Low-density parity-check code

Last updated

In information theory, a low-density parity-check (LDPC) code is a linear error correcting code, a method of transmitting a message over a noisy transmission channel. [1] [2] An LDPC code is constructed using a sparse Tanner graph (subclass of the bipartite graph). [3] LDPC codes are capacity-approaching codes, which means that practical constructions exist that allow the noise threshold to be set very close to the theoretical maximum (the Shannon limit) for a symmetric memoryless channel. The noise threshold defines an upper bound for the channel noise, up to which the probability of lost information can be made as small as desired. Using iterative belief propagation techniques, LDPC codes can be decoded in time linear in their block length.

Contents

LDPC codes are also known as Gallager codes, in honor of Robert G. Gallager, who developed the LDPC concept in his doctoral dissertation at the Massachusetts Institute of Technology in 1960. [4] [5] However, LDPC codes require computationally expensive iterative decoding, so they went unused for decades. In 1993 the newly invented turbo codes demonstrated that codes with iterative decoding could far outperform other codes used at that time, but turbo codes were patented and required a fee for use. This raised renewed interest in LDPC codes, which were shown to have similar performance, but were much older and patent-free. [6] Now that the fundamental patent for turbo codes has expired (on August 29, 2013), [7] [8] LDPC codes are still used for their technical merits.

LDPC codes have been shown to have ideal combinatorial properties. In his dissertation, Gallager showed that LDPC codes achieve the Gilbert–Varshamov bound for linear codes over binary fields with high probability. In 2020 it was shown that Gallager's LDPC codes achieve list decoding capacity and also achieve the Gilbert–Varshamov bound for linear codes over general fields. [9]

History

Impractical to implement when first developed by Gallager in 1963, [10] LDPC codes were forgotten until his work was rediscovered in 1996. [11] Turbo codes, another class of capacity-approaching codes discovered in 1993, became the coding scheme of choice in the late 1990s, used for applications such as the Deep Space Network and satellite communications. LDPC codes then received renewed interest as a patent-free alternative of similar performance. [6] Since then, advances in low-density parity-check codes have seen them surpass turbo codes in terms of error floor and performance in the higher code rate range, leaving turbo codes better suited for the lower code rates only. [12]

Applications

In 2003, an irregular repeat accumulate (IRA) style LDPC code beat six turbo codes to become the error-correcting code in the new DVB-S2 standard for digital television. [13] The DVB-S2 selection committee made decoder complexity estimates for the turbo code proposals using a much less efficient serial decoder architecture rather than a parallel decoder architecture. This forced the turbo code proposals to use frame sizes on the order of one half the frame size of the LDPC proposals.[ citation needed ]

In 2008, LDPC beat convolutional turbo codes as the forward error correction (FEC) system for the ITU-T G.hn standard. [14] G.hn chose LDPC codes over turbo codes because of their lower decoding complexity (especially when operating at data rates close to 1.0 Gbit/s) and because the proposed turbo codes exhibited a significant error floor at the desired range of operation. [15]

LDPC codes are also used for 10GBASE-T Ethernet, which sends data at 10 gigabits per second over twisted-pair cables. As of 2009, LDPC codes are also part of the Wi-Fi 802.11 standard as an optional part of 802.11n and 802.11ac, in the High Throughput (HT) PHY specification. [16] LDPC is a mandatory part of 802.11ax (Wi-Fi 6). [17]

Some OFDM systems add an additional outer error correction that fixes the occasional errors (the "error floor") that get past the LDPC correction inner code even at low bit error rates.

For example: The Reed-Solomon code with LDPC Coded Modulation (RS-LCM) uses a Reed-Solomon outer code. [18] The DVB-S2, the DVB-T2 and the DVB-C2 standards all use a BCH code outer code to mop up residual errors after LDPC decoding. [19]

5G NR uses polar code for the control channels and LDPC for the data channels. [20] [21]

Although LDPC code has had its success in commercial hard disk drives, to fully exploit its error correction capability in SSDs demands unconventional fine-grained flash memory sensing, leading to an increased memory read latency. LDPC-in-SSD [22] is an effective approach to deploy LDPC in SSD with a very small latency increase, which turns LDPC in SSD into a reality. Since then, LDPC has been widely adopted in commercial SSDs in both customer-grades and enterprise-grades by major storage venders. Many TLC (and later) SSDs are using LDPC codes. A fast hard-decode (binary erasure) is first attempted, which can fall back into the slower but more powerful soft decoding. [23]

