throbber
Marc Fossorier Hideki Imai
`Shu Lin Alain Poli (Eds.)
`
`Applied Algebra,
`Algebraic Algorithms and
`Error-Correcting Codes
`
`13th International Symposium, AAECC-13
`Honolulu, Hawaii, USA, November 15-19, 1999
`Proceedings
`
`1 3
`
`Apple 1224
`
`1
`
`

`
`Series Editors
`Gerhard Goos, Karlsruhe University, Germany
`Juris Hartmanis, Cornell University, NY, USA
`Jan van Leeuwen, Utrecht University, The Netherlands
`
`Volume Editors
`
`Marc Fossorier
`Shu Lin
`University of Hawaii, Department of Electrical Engineering
`2540 Dole St., # 483, Honolulu, HI 96822, USA
`E-mail: marc@aravis.eng.hawaii.edu; slin@spectra.eng.hawaii.edu
`
`Hideki Imai
`University of Tokyo, Institute of Industrial Science
`7-22-1, Roppongi, Minato-ku, Tokyo 106, Japan
`E-mail: imai@iis.u-tokyo.ac.jp
`
`Alain Poli
`Universit´e Paul Sabatier, AAECC/IRIT
`118, Route de Narbonne, F31067 Toulouse Cedex, France
`E-mail: poli@cict.fr
`
`Cataloging-in-Publication data applied for
`
`Die Deutsche Bibliothek - CIP-Einheitsaufnahme
`Applied algebra, algebraic algorithms and error-correcting codes
`: 13th international symposium ; proceedings / AAECC 13, Honolulu,
`Hawaii, USA, November 15 - 19, 1999. Marc Fossorier . . . (ed.). -
`Berlin ; Heidelberg ; New York ; Barcelona ; Hong Kong ; London ;
`Milan ; Paris ; Singapore ; Tokyo : Springer, 1999
`(Lecture notes in computer science ; Vol. 1719)
`ISBN 3-540-66723-7
`
`CR Subject Classi cation (1998): E.4, I.1, E.3, G.2, F.2
`
`ISSN 0302-9743
`ISBN 3-540-66723-7 Springer-Verlag Berlin Heidelberg New York
`
`This work is subject to copyright. All rights are reserved, whether the whole or part of the material is
`concerned, speci cally the rights of translation, reprinting, re-use of illustrations, recitation, broadcasting,
`reproduction on micro lms or in any other way, and storage in data banks. Duplication of this publication
`or parts thereof is permitted only under the provisions of the German Copyright Law of September 9, 1965,
`in its current version, and permission for use must always be obtained from Springer-Verlag. Violations are
`liable for prosecution under the German Copyright Law.
`c⃝ Springer-Verlag Berlin Heidelberg 1999
`Printed in Germany
`Typesetting: Camera-ready by author
`SPIN 10704606
`06/3142 – 5 4 3 2 1 0
`
`Printed on acid-free paper
`
`2
`
`

`
`New Sequences of Linear Time Erasure Codes
`Approaching the Channel Capacity
`
`M. Amin Shokrollahi
`
`Bell Labs, Room 2C-353, 700 Mountain Ave, Murray Hill, NJ 07974, USA
`amin@research.bell-labs.com
`
`Abstract. We will introduce a new class of erasure codes built from
`irregular bipartite graphs that have linear time encoding and decoding
`algorithms and can transmit over an erasure channel at rates arbitrarily
`close to the channel capacity. We also show that these codes are close
`to optimal with respect to the trade-off between the proximity to the
`channel capacity and the running time of the recovery algorithm.
`
`1
`
`Introduction
`
`A linear error-correcting code of block length n and dimension k over a finite field
`IFq—an [n, k]q-code for short—is a k-dimensional linear subspace of the standard
`vector space IFn
`q . The elements of the code are called codewords. To the code C
`there corresponds an encoding map Enc which is an isomorphism of the vector
`spaces IFk
`q and C. A sender, who wishes to transmit a vector of k elements in
`IFq to a receiver uses the mapping Enc to encode that vector into a codeword.
`The rate k/n of the code is a measure for the amount of real information in each
`codeword. The minimum distance of the code is the minimum Hamming distance
`between two distinct codewords. A linear code of block length n, dimension k,
`and minimum distance d over IFq is called an [n, k, d]q-code.
`Linear codes can be used to reliably transmit information over a noisy chan-
`nel. Depending on the nature of the errors imposed on the codeword during
`the transmission, the receiver then applies appropriate algorithms to decode the
`received word. In this paper, we assume that the receiver knows the position of
`each received symbol within the stream of all encoding symbols. We adopt as
`our model of losses the erasure channel, introduced by Elias [3], in which each
`encoding symbol is lost with a fixed constant probability p in transit indepen-
`dent of all the other symbols. As was shown by Elias [3], the capacity of this
`channel equals 1 − p.
`It is easy to see that a code of minimum distance d is capable of recovering
`d−1 or less erasures. In the best case, it can recover from any set of k coordinates
`of the encoding which means that d − 1 = n − k. Such codes are called MDS-
`codes. A standard class of MDS-codes is given by Reed-Solomon codes [10]. The
`connection of these codes with polynomial arithmetic allows for encoding and
`decoding in time O(n log2 n log log n). (See, [2, Chapter 11.7] and [10, p. 369]).
`However, these codes do not reach the capacity of the erasure channel, since
`there is no infinite sequence of such codes over a fixed field.
`
`Marc Fossorier et al. (Eds.): AAECC-13, LNCS 1719, pp. 65–76, 1999.
`c⃝ Springer-Verlag Berlin Heidelberg 1999
`
`3
`
`

