which is optimal) suffices for metric and dense MAXCUT. We also show how to solve in polynomial time $Ω(\sqrt{n})$-stable instances of MAXCUT, substantially improving the best previously known result.
△ Less
Submitted 22 May, 2012;
originally announced May 2012.
-
Clustering is difficult only when it does not matter
Authors:
Amit Daniely,
Nati Linial,
Michael Saks
Abstract:
Numerous papers ask how difficult it is to cluster data. We suggest that the more relevant and interesting question is how difficult it is to cluster data sets {\em that can be clustered well}. More generally, despite the ubiquity and the great importance of clustering, we still do not have a satisfactory mathematical theory of clustering. In order to properly understand clustering, it is clearly…
▽ More
Numerous papers ask how difficult it is to cluster data. We suggest that the more relevant and interesting question is how difficult it is to cluster data sets {\em that can be clustered well}. More generally, despite the ubiquity and the great importance of clustering, we still do not have a satisfactory mathematical theory of clustering. In order to properly understand clustering, it is clearly necessary to develop a solid theoretical basis for the area. For example, from the perspective of computational complexity theory the clustering problem seems very hard. Numerous papers introduce various criteria and numerical measures to quantify the quality of a given clustering. The resulting conclusions are pessimistic, since it is computationally difficult to find an optimal clustering of a given data set, if we go by any of these popular criteria. In contrast, the practitioners' perspective is much more optimistic. Our explanation for this disparity of opinions is that complexity theory concentrates on the worst case, whereas in reality we only care for data sets that can be clustered well.
We introduce a theoretical framework of clustering in metric spaces that revolves around a notion of "good clustering". We show that if a good clustering exists, then in many cases it can be efficiently found. Our conclusion is that contrary to popular belief, clustering should not be considered a hard task.
△ Less
Submitted 22 May, 2012;
originally announced May 2012.
-
No justified complaints: On fair sharing of multiple resources
Authors:
Danny Dolev,
Dror G. Feitelson,
Joseph Y. Halpern,
Raz Kupferman,
Nati Linial
Abstract:
Fair allocation has been studied intensively in both economics and computer science, and fair sharing of resources has aroused renewed interest with the advent of virtualization and cloud computing. Prior work has typically focused on mechanisms for fair sharing of a single resource. We provide a new definition for the simultaneous fair allocation of multiple continuously-divisible resources…
▽ More
Fair allocation has been studied intensively in both economics and computer science, and fair sharing of resources has aroused renewed interest with the advent of virtualization and cloud computing. Prior work has typically focused on mechanisms for fair sharing of a single resource. We provide a new definition for the simultaneous fair allocation of multiple continuously-divisible resources. Roughly speaking, we define fairness as the situation where every user either gets all the resources he wishes for, or else gets at least his entitlement on some bottleneck resource, and therefore cannot complain about not getting more. This definition has the same desirable properties as the recently suggested dominant resource fairness, and also handles the case of multiple bottlenecks. We then prove that a fair allocation according to this definition is guaranteed to exist for any combination of user requests and entitlements (where a user's relative use of the different resources is fixed). The proof, which uses tools from the theory of ordinary differential equations, is constructive and provides a method to compute the allocations numerically.
△ Less
Submitted 14 June, 2011;
originally announced June 2011.
-
Oblivious Collaboration
Authors:
Yehuda Afek,
Yakov Babichenko,
Uriel Feige,
Eli Gafni,
Nati Linial,
Benny Sudakov
Abstract:
Communication is a crucial ingredient in every kind of collaborative work. But what is the least possible amount of communication required for a given task? We formalize this question by introducing a new framework for distributed computation, called {\em oblivious protocols}.
We investigate the power of this model by considering two concrete examples, the {\em musical chairs} task $MC(n,m)$ and…
▽ More
Communication is a crucial ingredient in every kind of collaborative work. But what is the least possible amount of communication required for a given task? We formalize this question by introducing a new framework for distributed computation, called {\em oblivious protocols}.
