91
Multi-Edge Low-Density Parity-Check Coded Modulation by Lei Zhang A thesis submitted in conformity with the requirements for the degree of Master of Applied Science Graduate Department of Edward Rogers Sr. Department of Electrical and Computer Engineering University of Toronto Copyright c 2011 by Lei Zhang

Multi-Edge Low-Density Parity-Check Coded Modulation · 2013-10-24 · Abstract Multi-Edge Low-Density Parity-Check Coded Modulation Lei Zhang Master of Applied Science Graduate Department

  • Upload
    others

  • View
    27

  • Download
    0

Embed Size (px)

Citation preview

Multi-Edge Low-Density Parity-Check Coded Modulation

by

Lei Zhang

A thesis submitted in conformity with the requirementsfor the degree of Master of Applied Science

Graduate Department of Edward Rogers Sr. Department of Electricaland Computer Engineering

University of Toronto

Copyright c© 2011 by Lei Zhang

Abstract

Multi-Edge Low-Density Parity-Check Coded Modulation

Lei Zhang

Master of Applied Science

Graduate Department of Edward Rogers Sr. Department of Electrical and Computer

Engineering

University of Toronto

2011

A method for designing low-density parity-check (LDPC) codes for bandwidth-efficient

high-order coded modulation is proposed. Code structure utilizes the multi-edge-type

LDPC code ensemble to achieve an improved match between codeword bit protection ca-

pabilities and modulation bit-channel capacities over existing LDPC coded modulation

techniques. The multi-dimensional EXIT vector field for the specific multi-edge parame-

terization is developed for the analysis and design of code ensembles. A multi-dimensional

EXIT decoding convergence condition is derived to enable efficient optimization. Code

design results in terms of ensemble thresholds and finite-length Monte-Carlo simulations

indicate that the new technique improves on the state-of-the-art performance, with sig-

nificantly lower design and implementation complexity.

ii

Acknowledgements

This thesis would not have been written but for the great number of people from

whom I have received sage advice, unparalleled friendship and unconditional love. At the

forefront of this amazing group is my advisor, professor Frank Kschischang, whose over-

arching perspective, originality of ideas, seemingly boundless knowledge and meticulous

attention to detail have ensured a smooth and thoroughly enjoyable research experience

for me. In the last few years I’ve experienced tremendous academic and professional

growth under his guidance, inspired by his dedication and excellence in all aspects of an

academic career, from research, teaching, to professional commitments. It has been an

honour and a pleasure to work with professor Kschischang.

A special thank you to Dr. Benjamin Smith, whose wealth of knowledge, insights

and advice have greatly facilitated my research. I’ve also enjoyed the many stimulating

conversations we’ve had regarding careers and the minutiae of academia. It is one of

the many aspects which allowed him to become a valuable role model to an incipient

researcher such as myself.

To all my friends, I honestly believe that without your encouragement, company, and

coffee breaks, I would not have been able to overcome several particularly trying stages

throughout this endeavour. Thank you all. I hope to have the opportunity to repay each

and every one of you in kind.

Finally, to my parents, I dedicate this thesis to you as a small token of my appreciation

for your unconditional love and support. I love you both from the bottom of my heart.