`
`66
`
`M. Amin Shokrollahi
`
`Elias [3] showed that a random linear code can be used to transmit over the
`erasure channel at any rate R < 1 − p, and that encoding and decoding can be
`accomplished with O(n2) and O(n3) arithmetic operations, respectively. Hence,
`we have on the one hand codes that can be encoded and decoded faster than
`general linear codes, but do not reach the capacity of the erasure channel; and
`on the other hand we have random codes which reach the capacity but have
`encoding and decoding algorithms of higher complexity.
`The paper [1] was the first to design codes that could come arbitrarily close
`to the channel capacity while having linear time encoding and decoding algo-
`rithms. Improving these results, the authors of [8] took a different approach and
`designed fast linear-time algorithms for transmitting just below channel capac-
`ity. For all ϵ> 0 they were able to produce rate R = 1 − p(1 + ϵ) codes along
`with decoding algorithms that could recover from the random loss of a p fraction
`of the transmitted symbols in time proportional to n ln(1/ϵ) with high proba-
`bility, where n is the block length. These codes could also be encoded in time
`proportional to n ln(1/ϵ). They belong to the class of low-density parity check
`codes of Gallager [4]. In contrast to Gallager codes, however, the graphs used to
`construct the asymptotically good codes obtained in [8] are highly irregular.
`The purpose of the present paper is twofold. First, we prove a general trade-off
`theorem between the proximity of a given Gallager code to the channel capacity
`in terms of the loss fraction and the running time of the recovery algorithm of [8].
`We show that in this respect, the codes constructed in that paper are close to
`optimal. Next, we exhibit a different sequence of asymptotically close to optimal
`codes which have better parameters than the codes in [8]. An interesting feature
`of these codes is that the underlying bipartite graphs are right regular, i.e., all
`nodes on the right hand side of the graph have the same degree. Since they are
`theoretically better than their peers, we expect them to also perform better in
`practice.
`The organization of the paper is as follows. In the next section we will re-
`view the construction of Gallager codes. Next, we prove upper bounds on the
`maximum tolerable loss fraction in terms of the running time of the decoding al-
`gorithm. The last two sections are concerned with the derivation of the sequence
`of right regular erasure codes.
`
`2 Codes from Bipartite Graphs
`
`In this section, we will briefly review the class of codes we are interested in, and
`the erasure recovery algorithm associated to them.
`Our codes are similar to the Gallager codes [4] in that they are built from
`sparse bipartite graphs. In contrast to Gallager codes, however, our codes will
`be constructed from graphs that have a highly irregular degree pattern on the
`left.
`Let G be a bipartite graph with n nodes on the left and n − k nodes on the
`right. G gives rise to a binary code of block-length n and dimension ≥ k in the
`
`4
`
`