Operational use

LDPC codes functionally are defined by a sparse parity-check matrix. This sparse matrix is often randomly generated, subject to the sparsity constraints—LDPC code construction is discussed later. These codes were first designed by Robert Gallager in 1960. [5]

Below is a graph fragment of an example LDPC code using Forney's factor graph notation. In this graph, n variable nodes in the top of the graph are connected to (nk) constraint nodes in the bottom of the graph.

This is a popular way of graphically representing an (n, k) LDPC code. The bits of a valid message, when placed on the T's at the top of the graph, satisfy the graphical constraints. Specifically, all lines connecting to a variable node (box with an '=' sign) have the same value, and all values connecting to a factor node (box with a '+' sign) must sum, modulo two, to zero (in other words, they must sum to an even number; or there must be an even number of odd values).

Ldpc code fragment factor graph.svg

Ignoring any lines going out of the picture, there are eight possible six-bit strings corresponding to valid codewords: (i.e., 000000, 011001, 110010, 101011, 111100, 100101, 001110, 010111). This LDPC code fragment represents a three-bit message encoded as six bits. Redundancy is used, here, to increase the chance of recovering from channel errors. This is a (6, 3) linear code, with n = 6 and k = 3.

Again ignoring lines going out of the picture, the parity-check matrix representing this graph fragment is

In this matrix, each row represents one of the three parity-check constraints, while each column represents one of the six bits in the received codeword.

In this example, the eight codewords can be obtained by putting the parity-check matrix H into this form through basic row operations in GF(2):

Step 1: H.

Step 2: Row 1 is added to row 3.

Step 3: Row 2 and 3 are swapped.

Step 4: Row 1 is added to row 3.

From this, the generator matrix G can be obtained as (noting that in the special case of this being a binary code ), or specifically:

Finally, by multiplying all eight possible 3-bit strings by G, all eight valid codewords are obtained. For example, the codeword for the bit-string '101' is obtained by:

,

where is symbol of mod 2 multiplication.

As a check, the row space of G is orthogonal to H such that

The bit-string '101' is found in as the first 3 bits of the codeword '101011'.

Example encoder

LDPC encoder LDPC encoder Figure.png
LDPC encoder

During the encoding of a frame, the input data bits (D) are repeated and distributed to a set of constituent encoders. The constituent encoders are typically accumulators and each accumulator is used to generate a parity symbol. A single copy of the original data (S0,K-1) is transmitted with the parity bits (P) to make up the code symbols. The S bits from each constituent encoder are discarded.

The parity bit may be used within another constituent code.

In an example using the DVB-S2 rate 2/3 code the encoded block size is 64800 symbols (N=64800) with 43200 data bits (K=43200) and 21600 parity bits (M=21600). Each constituent code (check node) encodes 16 data bits except for the first parity bit which encodes 8 data bits. The first 4680 data bits are repeated 13 times (used in 13 parity codes), while the remaining data bits are used in 3 parity codes (irregular LDPC code).

For comparison, classic turbo codes typically use two constituent codes configured in parallel, each of which encodes the entire input block (K) of data bits. These constituent encoders are recursive convolutional codes (RSC) of moderate depth (8 or 16 states) that are separated by a code interleaver which interleaves one copy of the frame.

The LDPC code, in contrast, uses many low depth constituent codes (accumulators) in parallel, each of which encode only a small portion of the input frame. The many constituent codes can be viewed as many low depth (2 state) "convolutional codes" that are connected via the repeat and distribute operations. The repeat and distribute operations perform the function of the interleaver in the turbo code.

The ability to more precisely manage the connections of the various constituent codes and the level of redundancy for each input bit give more flexibility in the design of LDPC codes, which can lead to better performance than turbo codes in some instances. Turbo codes still seem to perform better than LDPCs at low code rates, or at least the design of well performing low rate codes is easier for turbo codes.

As a practical matter, the hardware that forms the accumulators is reused during the encoding process. That is, once a first set of parity bits are generated and the parity bits stored, the same accumulator hardware is used to generate a next set of parity bits.