iii

Contents

1 Introduction 1

1.1 Improving BICM-LDPC . . . . . . . . . . . . . . . . . . . . . . . . . . . 5

1.2 Literature Review . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8

1.3 Thesis Outline . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10

2 Technical Background 11

2.1 Target System . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11

2.1.1 IID channel adapter . . . . . . . . . . . . . . . . . . . . . . . . . 13

2.1.2 Non-iterative vs. Iterative demapping . . . . . . . . . . . . . . . . 14

2.1.3 Shaping . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16

2.2 Analysis and Design of Binary LDPC Codes . . . . . . . . . . . . . . . . 16

2.2.1 Ensemble-based design and density evolution . . . . . . . . . . . . 16

2.2.2 Extrinsic information transfer charts . . . . . . . . . . . . . . . . 23

2.3 Multi-edge-type LDPC Codes . . . . . . . . . . . . . . . . . . . . . . . . 26

3 Multi-edge LDPC Coded Modulation 31

3.1 Multi-edge Parameterization . . . . . . . . . . . . . . . . . . . . . . . . . 31

3.1.1 Check degree edge-type assignment . . . . . . . . . . . . . . . . . 35

3.2 Multi-edge Optimization . . . . . . . . . . . . . . . . . . . . . . . . . . . 38

3.2.1 Multi-dimensional EXIT vector field . . . . . . . . . . . . . . . . 39

3.2.2 Design of multi-edge coded modulation . . . . . . . . . . . . . . . 47

iv

4 Results 61

4.1 Threshold . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 61

4.1.1 Discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 62

4.1.2 High code rate designs . . . . . . . . . . . . . . . . . . . . . . . . 65

4.2 Finite-length . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 65

4.2.1 Rate 3/4 Gray-labelled 16-QAM . . . . . . . . . . . . . . . . . . . 68

4.2.2 Rate 1/2 Gray-labelled 16-QAM . . . . . . . . . . . . . . . . . . . 72

4.2.3 Fixed BER performance comparison . . . . . . . . . . . . . . . . 72

5 Conclusion 76

Bibliography 78

v

List of Tables

3.1 Density evolution thresholds for the (3,6) regular LDPC check degree edge-

type split pairings under Gray-labelled 16-QAM. . . . . . . . . . . . . . . 36

3.2 Density evolution thresholds for the (3,9) regular LDPC check degree edge-

type split pairings under Gray-labelled 16-QAM. . . . . . . . . . . . . . . 37

3.3 Density evolution thresholds for irregular LDPC check degree edge-type

split pairings under Gray-labelled 16-QAM. . . . . . . . . . . . . . . . . 38

3.4 Bit-channel capacities for Gray-labelled 2n-QAM at rate (n− 1)/n . . . . 59

4.1 ME-LCM-OPT optimized ensembles for 2n-QAM rate (n− 1)/n codes . 63

4.2 ME-LCM-OPT optimized ensembles of various code rates for 16-QAM . 66

4.3 ME-LCM-OPT optimized ensembles of various code rates for 16-QAM

(continued) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 67

4.4 Legend for data points in Fig. 4.6 . . . . . . . . . . . . . . . . . . . . . . 75

vi

List of Figures

1.1 Decoding block diagrams of BICM and MLC-MSD with 4 bit-levels. . . . 4

1.2 Tanner graph for parity check matrix in Eqn. 1.3. . . . . . . . . . . . . . 5

1.3 Gray-labelled 16-QAM constellation with labels corresponding to bits b0b1b2b3. 7

1.4 Histogram of LLRs for Gray-mapped 16-QAM bit-levels b0, b1, b2, b3. . . 7

2.1 Target system block diagram for the baseband-equivalent discrete-time

complex AWGN channel. . . . . . . . . . . . . . . . . . . . . . . . . . . . 12

2.2 Bit-channel and symbol-channel capacities of 16-QAM for set-partition

labelled MLC/MSD and Gray-labelled BICM. . . . . . . . . . . . . . . . 15

2.3 EXIT chart of optimized rate = 0.33 ensemble at threshold of -1.91 dB

Es/N0. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 25

2.4 A possible multi-edge-type representation of the Tanner graph in Fig. 1.2 27

3.1 Tanner graph of the 2 edge-type specified MET parameterization, node

degrees are illustrative and not meant to be realistic. . . . . . . . . . . . 33

3.2 Multi-dimensional EXIT vector field for 2 edge-types at threshold σ∗ =

0.3665. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 44

3.3 Multi-dimensional EXIT vector field for 2 edge-types at above threshold

σ = 0.3666. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 46

3.4 Backward-difference vector field of 2 edge-types at threshold σ∗ = 0.3665. 51

vii

3.5 Backward-difference vector field of 2 edge-types at above threshold σ =

0.3666. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 55

4.1 ME-LCM-OPT optimized ensembles for 2n-QAM rate (n− 1)/n codes. . 64

4.2 Probability of bit and errors for n = 4096 rate 3/4 code and Gray-labelled

16-QAM. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 69

4.3 Probability of bit and errors for n = 8192 rate 3/4 code and Gray-labelled

16-QAM. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 70

4.4 Probability of bit and errors for n = 16384 rate 3/4 code and Gray-labelled

16-QAM. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 71

4.5 Probability of bit and errors for n = 16200 rate 1/2 code and Gray-labelled

16-QAM. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 73

4.6 Comparison of Es/N0 required to achieve BER of 10−5 at different spectral

efficiencies for Gray-labelled 16-QAM. . . . . . . . . . . . . . . . . . . . . 74

viii

Chapter 1

Introduction

In many communication systems, the modulation and channel code are designed sepa-

rately. Several factors motivate this paradigm. A complicated modulation system can

often be encapsulated by a simple channel model, such as the binary symmetric channel

(BSC), to greatly simplify the channel code design without incurring a significant loss in

performance. For applications with high error tolerance, the uncoded modulation error

rate may be sufficiently low to obviate the need for a channel code. Even the pedagogical

tradition in undergraduate and graduate digital communication courses dictates sepa-

rating modulation and coding. The most dominant reason, however, is the complexity

of designing the modulation and channel code in conjunction. The trade-off between

the cost to design the more complex system and the performance gains from doing so

often results in the pragmatic engineering solution of the distinct modulation-coding

architecture.

For applications where bandwidth is a limiting resource, the gain in combining mod-

ulation and channel coding significantly outweighs the increase in system complexity. A

well-known example of coded modulation is Trellis Coded Modulation (TCM) [1]. TCM

combines the Euclidean distance properties of the modulation signal constellation with

the Hamming distance properties of the error-correcting convolutional (trellis) code in its

1

Chapter 1. Introduction 2

design process. Design rules map trellis transitions of the convolutional code to subsets

of the constellation of different Euclidean distances. In the early 1980’s, telephone line

modem designers considered 9600 kbit/s to be the limiting throughput under standard

bandwidth and power constraints. With the introduction of TCM in 1984, telephone

modems achieved 14.4 kbit/s and higher [2], which helped usher in the meteoric rise of

personal dial-up Internet services.

After the discovery of capacity-approaching turbo and low-density parity-check (LDPC)

codes in the early 90’s [3,4], coded modulation using these modern codes evolved follow-

ing two dominant approaches based on modulation bit-level capacities. Bit-Interleaved

Coded-Modulation (BICM) [5] uses an interleaver between the channel code encoder and

the bit-to-symbol mapper. BICM averages the Euclidean distances of different modula-

tion bit-levels so that the underlying error-correcting code experiences an average noise

degradation from the channel. The BICM bit-channel capacities, I(bi;Y ), can be shown

to be an approximation of the expanded symbol channel mutual information using the

“chain rule”

I(X;Y ) = I(b0, . . . , bn−1;Y )

=n−1∑i=0

I(bi;Y |b0, . . . , bi−1) (1.1)

≥n−1∑i=0

I(bi;Y ). (1.2)

Even though capacity is lost in the process, it has been shown in [5] that BICM

with Gray labelling can closely approximate the finite-constellation constrained channel

capacity. A capacity-approaching code can then be designed for each bit-channel capacity.