We investigate the power of this model by considering two concrete examples, the {\em musical chairs} task $MC(n,m)$ and the well-known {\em Renaming} problem. The $MC(n,m)$ game is played by $n$ players (processors) with $m$ chairs. Players can {\em occupy} chairs, and the game terminates as soon as each player occupies a unique chair. Thus we say that player $P$ is {\em in conflict} if some other player $Q$ is occupying the same chair, i.e., termination means there are no conflicts. By known results from distributed computing, if $m \le 2n-2$, no strategy of the players can guarantee termination. However, there is a protocol with $m = 2n-1$ chairs that always terminates. Here we consider an oblivious protocol where in every time step the only communication is this: an adversarial {\em scheduler} chooses an arbitrary nonempty set of players, and for each of them provides only one bit of information, specifying whether the player is currently in conflict or not. A player notified not to be in conflict halts and never changes its chair, whereas a player notified to be in conflict changes its chair according to its deterministic program. Remarkably, even with this minimal communication termination can be guaranteed with only $m=2n-1$ chairs. Likewise, we obtain an oblivious protocol for the Renaming problem whose name-space is small as that of the optimal nonoblivious distributed protocol.
Other aspects suggest themselves, such as the efficiency (program length) of our protocols. We make substantial progress here as well, though many interesting questions remain open.
△ Less
Submitted 10 June, 2011;
originally announced June 2011.
-
Tight products and Expansion
Authors:
Amit Daniely,
Nathan Linial
Abstract:
In this paper we study a new product of graphs called {\em tight product}. A graph $H$ is said to be a tight product of two (undirected multi) graphs $G_1$ and $G_2$, if $V(H)=V(G_1)\times V(G_2)$ and both projection maps $V(H)\to V(G_1)$ and $V(H)\to V(G_2)$ are covering maps. It is not a priori clear when two given graphs have a tight product (in fact, it is $NP$-hard to decide). We investigate…
▽ More
In this paper we study a new product of graphs called {\em tight product}. A graph $H$ is said to be a tight product of two (undirected multi) graphs $G_1$ and $G_2$, if $V(H)=V(G_1)\times V(G_2)$ and both projection maps $V(H)\to V(G_1)$ and $V(H)\to V(G_2)$ are covering maps. It is not a priori clear when two given graphs have a tight product (in fact, it is $NP$-hard to decide). We investigate the conditions under which this is possible. This perspective yields a new characterization of class-1 $(2k+1)$-regular graphs. We also obtain a new model of random $d$-regular graphs whose second eigenvalue is almost surely at most $O(d^{3/4})$. This construction resembles random graph lifts, but requires fewer random bits.
△ Less
Submitted 3 November, 2012; v1 submitted 20 January, 2010;
originally announced January 2010.
-
Are stable instances easy?
Authors:
Yonatan Bilu,
Nathan Linial
Abstract:
We introduce the notion of a stable instance for a discrete optimization problem, and argue that in many practical situations only sufficiently stable instances are of interest. The question then arises whether stable instances of NP--hard problems are easier to solve. In particular, whether there exist algorithms that solve correctly and in polynomial time all sufficiently stable instances of s…
▽ More
We introduce the notion of a stable instance for a discrete optimization problem, and argue that in many practical situations only sufficiently stable instances are of interest. The question then arises whether stable instances of NP--hard problems are easier to solve. In particular, whether there exist algorithms that solve correctly and in polynomial time all sufficiently stable instances of some NP--hard problem. The paper focuses on the Max--Cut problem, for which we show that this is indeed the case.
△ Less
Submitted 17 June, 2009;
originally announced June 2009.
-
On metric Ramsey-type phenomena
Authors:
Yair Bartal,
Nathan Linial,
Manor Mendel,
Assaf Naor
Abstract:
The main question studied in this article may be viewed as a nonlinear analogue of Dvoretzky's theorem in Banach space theory or as part of Ramsey theory in combinatorics. Given a finite metric space on n points, we seek its subspace of largest cardinality which can be embedded with a given distortion in Hilbert space. We provide nearly tight upper and lower bounds on the cardinality of this sub…
▽ More
The main question studied in this article may be viewed as a nonlinear analogue of Dvoretzky's theorem in Banach space theory or as part of Ramsey theory in combinatorics. Given a finite metric space on n points, we seek its subspace of largest cardinality which can be embedded with a given distortion in Hilbert space. We provide nearly tight upper and lower bounds on the cardinality of this subspace in terms of n and the desired distortion. Our main theorem states that for any epsilon>0, every n point metric space contains a subset of size at least n^{1-ε} which is embeddable in Hilbert space with O(\frac{\log(1/ε)}ε) distortion. The bound on the distortion is tight up to the log(1/ε) factor. We further include a comprehensive study of various other aspects of this problem.
△ Less
Submitted 20 June, 2007; v1 submitted 17 June, 2004;
originally announced June 2004.