Decoding

As with other codes, the maximum likelihood decoding of an LDPC code on the binary symmetric channel is an NP-complete problem, [24] shown by reduction from 3-dimensional matching. So assuming P != NP, which is widely believed, then performing optimal decoding for an arbitrary code of any useful size is not practical.

However, sub-optimal techniques based on iterative belief propagation decoding give excellent results and can be practically implemented. The sub-optimal decoding techniques view each parity check that makes up the LDPC as an independent single parity check (SPC) code. Each SPC code is decoded separately using soft-in-soft-out (SISO) techniques such as SOVA, BCJR, MAP, and other derivates thereof. The soft decision information from each SISO decoding is cross-checked and updated with other redundant SPC decodings of the same information bit. Each SPC code is then decoded again using the updated soft decision information. This process is iterated until a valid codeword is achieved or decoding is exhausted. This type of decoding is often referred to as sum-product decoding.

The decoding of the SPC codes is often referred to as the "check node" processing, and the cross-checking of the variables is often referred to as the "variable-node" processing.

In a practical LDPC decoder implementation, sets of SPC codes are decoded in parallel to increase throughput.

In contrast, belief propagation on the binary erasure channel is particularly simple where it consists of iterative constraint satisfaction.

For example, consider that the valid codeword, 101011, from the example above, is transmitted across a binary erasure channel and received with the first and fourth bit erased to yield ?01?11. Since the transmitted message must have satisfied the code constraints, the message can be represented by writing the received message on the top of the factor graph.

In this example, the first bit cannot yet be recovered, because all of the constraints connected to it have more than one unknown bit. In order to proceed with decoding the message, constraints connecting to only one of the erased bits must be identified. In this example, only the second constraint suffices. Examining the second constraint, the fourth bit must have been zero, since only a zero in that position would satisfy the constraint.

This procedure is then iterated. The new value for the fourth bit can now be used in conjunction with the first constraint to recover the first bit as seen below. This means that the first bit must be a one to satisfy the leftmost constraint.

Ldpc code fragment factor graph w erasures decode step 2.svg

Thus, the message can be decoded iteratively. For other channel models, the messages passed between the variable nodes and check nodes are real numbers, which express probabilities and likelihoods of belief.

This result can be validated by multiplying the corrected codeword r by the parity-check matrix H:

Because the outcome z (the syndrome) of this operation is the three × one zero vector, the resulting codeword r is successfully validated.

After the decoding is completed, the original message bits '101' can be extracted by looking at the first 3 bits of the codeword.

While illustrative, this erasure example does not show the use of soft-decision decoding or soft-decision message passing, which is used in virtually all commercial LDPC decoders.

Updating node information

In recent years[ when? ], there has also been a great deal of work spent studying the effects of alternative schedules for variable-node and constraint-node update. The original technique that was used for decoding LDPC codes was known as flooding. This type of update required that, before updating a variable node, all constraint nodes needed to be updated and vice versa. In later work by Vila Casado et al., [25] alternative update techniques were studied, in which variable nodes are updated with the newest available check-node information.[ citation needed ]

The intuition behind these algorithms is that variable nodes whose values vary the most are the ones that need to be updated first. Highly reliable nodes, whose log-likelihood ratio (LLR) magnitude is large and does not change significantly from one update to the next, do not require updates with the same frequency as other nodes, whose sign and magnitude fluctuate more widely.[ citation needed ] These scheduling algorithms show greater speed of convergence and lower error floors than those that use flooding. These lower error floors are achieved by the ability of the Informed Dynamic Scheduling (IDS) [25] algorithm to overcome trapping sets of near codewords. [26]

When nonflooding scheduling algorithms are used, an alternative definition of iteration is used. For an (n, k) LDPC code of rate k/n, a full iteration occurs when n variable and n  k constraint nodes have been updated, no matter the order in which they were updated.[ citation needed ]

Code construction

For large block sizes, LDPC codes are commonly constructed by first studying the behaviour of decoders. As the block size tends to infinity, LDPC decoders can be shown to have a noise threshold below which decoding is reliably achieved, and above which decoding is not achieved, [27] colloquially referred to as the cliff effect. This threshold can be optimised by finding the best proportion of arcs from check nodes and arcs from variable nodes. An approximate graphical approach to visualising this threshold is an EXIT chart.[ citation needed ]