An alternative method of coded modulation is Multi-Level Coding (MLC) [6, 7]. In

MLC, for a fixed constellation label, the bit-level capacities are given by (1.1) and achieves

the symbol channel capacity. Again, capacity-approaching codes can be designed for each

bit-channel. At the receiver, decoding progresses in a level-by-level fashion called Multi-

Chapter 1. Introduction 3

Stage Decoding (MSD) [7]. Initially only the decoder for bit-level b0 is active. Assuming

all b0 bits are correctly decoded, the receiver uses this side-information along with received

information to decode b1, and so on. A block diagram comparison between BICM and

MLC-MSD coded modulation is shown in Fig. 1.1. Even though MLC/MSD can achieve

channel capacity, it has significant disadvantages compared to BICM. Decoding latency of

MSD is a major issue for low latency systems. Error propagation from lower to higher bit-

levels may increase error rates. Furthermore, passing lower bit-level soft information to

higher bit-levels requires iterating through the symbol demapper which increases receiver

complexity.

Taking into account the factors of power efficiency, decoding latency, and implementa-

tion complexity, BICM outclasses MLC/MSD as the more suitable capacity-approaching

coded modulation technique. On-going standardization activity lends supports to this

conclusion. LDPC-based BICM (LDPC-BICM) is included in the 2nd-generation satel-

lite television standard (DVB-S2) [8] and the 2nd-generation digital cable television

standard (DVB-C2) [9]. LDPC-BICM is defined in the multiple access mode of the

wireless metropolitan area networks standard (WiMax) [10] as a high performance op-

tion. Protograph-based LDPC-BICM is proposed for deep space communication in the

Consultive Committee for Space Data Systems (CCSDS) 1.3.1-O-2 standard [11].

The complexity of LDPC-BICM is minimal since bit-interleaving can be built into

the LDPC parity-check matrix. However, in each of these standards except CCSDS, the

LDPC code used is based on an Irregular Repeat-Accumulate (IRA) code [12] initially de-

signed for a low-rate power-limited applications. There is currently no bandwidth-limited

application that uses high-rate LDPC codes specifically designed for coded modulation.

It appears that combining an available LDPC design with BICM has become the de facto

standard.

However, it may be imprudent to accept LDPC-BICM as the optimal LDPC coded

modulation technique. In the next section we provide an argument for developing true

Chapter 1. Introduction 4

Y b0 decoder

b1 decoder

b2 decoder

b3 decoder

b0

b1

b2

b3

(a) BICM

Y b0 decoder

b1 decoder

b2 decoder

b3 decoder

b0

b1

b2

b3

(b) MLC-MSD

Figure 1.1: Decoding block diagrams of BICM and MLC-MSD with 4 bit-levels.

Chapter 1. Introduction 5

LDPC coded modulation that accounts for modulation bit-level differences in LDPC

code design. Significant performance gains maybe achievable with such an improvement

to LDPC-BICM.

1.1 Improving BICM-LDPC

To introduce the argument for an improved BICM-LDPC design technique, we introduce

a few necessary details of LDPC codes. To keep the treatment brief, we relegate additional

technical details to Ch. 2.

LDPC codes are linear codes with sparse parity-check matrices. Each column of the

LDPC parity-check matrix represents a codeword bit. The non-zero entries in a column

denote the parity-check equations to which the particular codeword bit belongs. The

non-zero entries in a row of the parity-check matrix denote the codeword bits checked

by that parity-check equation. The parity-check matrix can be visually represented by

a bipartite graph called Tanner graph [13]. An example is shown in Fig. 1.2 for the

following parity-check matrix

H =

1 0 0 1 1 0 1

0 1 0 1 0 1 1

0 0 1 0 1 1 1

. (1.3)

The circles in Fig. 1.2 are called variable nodes and represent the columns of H. The

squares are called check nodes and represent the rows of H. The edges of the Tanner

+ + +

Figure 1.2: Tanner graph for parity check matrix in Eqn. 1.3.

Chapter 1. Introduction 6

graph connect codeword bits to the parity-check equations to which they belong.

The total number of edges that each variable/check node possesses is called the de-

gree of the variable/check node, corresponding to the number of non-zero entries in a

column/row of the parity-check matrix. If all variable nodes have the same degree, the

LDPC code is called regular. LDPC codes with variable nodes of different degrees are

called irregular codes. All capacity-approaching LDPC codes are irregular [14,15].

Decoding of LDPC codes uses the sum-product algorithm [16]. At the start, each vari-

able node receives a reliability measure from the demapper and sends it to neighbouring

checks. At the check nodes, the reliabilities are updated according to how well they sat-

isfy the parity-check constraints. After a complete iteration, variable nodes re-evaluate

their reliabilities according to repetition code constraints. This “message passing” action

continues until the variable node reliabilities are sufficiently high for a hard decision or

a maximum number of iterations has been reached.

We now outline the argument for seeking to improve the design of LDPC codes for

BICM systems. As repetition codes, LDPC variable nodes of different degrees offer dif-

ferent levels of error correction capability. High-degree variable nodes behave as very

long repetition codes and therefore are extremely reliable. However their low rates de-

crease the overall code rate significantly. On the other hand, degree 2 variable nodes

offer essentially no error correction but have the highest rate of all repetition codes. In

a capacity-approaching irregular LDPC code, there exists an inherent variation among

the different codeword bits.

Interestingly, high-order modulation also produces differences in the reliabilities of

received codeword bits. Distinct bit-levels in the symbol-labelling experience different

amounts of noise corruption due to Euclidean distance differences. For the constellation

and labelling given in Fig. 1.3, we plot the histogram of bit-channel output log-likelihood

ratios (LLR) in Fig. 1.4. The system signal-to-noise ratio is 9.32 dB Es/N0.

From the empirical means of the LLR distributions (black markers), we may conclude

Chapter 1. Introduction 7

−1 −0.5 0 0.5 1

−1

−0.5

0

0.5

1

In−Phase

Qua

drat

ure

0000

0001

0011

0010

0100

0101

0111

0110

1100

1101

1111

1110

1000

1001

1011

1010

Figure 1.3: Gray-labelled 16-QAM constellation with labels corresponding to bits

b0b1b2b3.