`
`New Sequences of Erasure Codes
`
`67
`
`following way: let the adjacency matrix of the graph G be given as
`H⊤
`
`A =! 0
`0 " ,
`H
`where H is some (n− k)× n matrix describing the connections in the graph. The
`code defined by the graph is the code with parity check matrix H. A different
`way of describing the code is as follows: we index the coordinate positions of
`the code with the n nodes on the left hand side of the graph. The code consists
`of all binary vectors (c1, . . . , cn) such that for each right node in the graph the
`sum of the coordinate places adjacent to it equals zero. The block-length of this
`code equals n, and its dimension is at least k since we are imposing n − k linear
`conditions on the coordinates of a codeword. Expressed in terms of the graph,
`the fraction of redundant symbols in a codeword is at most aL/aR where aL and
`aR are the average node degrees on the left and the right hand side of the graph,
`respectively. In other words, the rate of the code is at least 1 − aL/aR. This
`description of the rate will be useful in later analysis. In the following, we will
`assume that the rate is in fact equal to this value. This is because the statements
`we will prove below will become even stronger if the rate is larger.
`The above construction needs asymptotically O(n2) arithmetic operations to
`find the encoding of a message of length k, if the graph is sparse. One can apply
`a trick to reduce the running time to O(n) by a modification of the construction.
`Details can be found in [8].
`Suppose now that a codeword (c1, . . . , cn) is sent and that certain erasures
`have occurred. The erasure recovery algorithm works as follows. We first initialize
`the contents of the right hand nodes of G with zero. Then we collect the non-
`erased coordinate positions, add their value to the current value of their right
`neighbors, and delete the left node and all edges emanating from it from the
`graph. After this stage, the graph consists of the erased nodes on the left and
`the edges emanating from these nodes. In the next step we look for a right node in
`the graph of degree one, i.e., a node that has only one edge coming out of it. We
`transport the value of this node to its unique left neighbor ℓ, thereby recovering
`the value of cℓ. We add cℓ to the current value of all the right neighbors of ℓ,
`delete the edges emanating from ℓ, and repeat the process until we cannot find a
`node of degree one on the right, or until all nodes on the left have been recovered.
`It is obvious that, on a RAM with unit cost measure, the amount of arithmetic
`operations to finish the algorithm is at most proportional to the number of edges
`in the graph, i.e., to naL, where aL is the average node degree on the left. The
`aim is thus to find graphs with constant aL for which the recovery algorithm
`finishes successfully.
`The main contribution of [8] was to give an analytic condition on the maxi-
`mum fraction of tolerable losses in terms of the degree distribution of the graph.
`More precisely, define the left and the right degree of an edge in the graph as
`the degree of the left, resp. right node it is emanating from. Further, denote
`by λi and ρi the fraction of edges of left, resp. right degree i, and consider the
`
`generating functions λ(x) :=#i λixi−1 and ρ(x) :=#i ρixi−1. In the following
`
`we will call the pair (λ, ρ) a degree distribution.
`
`5
`
`

`
`68
`
`M. Amin Shokrollahi
`
`Table 1. Rate 1/2 codes built from heavy tail/Poisson distribution.
`
`N
`8
`16
`27
`47
`79
`132
`221
`
`aR
`5.9266
`7.0788
`8.0054
`9.0256
`10.007
`10.996
`12.000
`

`5.9105
`7.0729
`8.0027
`9.0243
`10.007
`10.996
`12.000
`
`δ/(1 − R)
`0.91968
`0.95592
`0.97256
`0.98354
`0.98990
`0.99378
`0.99626
`

`0.45984
`0.47796
`0.48628
`0.49177
`0.49495
`0.49689
`0.49813
`
`ˆδ
`0.49085
`0.49609
`0.49799
`0.49902
`0.49951
`0.49975
`0.49988
`
`δ/ˆδ
`0.93682
`0.96345
`0.97648
`0.98547
`0.99087
`0.99427
`0.99650
`
`The main theorem of [8] states that if the graph is chosen at random with
`degree distribution (λ, ρ) and if the erasures occur at random positions, then
`the above erasure recovery algorithm can correct a δ-fraction of losses if ρ(1 −
`δλ(x)) ≥ 1 − x for x ∈ [0, 1]. In a later paper [6], this condition was slightly
`relaxed to
`(1)
`δλ (1 − ρ(1 − x)) < x
`for x ∈ (0,δ ].
`The paper [8] further exhibited for any ϵ> 0 and any rate R an infinite
`sequence of degree distributions (λ, ρ) giving rise to codes of rate at least R such
`that the above inequality is valid for δ = (1−R)(1−ϵ), and such that the average
`left degree of the graph is O(log(1/ϵ)). In other words, δ can get arbitrarily close
`to its optimal value (1 − R) with a “logarithmic” sacrifice in the running time
`of the algorithm. Explicitly, for any given 0 <ϵ <1, one chooses an integer N
`close to 1/ϵ and considers the pair (λN ,ρ N) with
`
`λN (x) =
`
`N−1
`
`xk
`k
`
`$k=1
`equal to 1 − R, and H(N − 1) is the harmonic sum #N−1
`
`,ρ N(x) = eθN·(x−1),
`
`(2)
`
`1
`H(N − 1)
`where θN is chosen to make the fraction of the nodes on the left and the right
`k=1 1/k. This sequence
`is referred to as the heavy tail/Poisson sequence in the following. Table 1 gives
`an example of the performance of this sequence for some codes of rate 1/2. In
`that table, aR denotes the average right degree, δ is the maximum tolerable
`loss fraction, and ˆδ is a theoretical upper bound on δ, see Remark 1 following
`Corollary 1.
`In the following we will show that this sort of relationship between the average
`degree and the maximum tolerable loss fraction is the best possible. Furthermore,
`we will exhibit a new sequence of degree distributions for which the same type of
`relationship holds between the average degree and the maximum tolerable loss
`fraction.
`
`6
`
`