The construction of a specific LDPC code after this optimization falls into two main types of techniques:[ citation needed ]

Construction by a pseudo-random approach builds on theoretical results that, for large block size, a random construction gives good decoding performance. [11] In general, pseudorandom codes have complex encoders, but pseudorandom codes with the best decoders can have simple encoders. [28] Various constraints are often applied to help ensure that the desired properties expected at the theoretical limit of infinite block size occur at a finite block size.[ citation needed ]

Combinatorial approaches can be used to optimize the properties of small block-size LDPC codes or to create codes with simple encoders.[ citation needed ]

Some LDPC codes are based on Reed–Solomon codes, such as the RS-LDPC code used in the 10 Gigabit Ethernet standard. [29] Compared to randomly generated LDPC codes, structured LDPC codes—such as the LDPC code used in the DVB-S2 standard—can have simpler and therefore lower-cost hardware—in particular, codes constructed such that the H matrix is a circulant matrix. [30]

Yet another way of constructing LDPC codes is to use finite geometries. This method was proposed by Y. Kou et al. in 2001. [31]

Compared to turbo codes

LDPC codes can be compared with other powerful coding schemes, e.g. turbo codes. [32] In one hand, BER performance of turbo codes is influenced by low codes limitations. [33] LDPC codes have no limitations of minimum distance, [34] that indirectly means that LDPC codes may be more efficient on relatively large code rates (e.g. 3/4, 5/6, 7/8) than turbo codes. However, LDPC codes are not the complete replacement: turbo codes are the best solution at the lower code rates (e.g. 1/6, 1/3, 1/2). [35] [36]

See also

People

Theory

Applications

Other capacity-approaching codes

Related Research Articles

<span class="mw-page-title-main">Error detection and correction</span> Techniques that enable reliable delivery of digital data over unreliable communication channels

In information theory and coding theory with applications in computer science and telecommunication, error detection and correction (EDAC) or error control are techniques that enable reliable delivery of digital data over unreliable communication channels. Many communication channels are subject to channel noise, and thus errors may be introduced during transmission from the source to a receiver. Error detection techniques allow detecting such errors, while error correction enables reconstruction of the original data in many cases.

In telecommunication, a convolutional code is a type of error-correcting code that generates parity symbols via the sliding application of a boolean polynomial function to a data stream. The sliding application represents the 'convolution' of the encoder over the data, which gives rise to the term 'convolutional coding'. The sliding nature of the convolutional codes facilitates trellis decoding using a time-invariant trellis. Time invariant trellis decoding allows convolutional codes to be maximum-likelihood soft-decision decoded with reasonable complexity.

<span class="mw-page-title-main">Hamming code</span> Family of linear error-correcting codes

In computer science and telecommunication, Hamming codes are a family of linear error-correcting codes. Hamming codes can detect one-bit and two-bit errors, or correct one-bit errors without detection of uncorrected errors. By contrast, the simple parity code cannot correct errors, and can detect only an odd number of bits in error. Hamming codes are perfect codes, that is, they achieve the highest possible rate for codes with their block length and minimum distance of three. Richard W. Hamming invented Hamming codes in 1950 as a way of automatically correcting errors introduced by punched card readers. In his original paper, Hamming elaborated his general idea, but specifically focused on the Hamming(7,4) code which adds three parity bits to four bits of data.

In information theory, turbo codes are a class of high-performance forward error correction (FEC) codes developed around 1990–91, but first published in 1993. They were the first practical codes to closely approach the maximum channel capacity or Shannon limit, a theoretical maximum for the code rate at which reliable communication is still possible given a specific noise level. Turbo codes are used in 3G/4G mobile communications and in satellite communications as well as other applications where designers seek to achieve reliable information transfer over bandwidth- or latency-constrained communication links in the presence of data-corrupting noise. Turbo codes compete with low-density parity-check (LDPC) codes, which provide similar performance.

Claude Berrou is a French professor in electrical engineering at École Nationale Supérieure des Télécommunications de Bretagne, now IMT Atlantique. He is the sole inventor of a groundbreaking quasi-optimal error-correcting coding scheme called turbo codes as evidenced by the sole inventorship credit given on the fundamental patent for turbo codes. The original patent filing for turbo codes issued in the US as US Patent 5,446,747.