−40 −35 −30 −25 −20 −15 −10 −5 0 5 10 15 20 25 30 35 400

0.002

0.004

0.006

0.008

0.01

0.012

0.014

0.016

0.018

0.02

Log−likelihood ratio

Nor

mal

ized

occ

uran

ce

b1,b3

b0,b2

Figure 1.4: Histogram of LLRs for Gray-mapped 16-QAM bit-levels b0, b1, b2, b3.

Chapter 1. Introduction 8

that bit-levels b0 and b2 are of higher quality than the remaining two bits. Other measures

of bit-channel quality such as probability of error and mutual information also point to

this conclusion. Therefore, high-order modulation inherently produces different qualities

of bit-levels hence varying reliabilities over channel codeword bits.

BICM averages over these different bit-channel output reliabilities. For codes which

do not have varying levels of protection over codeword bits, this is a good coded modu-

lation technique. However, irregular LDPC codes implement varying levels of protection

over the codeword bits. We believe that the performance of high-order LDPC-BICM

coded modulation can be improved if the differences in bit-level output reliabilities are

integrated into the LDPC code design procedure. Much like the classical TCM scheme,

LDPC coded modulation must exploit bit-level reliability differences by optimizing the

variable node degree distributions and mapping assignments simultaneously.

1.2 Literature Review

While the design of capacity-approaching LDPC codes for binary-input memoryless sym-

metric channels such as the BSC and the additive white Gaussian channel (BIAWGNC)

has been thoroughly studied [14, 17, 18], much less is known about the design of LDPC

codes for high-order coded modulation. The few available references on improving LDPC

coded modulation focus on two general methodologies: the bit-to-symbol interleaving of

a fixed LDPC code [19–21], and code design incorporating differences in bit-level relia-

bilities [22–26]. The first method is nothing more than finding a particular interleaver

for BICM without code design. Only the second method can be truly considered to be

the design of LDPC codes for coded modulation.

In search of an improved interleaver, the authors of [19] proposed a mapping scheme

where less-protected variable nodes were mapped to low-reliability bit-levels while more-

protected variable nodes were mapped to highly reliable bit-levels. The mapping provided

Chapter 1. Introduction 9

0.15-0.20 dB of improvement at no complexity increase. Intuitively, the improvement may

have been the result of allowing the most-reliable messages to propagate widely from high-

degree variable nodes. In [20], a mapping was proposed to minimize the connections of

each check node to variable nodes with low-reliability channel output, resulting in 0.3-0.7

dB of improvement. Finally, [21] proposed an improved interleaver for the DVB-S2 LDPC

code after taking into account the bit-level reliability differences. These mapping-based

methods certainly improved LDPC-BICM performance, but since the underlying LDPC

code was fixed the improvement was limited.

The most significant work on the design of LDPC codes for coded modulation has

been [22]. The work applies density evolution [14] to design LDPC codes for the dis-

tinct bit-channels of MLC and BICM. The problem is reduced to several binary LDPC

code designs. In [23] a powerful class of low-complexity, low error-floor LDPC codes

based on protographs are applied to high-order modulation with impressive performance.

Together, the references [22, 23] provide the most significant references for our work.

In [27, 28] LDPC codes are only used for low to medium quality bit channels, while

very high quality bit-channels are either uncoded or protected by very simple classical

codes. In [24–26], the multi-edge-type concepts are used in LDPC coded modulation.

Multi-edge-type LDPC codes [29, pp. 382-397] can incorporate the distinct bit-channel

reliabilities into code optimization. They also give the designer flexible control over code

structure to trade-off between complexity and performance. Although [24–26] did allude

to multi-edge ideas, they fall short of providing a specific multi-edge parameterization

with efficient analysis and design techniques. The key contribution of this thesis is the

design of multi-edge-type LDPC codes for LDPC-BICM.

Chapter 1. Introduction 10

1.3 Thesis Outline

Ch. 2 provides all the necessary technical background used in the rest of this thesis.

The target system model, fundamentals of LDPC design, density evolution and extrinsic

information transfer (EXIT) charts are a few of the topics reviewed in the chapter. Ch.

3 develops the multi-edge LDPC code design technique for coded modulation and forms

the main body of the thesis. The development follows from the initial specialization

of the multi-edge parameterization, to the multi-dimensional EXIT chart method for

analyzing such ensembles, to the innovative technique for code design based on the multi-

dimensional EXIT chart. Results of the new LDPC coded modulation design technique

are given in Ch. 4 in terms of both the ensemble thresholds and finite-length simulations.

Ch. 5 concludes the thesis and provides directions for future work.

Chapter 2

Technical Background

The goal of this chapter is to provide a comprehensive review of the technical knowledge

required for understanding the rest of the thesis. Sec. 2.1 describes the target system.

Several alternatives are discussed and justifications are given for choosing to limit the

scope of the thesis to one system. Binary LDPC analysis and design techniques are

reviewed in Sec. 2.2. A thorough understanding of the details and intuition of these

techniques is essential since the solutions developed in this thesis are based on these

binary design techniques. Lastly, Sec. 2.3 provides details on multi-edge-type LDPC

codes.

Throughout this thesis, bold font always denotes a vector quantity of length clear

from context. A bold constant denotes a vector of repeated entries, all of which are

equal to the indicated value. For example, 1 = (1, . . . , 1). A bold variable either denotes

a vector of variables, for example x = (x1, . . . , xn) or a vector field function f(x) =

(f1(x), . . . , fn(x)). The difference between them will be clear from context.