`
`New Sequences of Erasure Codes
`
`69
`
`3 Upper Bounds for the Maximum Tolerable Loss
`Fraction
`
`In this section we will prove some upper bounds on the maximum tolerable loss
`fraction δ for which the algorithm given in the previous section is successful when
`applied to a random graph with a given degree distribution. Our main tool will
`be the inequality (1). We will first need a preliminary result.
`Lemma 1. Let G be a bipartite graph with edge degree distribution (λ, ρ). Then,
`the average left, resp. right, node degree of G equal
`1
`1
`0 λ(x)dx
`0 ρ(x)dx
`respectively. Let the polynomials Λ and R be defined by
`0 λ(t)dt
`0 ρ(t)dt
`0 λ(t)dt
`0 ρ(t)dt
`Then the coefficient of xi in Λ(x), resp.R (x) equals the fraction of nodes of
`degree i on the LHS, resp. RHS of G.
`
`,
`
`,
`
`% 1
`Λ(x) := % x
`% 1
`
`,
`
`% 1
`R(x) := % x
`% 1
`
`.
`
`Proof. Let L denote the number of nodes on the LHS of G. Then the number of
`edges in G equals aLL, and the number of edges of degree i equals λiaLL. Hence,
`the number of nodes of degree i on the LHS of the graph equals λiaLL/i, since
`each such node contributes to exactly i edges of degree i. Hence, the fraction of
`nodes of degree i equals λiaL/i. Since the sum of all these fractions equals 1, we
`obtain a−1
`0 λ(x)dx and this yields both assertions for the LHS
`of the graph. The assertions on the RHS follow similarly.
`⊓&
`
`L = #i λi/i = % 1
`
`The following theorem shows a rather strong upper bound on δ.
`Theorem 1. Let λ and ρ denote the right and left edge degree distributions of
`a bipartite graph. Let δ be a positive real number such that
`δλ (1 − ρ(1 − x)) ≤ x
`
`for 0 < x ≤ δ. Then we have
`aL
`(1 − (1 − δ)aR) ,
`δ ≤
`aR
`where aL and aR are the average node degrees of the graph on the left, resp. right,
`hand side.
`
`Proof. As a real valued function the polynomial λ(x) is strictly increasing for
`positive values of x. Hence it has a unique inverse λ−1 which is also strictly
`increasing. The first of the above inequalities is thus equivalent to
`1 − ρ(1 − x) ≤ λ−1 (x/δ)
`
`7
`
`

`
`70
`
`M. Amin Shokrollahi
`
`0
`
`0
`
`0
`
`0 λ(x)dx.) Invoking the previous lemma, we thus
`
`for 0 < x ≤ δ. Integrating both sides from 0 to δ and simplifying the expressions,
`we obtain
`ρ(x)dx −& 1−δ
`& 1
`ρ(x)dx ≥ δ& 1
`λ(x)dx.
`(Note that% 10 λ−1(x)dx = 1−% 1
`
`aR ’1 − % 1−δ
`0 ρ(x)dx ( = aL
`aR − aL& 1−δ
`ρ(x)dx = aL
`δ ≤
`% 1
`where the polynomial R is defined as in the statement of Lemma 1. To finish
`the proof we only need to show that R(1− δ) ≥ (1 − δ)aR. To see this, first note
`that if a1, . . . , aM are nonnegative real numbers adding up to 1 and µ is any
`nonnegative real number, then
`
`have
`
`aL
`
`0
`
`ρ(x)dx
`
`0
`
`(1 − R(1 − δ)) ,
`
`aR
`
`iai.
`
`(3)
`
`aiµi ≥ µ#i
`$i
`This follows from taking logarithms of both sides and noting that the log-function
`is concave. Denoting by ai the coefficients of R(x) and setting µ := 1 − δ, this
`gives our desired inequality: the ai are by the previous lemma the fractions of
`nodes of degree i, and hence #i iai is the average right degree aR.
`⊓&
`
`Corollary 1. With the same assumptions as in the previous theorem, we have
`
`aL
`δ ≤
`aR
`Proof. Follows from δ ≤ aL/aR.
`
`(1 − (1 − aL/aR)aR) .
`
`⊓&
`
`Remark 1. A more refined upper bound for δ is given by the following. Let r
`denote the quantity aL/aR, i.e., one minus the rate of the code. Suppose that
`aR > 1/r—an assumption that is automatically satisfied if the average left degree
`is larger than one—and consider the function f(x) = x−r(1−(1−x)aR). We have
`f(0) = 0, f(1) = 1 − r > 0, f′(0) < 0, and f′(x) has exactly one root. Hence, f
`has exactly one nonzero root, denoted by ˆδaR,r and this root is between 0 and 1.
`The previous theorem asserts that the maximum δ satisfying the inequality (1)
`is smaller than ˆδaR,aL/aR.
`The following definition is inspired by Corollary 1.
`Definition 1. A sequence (λm,ρ m) of degree distributions giving rise to codes
`of rate R ≥ R0 for a fixed R0 > 0 is called asymptotically quasi-optimal if there
`exists a constant µ (possibly depending on R) and for each m there exists a
`positive δm with δmλm(1 − ρm(1 − x)) ≤ x for x ∈ (0,δ m] such that the average
`right degree aR satisfies aR ≤ µ log(1 − δ/(1 − R)).
`
`8
`
`