A Sparse graph code is a code which is represented by a sparse graph.

In coding theory, a linear code is an error-correcting code for which any linear combination of codewords is also a codeword. Linear codes are traditionally partitioned into block codes and convolutional codes, although turbo codes can be seen as a hybrid of these two types. Linear codes allow for more efficient encoding and decoding algorithms than other codes.

In information theory, the noisy-channel coding theorem, establishes that for any given degree of noise contamination of a communication channel, it is possible to communicate discrete data nearly error-free up to a computable maximum rate through the channel. This result was presented by Claude Shannon in 1948 and was based in part on earlier work and ideas of Harry Nyquist and Ralph Hartley.

Hybrid automatic repeat request is a combination of high-rate forward error correction (FEC) and automatic repeat request (ARQ) error-control. In standard ARQ, redundant bits are added to data to be transmitted using an error-detecting (ED) code such as a cyclic redundancy check (CRC). Receivers detecting a corrupted message will request a new message from the sender. In Hybrid ARQ, the original data is encoded with an FEC code, and the parity bits are either immediately sent along with the message or only transmitted upon request when a receiver detects an erroneous message. The ED code may be omitted when a code is used that can perform both forward error correction (FEC) in addition to error detection, such as a Reed–Solomon code. The FEC code is chosen to correct an expected subset of all errors that may occur, while the ARQ method is used as a fall-back to correct errors that are uncorrectable using only the redundancy sent in the initial transmission. As a result, hybrid ARQ performs better than ordinary ARQ in poor signal conditions, but in its simplest form this comes at the expense of significantly lower throughput in good signal conditions. There is typically a signal quality cross-over point below which simple hybrid ARQ is better, and above which basic ARQ is better.

In computing, telecommunication, information theory, and coding theory, forward error correction (FEC) or channel coding is a technique used for controlling errors in data transmission over unreliable or noisy communication channels.

In computer science, repeat-accumulate codes are a low complexity class of error-correcting codes. They were devised so that their ensemble weight distributions are easy to derive. RA codes were introduced by Divsalar et al.

The error floor is a phenomenon encountered in modern iterated sparse graph-based error correcting codes like LDPC codes and turbo codes. When the bit error ratio (BER) is plotted for conventional codes like Reed–Solomon codes under algebraic decoding or for convolutional codes under Viterbi decoding, the BER steadily decreases in the form of a curve as the SNR condition becomes better. For LDPC codes and turbo codes there is a point after which the curve does not fall as quickly as before, in other words, there is a region in which performance flattens. This region is called the error floor region. The region just before the sudden drop in performance is called the waterfall region.

<span class="mw-page-title-main">EXIT chart</span>

An extrinsic information transfer chart, commonly called an EXIT chart, is a technique to aid the construction of good iteratively-decoded error-correcting codes.

<span class="mw-page-title-main">Hamming(7,4)</span> Linear error-correcting code

In coding theory, Hamming(7,4) is a linear error-correcting code that encodes four bits of data into seven bits by adding three parity bits. It is a member of a larger family of Hamming codes, but the term Hamming code often refers to this specific code that Richard W. Hamming introduced in 1950. At the time, Hamming worked at Bell Telephone Laboratories and was frustrated with the error-prone punched card reader, which is why he started working on error-correcting codes.

In coding theory, concatenated codes form a class of error-correcting codes that are derived by combining an inner code and an outer code. They were conceived in 1966 by Dave Forney as a solution to the problem of finding a code that has both exponentially decreasing error probability with increasing block length and polynomial-time decoding complexity. Concatenated codes became widely used in space communications in the 1970s.

<span class="mw-page-title-main">Expander code</span>

In coding theory, expander codes form a class of error-correcting codes that are constructed from bipartite expander graphs. Along with Justesen codes, expander codes are of particular interest since they have a constant positive rate, a constant positive relative distance, and a constant alphabet size. In fact, the alphabet contains only two elements, so expander codes belong to the class of binary codes. Furthermore, expander codes can be both encoded and decoded in time proportional to the block length of the code.