2.1 Target System

We now describe the system at which the design techniques given in this thesis are

aimed. As illustrated in Fig. 2.1, the source of the target system generates a sequence of

11

Chapter 2. Technical Background 12

ENC MAP

n

DMAP DECm c x y L c

Figure 2.1: Target system block diagram for the baseband-equivalent discrete-time com-

plex AWGN channel.

uniformly distributed independent and identically distributed binary random variables.

The channel code encoder takes k source bits per input block and maps them to a length

n (where n ≥ k) codeword c from the channel codebook C. The code rate is r = k/n.

A mapping function µ maps each codeword c to a sequence of modulation symbols

x, where each symbol xi is selected from the constellation X . In M bits per symbol

modulation, there are 2M distinct points in the constellation. Assuming n is a multiple

of M , the sequence x contains n/M symbols. For an arbitrary constellation labelling

scheme, let m = (0, . . . ,M − 1) index the bit-levels of the labelled constellation and let

bmi denote the m-th bit in the label of the transmitted symbol xi.

The sequence of symbols x, in discrete-time baseband-equivalent representation of

bandpass transmission, is corrupted by complex additive white Gaussian noise n = nI +

jnQ of variance σ2 = N0/2 per dimension, where N0/2 is the two-sided power spectral

density of the Gaussian noise. Note that all simulations in this thesis are performed

by assuming unity average symbol energy while scaling noise variance to the desired

signal-to-noise ratio.

At the receiver, the received sequence y = x + n is demapped by taking the bit-wise

log-likelihood ratio (LLR)

LMi+m = logP (yi|bmi = 0)

P (yi|bmi = 1)(2.1)

where i = (0, . . . , n/M − 1) indexes the received sequence y.

The uncoded maximum-likelihood decision rule is to decide 0 if LMi+m > 0, decide 1

if LMi+m < 0 and decide 0 or 1 at random with equal probability if LMi+m = 0.

Chapter 2. Technical Background 13

For complex additive white Gaussian channel with N0 = σ2 the LLR is given by

LMi+m = log

∑a∈X 0

m

exp

(− 1

N0

‖yi − a‖2)

∑b∈X 1

m

exp

(− 1

N0

‖yi − b‖2) . (2.2)

where X 0m and X 1

m partition the signalling constellation X into sets of points where bm = 0

and bm = 1, respectively.

The length n sequence L is decoded by the channel code decoder to the decoded word

c. A bit error occurs if ci 6= ci for some i, a frame error occurs if one or more bit errors

occur in the decoded word. After correct decoding, the transmitted message m can be

extracted from c if C is systematic. In this thesis, we judge the system error performance

only by comparing c to c.

2.1.1 IID channel adapter

As will be explained in Sec. 2.2, it is highly desirable for the bit-wise channels from

the transmitted codeword bits ci to the demapped LLR bit reliabilities Li to be output

symmetric. By definition, a binary-input channel is output symmetric if

P (Li|ci = 0) = P (−Li|ci = 1). (2.3)

High-order modulation systems are in general not output symmetric. A work-around

to this difficulty was introduced in [22] by inserting independent and identically dis-

tributed (iid) channel adapters into the system. At the transmitter, the iid channel

adapter XORs codeword c with a random binary sequence u generated from identically

distributed, uniform Bernoulli random variables. At the receiver, the sequence L is multi-

plied bit-wise by 1−2u. It is easy to verify that these two operations produce bit-channel

output symmetry while maintaining the same bit-channel capacity as the original sys-

tem. For proof please see the reference [22]. Note that iid channel adapters can be easily

Chapter 2. Technical Background 14

implemented in practice using synchronized pseudo-random binary sequence generators

at the transmitter and receiver.

2.1.2 Non-iterative vs. Iterative demapping

The system in Fig. 2.1 performs the single demap and decode operation used in BICM

systems. In MLC/MSD, the decoders for the lower bit-levels pass decoded bit informa-

tions back to the demapper to assist the next bit-level decoder. As mentioned, MSD can

achieve a capacity higher than non-iterative demapping. However, it has been shown

in [5] that if binary-reflected Gray labelling (BRGL) is used to label the constellation,

then the difference between the iterative and non-iterative demapping schemes is ex-

tremely small at high code rates. Fig. 2.2 plots the bit-channel and symbol-channel

capacities for 16-QAM under iterative and non-iterative decoding. The iterative scheme

is labelled using set-partition while non-iterative scheme uses BRGL. Note the small

difference (0.0037 dB) between MLC/MSD and BICM at rate 3/4.

The three types of capacities shown in Fig. 2.2 should be carefully distinguished.

The ultimate Shannon limit (red curve) is the capacity achieved under a continuous,

capacity-achieving input distribution with iterative demapping. The iterative demap-

ping capacity (dashed blue curve) can be achieved under discrete, uniformly distributed

16-QAM constellation with iterative demapping. The non-iterative demapping capacity

(solid blue curve) can be achieved under discrete, uniformly distributed 16-QAM constel-

lation without iterative demapping. We will always refer to the non-iterative demapping

capacity in this thesis unless otherwise noted.

Iterative demapping requires higher receiver complexity and latency. In light of the

negligible loss in capacity at the operating point marked in Fig. 2.2, we are justified to

focus on the low-complexity non-iterative Gray-labelled BICM scheme.

Chapter 2. Technical Background 15

−10 −5 0 5 10 15 200

0.5

1

1.5

2

2.5

3

3.5

4

4.5

5

Es/N

0 (dB)

bit/c

hann

el u

se

b

0 GY

b1 GY

b2 GY

b3 GY

b0 SP

b1 SP

b2 SP

b3 SP

sym GYsym SPC

Sh

1.53 dB shaping gain

∆ 0.0037 dB

Figure 2.2: Bit-channel and symbol-channel capacities of 16-QAM for set-partition la-

belled MLC/MSD and Gray-labelled BICM.