`
`New Sequences of Erasure Codes
`
`71
`
`The significance of quasi-optimal is the following: we want to construct codes
`that can recover as many erasures as possible in as little as possible time. The
`running time of the erasure recovery algorithm is proportional to the block-
`length of the code times aR, the average right degree. Hence, we are looking for
`a trade-off between aR and δ. Corollary 1 shows that we cannot make aR too
`small. In fact, it implies that aR ≥ log(1− δ/(1− R))/ log R. However, we might
`hope that we can make aR smaller than log(1 − δ/(1 − R)) times a (negative)
`constant µ. In this way, we have maintained a qualitative optimality of the code
`sequence in the sense that the running time increases only logarithmically with
`the relative increase of the erasure correction capability.
`The heavy tail/Poisson sequence (2) is asymptotically quasi-optimal as is
`shown in [8]. Starting in the next section, we will introduce another quasi-optimal
`sequence which has certain advantages over the heavy tail/Poisson sequence.
`We close this section by stating another useful upper bound on the maximum
`possible value of δ.
`Lemma 2. Suppose that λ and ρ are polynomials and δ is such that δλ(1−ρ(1−
`x)) < x for x ∈ (0,δ ]. Then δ ≤ ρ′(1)/λ′(0).
`Proof. Let f(x) = δλ(1 − ρ(1 − x)) − x. By assumption, f(x) < 0 for x ∈ (0,δ ].
`This implies that f′(0) ≤ 0, where f′(x) is the derivative of f with respect to x.
`But f′(0) = δλ′(0)ρ′(1) − 1, which gives the assertion.
`⊓&
`4 Fractional Binomial Coefficients
`
`As can be seen from their description (2), the heavy tail/Poisson sequence is
`closely related to the Taylor series of − ln(1− x). The new sequence that we will
`describe in the next section will be related to the Taylor expansion of (1 − x)α
`where 1/α is an integer. The coefficients of this expansion are fractional binomial
`coefficients. For this reason, we will recall some well-known facts about these
`numbers.
`For real α and a positive integer N we define
`
`
`! αN" := α(α − 1)··· (α − N + 1)
`N !1 −
`N − 1"···)1 −
`= (−1)N−1 α
`* (1 − α) .
`For convenience, we also define +α
`0, := 1. We have the Taylor expansion
`
`=1!αk"(−1)k+1xk.
`1 − (1 − x)α =
`Note that for 0 <α < 1 the coefficients of the right hand power series above are
`all positive. Furthermore, we have
`
`α2
`
`N!
`

`
`(4)
`
`∞$k
`
`(5)
`
`
`
`! αN"(−1)N +1,
`
`N α
`
`N−1
`
`
`
`$k=1!αk"(−1)k+1 = 1 −
`
`9
`
`

`
`72
`
`M. Amin Shokrollahi
`
`and
`
`N−1
`
`k + 1
`
`N,(−1)N +1
`= α −+ α
`
`α + 1
`
`,
`
`(6)
`
`k"(−1)k+1
`$k=1!α
`
`cα
`
`as can easily be proved by induction.
`In the following, we will derive some estimates on the size of the binomial
`coefficients.
`Proposition 1. Let α be a positive real number less than 1/2 and let N ≥ 2 be
`an integer. There is a constant c independent of α and N such that
`
`N α+1 ≤! αN"(−1)N +1 ≤

`N α+1 .
`N,(−1)N +1 is positive. Taking its logarithm and expand-
`Proof. First note that+ α
`ing the series − ln(1 − x) =#i xi/i, we obtain
`
`ln!! αN"(−1)N +1" = ln(α/N) − α
`2 −··· .
`−
`
`(7)
`
`1 k
`N$k
`
`=1
`
`α2
`2
`
`1 k
`
`N$k
`
`=1
`
`(Note that the series involved are absolutely convergent, so that we can rearrange
`the orders of the summations.) For an upper bound on this sum we replace
`#N
`k=1 1/ks by 1 for s ≥ 2 and use the left hand side of the inequality
`1
`2N
`
`1 k
`
`N$k
`
`=1
`
`ln(N) + γ<
`
`< ln(N) + γ +
`
`,
`
`This gives, after exponentiation,
`