In information theory, a polar code is a linear block error-correcting code. The code construction is based on a multiple recursive concatenation of a short kernel code which transforms the physical channel into virtual outer channels. When the number of recursions becomes large, the virtual channels tend to either have high reliability or low reliability (in other words, they polarize or become sparse), and the data bits are allocated to the most reliable channels. It is the first code with an explicit construction to provably achieve the channel capacity for symmetric binary-input, discrete, memoryless channels (B-DMC) with polynomial dependence on the gap to capacity. Notably, polar codes have modest encoding and decoding complexity O(n log n), which renders them attractive for many applications. Moreover, the encoding and decoding energy complexity of generalized polar codes can reach the fundamental lower bounds for energy consumption of two dimensional circuitry to within an O(nε polylog n) factor for any ε > 0.

In coding theory, Zemor's algorithm, designed and developed by Gilles Zemor, is a recursive low-complexity approach to code construction. It is an improvement over the algorithm of Sipser and Spielman.

Serial concatenated convolutional codes (SCCC) are a class of forward error correction (FEC) codes highly suitable for turbo (iterative) decoding. Data to be transmitted over a noisy channel may first be encoded using an SCCC. Upon reception, the coding may be used to remove any errors introduced during transmission. The decoding is performed by repeated decoding and [de]interleaving of the received symbols.

Sudoku codes are non-linear forward error correcting codes following rules of sudoku puzzles designed for an erasure channel. Based on this model, the transmitter sends a sequence of all symbols of a solved sudoku. The receiver either receives a symbol correctly or an erasure symbol to indicate that the symbol was not received. The decoder gets a matrix with missing entries and uses the constraints of sudoku puzzles to reconstruct a limited amount of erased symbols.