Chapter 2. Technical Background 16

2.1.3 Shaping

The gap between the ultimate Shannon limit and MLC/MSD capacity is due to the use

of a discrete, uniformly-distributed M -QAM constellation for the continuous-input com-

plex AWGN channel. From information theory we know the capacity-achieving input

distribution for this channel is a two-dimensional circularly symmetric Gaussian distri-

bution. The technique of approximating this ideal input distribution using a discrete

distribution is called shaping [30, 31]. An asymptotic shaping gain of 1.53 dB can be

achieved as shown in Fig. 2.2. We neglect shaping in our system to limit the scope of our

research. This simplification has also been made in all of prior works cited in Sec. 1.2.

We believe that shaping techniques can be applied to our code designs without affecting

the coded modulation gains.

2.2 Analysis and Design of Binary LDPC Codes

In this section we review well-known design and analysis techniques for capacity-approaching

irregular LDPC codes. Asymptotic ensemble analysis using density evolution and its ap-

proximation using extrinsic information transfer (EXIT) charts is explained in detail.

2.2.1 Ensemble-based design and density evolution

An irregular LDPC code of length n is fully specified by the number of variable nodes

and their degrees, the number of check nodes and their degrees, and the edge connections

between variable and check nodes. The number of variable and check nodes and their

degrees can be conveniently represented by using degree distribution polynomials

Λ(x) =dv∑i=1

Λixi , P (x) =

dc∑i=2

Pixi (2.4)

where dv and dc are maximum variable and check node degrees, Λi and Pi are the number

of variable and check nodes of each degree. Note that check degrees are greater or equal

Chapter 2. Technical Background 17

to 2 since a parity check equation of 1 term is useless. The total number of variable

nodes is Λ(1) = n and of check nodes is P (1) = (1− r)n.

It is more useful to normalize (2.4) by the number of variable and check nodes. We

define the node-perspective normalized degree distribution polynomials

L(x) =Λ(x)

Λ(1), R(x) =

P (x)

P (1). (2.5)

Node-perspective indicates that the coefficient in each term of the degree polynomial

denotes the fraction of nodes of that degree. An alternative edge-perspective degree

distribution polynomial indicates that the coefficients refer to the fraction of edges (in

total number edges) that are connected to nodes of that degree. It is easy to convert

node-perspective to edge-perspective degree distribution polynomials by

λ(x) =L′(x)

L′(1), ρ(x) =

R′(x)

R′(1)(2.6)

where ′ denotes differentiation.

Converting back to node-perspective degree distribution polynomials is achieved by

L(x) =

∫ x0λ(s)ds∫ 1

0λ(s)ds

, R(x) =

∫ x0ρ(s)ds∫ 1

0ρ(s)ds

. (2.7)

An ensemble is the set of all parity-check matrices (equivalently Tanner graphs) that

satisfy the degree distribution polynomials. By “satisfy”, we mean satisfy to within some

small tolerance, since in finite block-length it is often impossible to exactly satisfy the

distribution polynomials. Consider the degree of a node to be the number of “sockets”

it has available for edges to plug into. An edge-count constraint L′(1) = R′(1) is placed

on the variable and check degree distribution polynomials to ensure an equal number

of sockets on both sides. Let π be a permutation for a particular connection of edges

between variable node sockets and check node sockets. An ensemble is defined to be the

collection of all Tanner graphs which satisfy the degree distribution polynomials, under

all possible permutations π, and all possible channel outputs [18].

Chapter 2. Technical Background 18

The channel outputs are assumed to be independent between all codeword bits. The

bit-channel is assumed to be output symmetric as defined in (2.3). For a symmetric

channel, one can show that assuming only the all-zeros codeword is sent is equivalent

in performance to assuming all possible codewords are sent [29, pp. 215-216]. Hence

under the symmetric channel assumption, the ensemble also encompasses all possible

transmitted codewords. This is why iid channel adapters are necessary in our target

system. Ensemble-based analysis cannot be used for code design if the bit-channels are

not output symmetric.

The sum-product algorithm

Ensemble-based analysis evaluates the expected probability of message errors for some

decoding algorithm, averaged over all Tanner graphs and channel outputs in the ensem-

ble. The decoding algorithm for LDPC codes in AWGN is a specialized instance of the

message-passing sum-product algorithm [16] called Belief Propagation (BP) [4].

For any variable node vi in a LDPC code Tanner graph, denote its channel output

LLR by µi. Let J denote the indices of its neighbouring check nodes. For any check

node cj, let I denote the indices of its neighbouring variable nodes. Let µvi→j and µci→j

represent the messages passed from variable and check node of index i to a node of index

j. Initially, all µci→j = 0 and all µvi→j = µi.

In subsequent iterations, the variable node update equation is

µvi→j =∑j′∈J\j

µcj′→i + µi, ∀j ∈ J . (2.8)

The check node update equation is

µcj→i = 2 tanh−1

∏i′∈I\i

tanh

(µvi′→j

2

) , ∀i ∈ I. (2.9)

A full decoding iteration includes one execution of variable and check updates. At

Chapter 2. Technical Background 19

the end of every iteration, a hard decision is made for variable vi based on µvi→j + µcj→i

using the decision rule from Sec. 2.1. Decoding ends when the hard-decision codeword

passes parity check or a maximum number of decoding iterations has been reached.

Concentration, decoding-tree and density evolution

Several key results justify ensemble-based analysis and its main tool, density evolution.

The concentration theorem [18] states that if P n(l) is the expected fraction of incorrect

messages passed during the l-th decoding iteration for a block-length n code ensemble,

then the probability of the actual fraction of incorrect messages for a sample code of the

ensemble being outside of (P n(l)− δ, P n(l) + δ) tends to 0 exponentially with n, for any

δ > 0.

Given the concentration theorem, the problem of analyzing the error performance of

a particular code Tanner graph is converted to analyzing the expected performance of all

Tanner graphs in the ensemble. At first this appears to be an even more difficult problem,

but the expansion of a code to its ensemble allows for a second simplifying theorem to

be applied.