`
`where γ is Euler’s constant [5, pp. 480–481]. For the lower bound we use the
`right hand side of the above inequality and replace #N
`k=1 1/ks for s ≥ 2 by 2.
`
`N α+1 eα(2−γ−1/2N )(1 − α)2 ≤! αN"(−1)N +1 ≤

`N α+1 eα(1−γ)(1 − α).
`Noting that 0 <α ≤ 1/2, we obtain our assertion.
`
`⊓&
`
`5 Right Regular Sequences
`
`A closer look at the proof of Theorem 1 reveals the following: one of the reasons
`we might obtain a lower than optimal upper bound for the largest δ satisfying (1)
`is that the inequality (3) is not sharp. In fact, that inequality is sharp iff all
`the ai except for one are zero, i.e., iff ρ(x) has only one nonzero coefficient,
`i.e., iff the graph has only one degree on the right hand side. We call such
`graphs right regular in the following. We will study in this section right regular
`degree distributions and design the left hand side in such a way as to obtain
`asymptotically quasi-optimal sequences. We remark that right regular sequences
`
`10
`
`

`
`New Sequences of Erasure Codes
`
`73
`
`,
`
`(8)
`
`(x).
`
`(9)
`
`were previously studied in an unpublished manuscript [9]. Our approach differs
`however from the one given in that paper.
`From a theoretical point of view, codes obtained from these sequences should
`perform better than the heavy tail/Poisson distribution, since they allow for a
`larger loss-fraction for a given rate and a given average left degree. Furthermore,
`the analysis given in [6] suggests that the actual performance of the code is
`related to how accurately the neighborhood of a message node is described by a
`tree given by the degree distributions. For instance, the performance of regular
`graphs is much more sharply concentrated around the value predicted by the
`theoretical analysis, than the performance of irregular graphs. Hence, if the graph
`is right regular, one should expect a smaller variance in the actual performance
`than for corresponding irregular codes.
`For integers a ≥ 2 and N ≥ 2 we define
`α,N (x) := α#N−1
`k=1 +αk,(−1)k+1xk
`
`ρa(x) := xa−1,λ
`N,(−1)N +1
`α − N+ α
`where here and in the following we set α := 1/(a− 1). Note that ρa(1) = 1, that
`λa,N(1) = 1 by (5), and that λa,N has positive coefficients. Hence, (ρa,λ a,N )
`indeed defines a degree distribution.
`Let ν< 1 be a positive constant. The sequence we are interested in is given
`by the following degree distributions:
`ρa(x),λ
`a,⌊ν−1/α⌋
`First we consider the rate of the codes defined by these distributions.
`Proposition 2. The rate R = Ra,µ of the code defined by the distributions
`in (9) satisfies
`1 − cν
`1 − ν
`1 − cνν 1/α ≤ 1 − R ≤
`1 − νν1/α ,
`where c is the constant from Proposition 2.
`Proof. In the following, we denote by N the quantity ⌊ν−1/α⌋ and by λ(x) the
`function λa,N(x). We need to compute the integral of λ(x). We have
`N,(−1)N +1
`α −+ α
`& 1
`N,(−1)N +1 .
`α − N+ α
`N,(−1)N +1
`α − N+ α
`1 −
`:= 1 − rα,N .
`N,(−1)N +1
`α −+ α
`Next, we estimate rα,N using Proposition 2 again:
`1 − ν
`1 − cν
`1 − cνν 1/α ≤ rα,N ≤
`1 − νν1/α .
`This gives the desired assertion on the rate.
`
`0 ρa(x)dx = 1/a = α/(α + 1). Hence, the rate of the code is
`
`λ(x)dx = α
`α + 1
`
`0
`
`Further,% 1
`
`(10)
`
`⊓&
`
`11
`
`

`
`74
`
`M. Amin Shokrollahi
`
`δα
`
`.
`
`Theorem 2. The sequence of degree distributions given in (9) is asymptotically
`quasi-optimal.
`Proof. Let us compute the maximum value of δ such that δλ(1 − ρ(1 − x)) < x
`for x ∈ (0,δ ), where (λ, ρ) is the pair given in (9). We have
`δα
`δλ(1 − ρ(1 − x)) <
`(1 − ρ(1 − x)α)
`N,(−1)N +1
`α − N+ α
`=
`N,(−1)N +1 x.
`α − N+ α
`So, for δλ(1 − ρ(1 − x)) < x we need
`N,(−1)N +1
`α − N+ α
`δ ≤