References

  1. MacKay, David J. (2003). Information theory, Inference and Learning Algorithms. Cambridge University Press. ISBN   0-521-64298-1.
  2. Moon, Todd K. (2005). Error Correction Coding, Mathematical Methods and Algorithms. Wiley. ISBN   0-471-64800-0. (Includes code)
  3. Amin Shokrollahi, LDPC Codes: An Introduction (PDF), archived from the original (PDF) on May 17, 2017
  4. Hardesty, L. (January 21, 2010). "Explained: Gallager codes". MIT News. Retrieved August 7, 2013.
  5. 1 2 Gallager, R.G. (January 1962). "Low density parity check codes". IRE Trans. Inf. Theory. 8 (1): 21–28. doi:10.1109/TIT.1962.1057683. hdl:1721.1/11804/32786367-MIT. S2CID   260490814.
  6. 1 2 Erico Guizzo (March 1, 2004). "CLOSING IN ON THE PERFECT CODE". IEEE Spectrum. "Another advantage, perhaps the biggest of all, is that the LDPC patents have expired, so companies can use them without having to pay for intellectual-property rights."
  7. US 5446747
  8. Mackenzie, D. (July 9, 2005). "Communication speed nears terminal velocity". New Scientist.
  9. Mosheiff, J.; Resch, N.; Ron-Zewi, N.; Silas, S.; Wootters, M. (2020). "Low-Density Parity-Check Codes Achieve List-Decoding Capacity". SIAM Journal on Computing (FOCS 2020): 38–73. doi:10.1137/20M1365934. S2CID   244549036.
  10. Gallager, Robert G. (1963). Low Density Parity Check Codes (PDF). M.I.T. Press. Retrieved August 7, 2013.
  11. 1 2 David J.C. MacKay and Radford M. Neal, "Near Shannon Limit Performance of Low Density Parity Check Codes," Electronics Letters, July 1996
  12. Telemetry Data Decoding, Design Handbook
  13. Presentation by Hughes Systems Archived 2006-10-08 at the Wayback Machine
  14. HomePNA Blog: G.hn, a PHY For All Seasons
  15. IEEE Communications Magazine paper on G.hn Archived 2009-12-13 at the Wayback Machine
  16. IEEE Standard, section 20.3.11.6 "802.11n-2009", IEEE, October 29, 2009, accessed March 21, 2011.
  17. "IEEE SA - IEEE 802.11ax-2021". IEEE Standards Association. Retrieved May 22, 2022.
  18. Chih-Yuan Yang, Mong-Kai Ku. http://123seminarsonly.com/Seminar-Reports/029/26540350-Ldpc-Coded-Ofdm-Modulation.pdf "LDPC coded OFDM modulation for high spectral efficiency transmission"
  19. Nick Wells. "DVB-T2 in relation to the DVB-x2 Family of Standards" Archived 2013-05-26 at the Wayback Machine
  20. "5G Channel Coding" (PDF). Archived from the original (PDF) on December 6, 2018. Retrieved January 6, 2019.
  21. Maunder, Robert (September 2016). "A Vision for 5G Channel Coding" (PDF). Archived from the original (PDF) on December 6, 2018. Retrieved January 6, 2019.
  22. Kai Zhao; Wenzhe Zhao; Hongbin Sun; Tong Zhang; Xiaodong Zhang; Nanning Zheng (2013). LDPC-in-SSD: Making Advanced Error Correction Codes Work Effectively in Solid State Drives (PDF). FAST' 13. pp. 243–256.
  23. "Soft-Decoding in LDPC based SSD Controllers". EE Times. 2015.
  24. Robert McEliece, E. R. Berlekamp and H. Van Tilborg (1978). "On the Inherent Intractability of Certain Coding Problems". IEEE Trans. Inf. Theory. IEEE: 384–386. doi:10.1109/TIT.1978.1055873.
  25. 1 2 Casado, A.I.V.; Griot, M.; Wesel, R.D. (2007). Informed Dynamic Scheduling for Belief-Propagation Decoding of LDPC Codes. 2007 IEEE International Conference on Communications, Glasgow, UK. pp. 932–7. arXiv: cs/0702111 . doi:10.1109/ICC.2007.158.
  26. Richardson, T. (October 2003). "Error floors of LDPC codes" (PDF). Proceedings of the Annual Allerton Conference on Communication Control and Computing. 41 (3): 1426–35. ISSN   0732-6181.
  27. Richardson, T.J.; Shokrollahi, M.A.; Urbanke, R.L. (February 2001). "Design of capacity-approaching irregular low-density parity-check codes". IEEE Transactions on Information Theory. 47 (2): 619–637. doi:10.1109/18.910578.
  28. Richardson, T.J.; Urbanke, R.L. (February 2001). "Efficient encoding of low-density parity-check codes". IEEE Transactions on Information Theory. 47 (2): 638–656. doi:10.1109/18.910579.
  29. Ahmad Darabiha, Anthony Chan Carusone, Frank R. Kschischang. "Power Reduction Techniques for LDPC Decoders"
  30. Zhang, Z.; Anantharam, V.; Wainwright, M.J.; Nikolic, B. (April 2010). "An Efficient 10GBASE-T Ethernet LDPC Decoder Design With Low Error Floors" (PDF). IEEE Journal of Solid-State Circuits. 45 (4): 843–855. Bibcode:2010IJSSC..45..843Z. doi:10.1109/JSSC.2010.2042255. S2CID   10431486.
  31. Kou, Y.; Lin, S.; Fossorier, M.P.C. (November 2001). "Low-density parity-check codes based on finite geometries: a rediscovery and new results". IEEE Transactions on Information Theory. 47 (7): 2711–36. CiteSeerX   10.1.1.100.3023 . doi:10.1109/18.959255.
  32. Tahir, B.; Schwarz, S.; Rupp, M. (2017). BER comparison between Convolutional, Turbo, LDPC, and Polar codes. 2017 24th International Conference on Telecommunications (ICT), Limassol, Cyprus. pp. 1–7. doi:10.1109/ICT.2017.7998249.
  33. Moon Todd, K. (2005). Error correction coding: mathematical methods and algorithms. Wiley. p. 614. ISBN   0-471-64800-0.
  34. Moon Todd 2005 , p. 653
  35. Andrews, Kenneth S., et al. "The development of turbo and LDPC codes for deep-space applications." Proceedings of the IEEE 95.11 (2007): 2142-2156.
  36. Hassan, A.E.S., Dessouky, M., Abou Elazm, A. and Shokair, M., 2012. Evaluation of complexity versus performance for turbo code and LDPC under different code rates. Proc. SPACOMM, pp.93-98.
  37. "IEEE Spectrum: Does China Have the Best Digital Television Standard on the Planet?". spectrum.ieee.org. Archived from the original on December 12, 2009.