A second key theorem in [18] states that the ensemble expected fraction of incorrect

messages P n(l) converges to P∞(l) as n tends to infinity, where P∞(l) is the fraction

of incorrect messages passed during iteration l assuming the decoding neighbourhood of

depth l is cycle-free.

The decoding neighbourhood of depth l for a variable node vi is the recursive expan-

sion of edges and neighbouring nodes of vi in l decoding iterations. An additional level

of check and variable nodes is added with every iteration. A length 2l cycle exists if vi

appears in its own decoding neighbourhood of depth l. Intuitively, the presence of cycles

means the message received by vi after l decoding iterations is necessarily dependent on

previous messages from vi, hence messages are correlated. Correlated message passing is

extremely difficult to analyze. The convergence of ensemble expectation to the cycle-free

Chapter 2. Technical Background 20

case allows for the assumption that all messages are independent, greatly simplifying

analysis. The results for the cycle-free graph directly apply to the ensemble expectation,

since they are equal as n tends to infinity.

The fraction of error messages of the cycle-free graph is analyzed using Density Evo-

lution (DE) [18]. DE tracks all messages in the BP decoding algorithm for the cycle-free

graph realization of the code ensemble, over all possible channel outputs and transmitted

codewords. As the name suggests, DE operates on probability densities of the channel

outputs and messages.

Let P0 denote the density of the channel output LLR. Initially, the variable nodes all

send their channel output, therefore the density of the µvi→j messages is P0. At the check

nodes, let Γ() and Γ−1() be a transform and its inverse that implements (2.9) and allows

for the message calculation in the transform domain to be a sum. Such a transform is

given in [18]. For a degree i check node, with independent incoming messages, the output

message density is given by

Γ−1(Γ(P0)⊗(i−1)) (2.10)

where⊗ denotes convolution and exponentiation is a shorthand for repeated convolutions.

Averaging over all check node degrees and their respective edge-perspective distribu-

tion coefficients, we obtain the µcj→i message density after 1 iteration as

Q1 = Γ−1(ρ(Γ(P0))) = Γ−1

(dc∑i=2

ρiΓ(P0)⊗(i−1)

). (2.11)

In the cycle-free decoding neighbourhood, all messages remain independent after node

updates. Therefore the variable message density after 1 iteration is

P1 = P0 ⊗ λ(Q1) = P0 ⊗dv∑i=1

λ⊗(i−1)i . (2.12)

For any iteration l, the recursive density update for µvi→j messages is

Chapter 2. Technical Background 21

Pl = P0 ⊗ λ(Γ−1(ρ(Γ(Pl−1)))). (2.13)

The density evolution threshold for the AWGN channel is defined to be σ∗ such that

Pl → 0 as l →∞ for all σ < σ∗. In [15] a quantized version of density evolution named

discrete density evolution is given with good implementation and numerical stability

characteristics. All thresholds reported in this thesis are evaluated using discrete density

evolution.

To summarize, in ensemble-based LDPC code analysis, for a given pair of variable and

check node degree distributions, the goal is to evaluate the expected fraction of incorrect

messages for the ensemble of all code Tanner graphs, channel outputs and transmitted

codewords. At large block-lengths, the faction of incorrect messages of any specific code

Tanner graph is concentrated around the ensemble expectation. The expectation is shown

to be asymptotically equal to the fraction of incorrect messages of the cycle-free decoding

neighbourhood, which can be analytically determined using density evolution.

Gaussian-approximated density evolution

Density evolution is an effective analytical tool for finding the threshold of LDPC code

ensembles parameterized by degree distribution pairs. However, as a synthesis method

to find good degree distribution pairs it is overly complex to be useful in optimization.

Originally, [14] used the genetic algorithm differential evolution [32] to optimize degree

distribution pairs. Such heuristic algorithms are prone to being trapped in local minima

and does not give any convergence guarantee. In addition, since every optimization

iteration requires many threshold evaluations, this leads to extremely high runtimes.

The complexity of density evolution can be reduced if the message densities (Pl, Ql)

are approximated by using symmetric Gaussian densities. Intuitively, since variable node

updates are convolutions of independent input message densities, by the central limit

theorem [33] the output message density is approximately Gaussian for high node degrees.

Chapter 2. Technical Background 22

For the purpose of this discussion, we define a symmetric Gaussian density to be a

Gaussian density with σ2 = 2µ [29]. Consequently, only one parameter µ is needed

to fully specify the density function. This Gaussian-approximated density evolution

(GA-DE) was introduced in [34] where the authors found degree distribution pairs with

thresholds within 0.02 dB of full density evolution designs.

The key simplifying aspect of GA-DE is the node update equations are no longer

operations on densities but on the single representative parameter. The convolution of

symmetric Gaussian densities with mean µ at a degree i variable node simply results in

an output Gaussian of mean (i−1)µl−1 +µ0 where µ0 is the mean of the channel output.

The check node update can be similarly condensed into an expression involving the input

message means only. An entire GA-DE iteration can be expressed from the perspective

of the average variable output message mean µvl as

µvl = f(µv0, µvl−1) (2.14)

where f is the single-variable function model of (2.13). It has been shown in [34] that

the condition f(µv0, µvl−1) > µvl−1 is necessary and sufficient to ensure convergence to zero

incorrect messages in GA-DE. More importantly, this convergence condition is linear,

thus allowing optimization to be achieved by using efficient linear programming.

GA-DE was an early instance of single-parameter approximations of density evolu-

tion. The idea of single-parameter approximation is to apply the symmetric Gaussian

approximation to reduce the cumbersome operations of (2.13) to iterated functions of one

parameter. Powerful optimization tools can then be applied to the iterated functions.