`This is, by the way, the same upper bound as the one obtained from Lemma 2. In
`the following we assume that δ is equal to the above upper bound, and compute
`δ/(1 − R), R being the rate of code estimated in Proposition 2:
`N,(−1)N +1
`= α −+ α
`1

`≥ 1 −
`N α+1 .
`1 − R

`Ignoring diophantine constraints, we assume in the following that N = ν−1/α.
`This gives
`
`(11)
`

`
`1 − R ≥ 1 − ν(α+1)/α = 1 − νaR,
`where aR = a = (α + 1)/α is the average right degree of the code. This proves
`the assertion.
`⊓&
`In practical situations, one is interested in designing a code of a given fixed rate.
`Since the parameter α in the definition of the sequence in (9) is the inverse
`of an integer, the range of this parameter is discrete. Hence, we do not have
`a continuous range of rates for these codes. However, we can come arbitrarily
`close to our desired rate by making α smaller, i.e., by allowing high degrees
`on the RHS of the graph. Some examples for different target rates are given in
`Table 2. The value of N has been hand-optimized in these examples to come as
`close to the desired rate as possible. The last column in that table corresponds
`to the value ˆδ defined in Remark 1 following Corollary 1. It gives a theoretical
`upper bound on the best value of δ. As can be observed from the table, the
`maximum value of δ converges very quickly to the maximum possible value.
`Also, a comparison between these codes and the heavy tail/Poisson distribution
`in Table 1 reveals that the new codes are better in terms of the trade-off between
`δ and aR. However, the new codes need larger degrees on the left than the heavy
`tail/Poisson distribution.
`To obtain codes that have a fixed rate, one can modify α slightly. Such ex-
`amples are given in Table 3. Both parameters (α and N) of these codes are
`hand-optimized to give the best results.
`
`12
`
`

`
`Table 2. Right regular sequences for rates R close to 2/3, 1/2, and 1/3.
`
`New Sequences of Erasure Codes
`
`75
`
`N
`2
`3
`6
`11
`17
`27
`42
`64
`13
`29
`60
`12.
`257
`523
`1058
`111
`349
`1077
`3298
`
`aR
`6
`7
`8
`9
`10
`11
`12
`13
`6
`7
`8
`9
`10
`11
`12
`6
`7
`8
`9
`
`1 − Rδ
`0.33333
`0.31677
`0.32886
`0.33645
`0.33357
`0.33392
`0.33381
`0.33312
`0.50090
`0.50164
`0.49965
`0.49985
`0.50000
`0.50002
`0.49999
`0.66677
`0.66667
`0.66663
`0.66669
`
`/(1 − R)
`0.60000
`0.74537
`0.88166
`0.93777
`0.96001
`0.97502
`0.98401
`0.98953
`0.96007
`0.98251
`0.99159
`0.99598
`0.99805
`0.99904
`0.99953
`0.99698
`0.99904
`0.99969
`0.99990
`

`0.20000
`0.23611
`0.28994
`0.31551
`0.32024
`0.32558
`0.32847
`0.32963
`0.48090
`0.49287
`0.49545
`0.49784
`0.49903
`0.49954
`0.49975
`0.66475
`0.66603
`0.66642
`0.66662
`
`ˆδ
`0.29099
`0.28714
`0.31243
`0.32690
`0.32724
`0.32984
`0.33113
`0.33134
`0.49232
`0.49759
`0.49762
`0.49885
`0.49951
`0.49977
`0.49986
`0.66584
`0.66636
`0.66653
`0.66665
`
`δ/ˆδ
`0.68731
`0.82230
`0.92801
`0.96514
`0.97860
`0.98711
`0.99197
`0.99484
`0.97679
`0.99052
`0.99563
`0.99797
`0.99904
`0.99953
`0.99977
`0.99837
`0.99950
`0.99984
`0.99995
`
`6 Conclusions and Open Questions
`
`In this paper we have analyzed the theoretical performance of a simple erasure
`recovery algorithm applied to Gallager codes by deriving upper bounds on the
`maximum fraction of tolerable losses given a parameter that describes the run-
`ning time for encoding and decoding of the codes. We have shown that there is
`a trade-off between proximity to the optimal value of tolerable losses, i.e., one
`minus the rate of the code, and the average degree of nodes in the graph, in
`the sense that multiplying the average degree by a constant factor implies an
`exponential relative increase of the maximum tolerable loss fraction. Further,
`we have introduced a new sequence of graphs which are asymptotically close to
`optimal with respect to this criterion. These graphs are right regular and their
`node degree distribution on the left hand side is closely related to the power
`series expansion of (1 − x)α, where α is the inverse of an integer. Previously,
`the only known such sequence was the heavy tail/Poisson sequence introduced
`in [8]. We have included examples which show that the new codes tolerate a
`higher fraction of losses if the average degree of the graph is fixed.
`It would be very interesting to extend the analysis given in this paper to
`other decoding algorithms, e.g., to the simple decoding algorithm of Gallager
`for error correcting codes [4,7]. Such an analysis would probably give clues for
`constructing infinite sequences of graphs that perform asymptotically optimally
`with respect to the decoding algorithm in question.
`
`13
`
`

`
`76
`
`M. Amin Shokrollahi
`
`Table 3. Rate 1/2 codes, ρ(x) = xaR−1, λα,N (x) as defined in (8) for arbitrary α.
`
`Nα
`11
`27
`59
`124
`256
`521
`1057
`
`0.17662
`0.16412
`0.14225
`0.12480
`0.11112
`0.09993
`0.09090
`
`aR
`6
`7
`8
`9
`10
`11
`12
`
`δ/(1 − R)
`0.95982
`0.98190
`0.99142
`0.99596
`0.99800
`0.99896
`0.99950
`

`0.47991
`0.49095
`0.49571
`0.49798
`0.49900
`0.49948
`0.49975
`
`ˆδ
`0.49134
`0.49586
`0.49798
`0.49901
`0.49951
`0.49975
`0.49988
`
`δ/ˆδ
`0.97673
`0.99009
`0.99543
`0.99794
`0.99898
`0.99945
`0.99974
`
`References
`
`1. N. Alon and M. Luby. A linear time erasure-resilient code with nearly optimal
`recovery. IEEE Trans. Inform. Theory, 42:1732–1736, 1996.
`2. R.E. Blahut. Theory and Practice of Error Control Codes. Addison Wesley, Read-
`ing, MA, 1983.
`3. P. Elias. Coding for two noisy channels. In Information Theory, Third London
`Symposium, pages 61–76, 1955.
`4. R. G. Gallager. Low Density Parity-Check Codes. MIT Press, Cambridge, MA,
`1963.
`5. R.L. Graham, D.E. Knuth, and P. Patashnik. Concrete Mathematics. Addison-
`Wesley, 1994.
`6. M. Luby, M. Mitzenmacher, and M.A. Shokrollahi. Analysis of random processes
`via and-or tree evaluation. In Proceedings of the 9th Annual ACM-SIAM Sympo-
`sium on Discrete Algorithms, pages 364–373, 1998.
`7. M. Luby, M. Mitzenmacher, M.A. Shokrollahi, and D. Spielman. Analysis of low
`density codes and improved designs using irregular graphs. In Proceedings of the
`30th Annual ACM Symposium on Theory of Computing, pages 249–258, 1998.
`8. M. Luby, M. Mitzenmacher, M.A. Shokrollahi, D. Spielman, and V. Stemann.
`Practical loss-resilient codes. In Proceedings of the 29th annual ACM Symposium
`on Theory of Computing, pages 150–159, 1997.
`9. M. Luby and M.A. Shokrollahi. Right regular erasure codes. Unpublished, 1998.
`10. F.J. MacWilliams and N.J.A. Sloane. The Theory of Error-Correcting Codes.
`North-Holland, 1988.
`
`14

This document is available on Docket Alarm but you must sign up to view it.


Or .

Accessing this document will incur an additional charge of $.

After purchase, you can access this document again without charge.

Accept $ Charge
throbber

Still Working On It

This document is taking longer than usual to download. This can happen if we need to contact the court directly to obtain the document and their servers are running slowly.

Give it another minute or two to complete, and then try the refresh button.

throbber

A few More Minutes ... Still Working

It can take up to 5 minutes for us to download a document if the court servers are running slowly.

Thank you for your continued patience.

This document could not be displayed.

We could not find this document within its docket. Please go back to the docket page and check the link. If that does not work, go back to the docket and refresh it to pull the newest information.

Your account does not support viewing this document.

You need a Paid Account to view this document. Click here to change your account type.

Your account does not support viewing this document.

Set your membership status to view this document.

With a Docket Alarm membership, you'll get a whole lot more, including:

  • Up-to-date information for this case.
  • Email alerts whenever there is an update.
  • Full text search for other cases.
  • Get email alerts whenever a new case matches your search.

Become a Member

One Moment Please

The filing “” is large (MB) and is being downloaded.

Please refresh this page in a few minutes to see if the filing has been downloaded. The filing will also be emailed to you when the download completes.

Your document is on its way!

If you do not receive the document in five minutes, contact support at support@docketalarm.com.

Sealed Document

We are unable to display this document, it may be under a court ordered seal.

If you have proper credentials to access the file, you may proceed directly to the court's system using your government issued username and password.


Access Government Site

We are redirecting you
to a mobile optimized page.





Document Unreadable or Corrupt

Refresh this Document
Go to the Docket

We are unable to display this document.

Refresh this Document
Go to the Docket