The single-parameter approximation design technique was further studied in [35] using

a semi-Gaussian approximation. The modification improved threshold accuracy and de-

sign flexibility as the original GA-DE did not work well for variable degrees greater than

10 [34].

Chapter 2. Technical Background 23

2.2.2 Extrinsic information transfer charts

The most widely used single-parameter approximation of density evolution is the extrinsic

information transfer (EXIT) technique. In essence, EXIT uses the “extrinsic information”

parameter as the single-parameter approximation for GA-DE. The definition of extrinsic

information is based on the extrinsic processing principle of iterative decoding algorithms.

Extrinsic processing is evident in BP update equations (2.8), (2.9) where the out-going

message from node i to node j always excludes the incoming message from node j to

node i. On a cycle-free decoding graph, the extrinsic processing principle ensures that a

node will never receive messages dependent on itself.

Assume a variable node initially receives an incorrect channel output LLR, to correct

this bit it must eventually receive sufficiently correct extrinsic messages. This means the

mutual information between the extrinsic messages of bit i and the value of bit i must

eventually converge to 1 as l→∞. The mutual information is defined in [36] as

I(X;L) = H(X)−H(X|L)

= 1−∫∞−∞

e−(ξ−σ2/2)2/2σ2√2πσ2

log2 [1 + e−ξ]dξ

≡ J(σ)

(2.15)

where X is the uniform binary random variable representing a codeword bit and L is the

demapped LLR output from a symmetric AWGN channel of variance σ2.

Using this conversion between σ and extrinsic mutual information, the update equa-

tions for EXIT-based GA-DE are

Ivl =dv∑i=1

λiJ

(√(i− 1)[J−1(Icl )]

2 + σ2ch

), (2.16)

Icl = 1−dc∑i=2

ρiJ(√

(i− 1)[J−1(1− Ivl−1)]2)

(2.17)

where v, c superscripts indicate the extrinsic mutual information updates due to variable

or check nodes, l is the decoding iteration, and σ2ch is the channel output LLR variance.

Chapter 2. Technical Background 24

The functions J(σ) and J−1(I) can be pre-calculated or approximated as in [36]. Note

the extrinsic informations are averaged over different variable and check node degrees.

An approximation is made in (2.17) to find the check node mutual information update

based on duality between parity-check and repetition codes. For a full justification please

refer to [36, 37]. The condition for successful decoding is Ivl → 1 as l → ∞. Successful

decoding is defined to be the existence of a sequence of codes of block-length n such that

the probability of bit error goes to 0 as n → ∞ and l → ∞. Conversely, decoding is

unsuccessful for a code ensemble if the probability of bit error is bounded away from 0

as n→∞ and l→∞ [29].

Combining (2.16) and (2.17) into a function f , the equivalent convergence condition

for successful decoding is [35]

f(Iv, σ2ch) > Iv, ∀Iv ∈ [J(σch), 1) (2.18)

Observe that (2.16) is a linear combination with coefficients λi, so that (2.18) can be

re-written as

Ivl =dv∑i=1

λifi(Ivl−1, σ

2ch), (2.19)

where fi captures the extrinsic mutual information transfer of only degree i variable nodes

and one particular check degree. We call these functions elementary EXIT functions. The

code design optimization problem is a linear programming problem that maximizes the

code rate 1− (∑ρj/j/

∑λi/i) over variable node degree distributions λi given by

maxλi

∑i≥1

λii

λi ≥ 0∑i≥1 λi = 1∑

i≥1 λifi(Ivl−1) > Ivl−1, ∀Ivl−1 ∈ [J(σch), 1).

(2.20)

Chapter 2. Technical Background 25

Fig. 2.3 shows the EXIT chart of an optimized ensemble from [35]. The dotted lines

are the elementary EXIT functions for dv = (2, . . . , 30) and a check node of degree 6. The

EXIT curve, in solid blue, is the linear combination of the elementary EXIT functions

weighted by the optimized variable node degree distribution.

The black reference line demarcates improving and degrading variable node mutual

information. Notice that the EXIT curve always lies above the reference line, satisfying

the convergence condition, thus allowing successful decoding as shown by the staircase

line.

A final point on EXIT optimization. The problem setup assumes a concentrated check

degree. The negligible performance impact of this simplification has been justified in past

literature [15,16,36]. We will also follow this simplification in our work.

0.4 0.5 0.6 0.7 0.8 0.9 1

0.4

0.5

0.6

0.7

0.8

0.9

1

Il−1v

I lv

2

30

Figure 2.3: EXIT chart of optimized rate = 0.33 ensemble at threshold of -1.91 dB Es/N0.

Chapter 2. Technical Background 26

2.3 Multi-edge-type LDPC Codes

The key motivation for the development of multi-edge-type LDPC codes is to impose

structure over the random single-edge-type code ensemble defined by pairs of variable

and check node degree distribution polynomials. Here we describe two examples where

prudently imposed structure leads to complexity reduction or performance improvement

over completely structureless code ensembles.

Given a maximum variable node degree, it has been observed that under density

evolution optimization, variable node degree distributions heavily utilize variable nodes

of the highest degree in order to achieve capacity-approaching performance [14]. In

practical implementation, the complexity of the decoder hardware scales directly with

the maximum variable degree. Therefore, it is desirable to reduce the maximum variable

degree while maintaining the capacity-approaching performance. High degree variable

nodes are appealing since they help to propagate any reliable intrinsic information and

extrinsic information that they are likely to produce. Under the purely random socket

assignment of single-edge-type LDPC codes, a very high variable node degree is required

to achieve this “spreading” effect with sufficiently high probability. However, if a code

designer imposes structure on maximum degree variable nodes, for example by avoiding

connections to many degree 2 nodes, then the same effect can be achieved with high

probability for a lower maximum variable degree [29, pp. 384-389].

A second example concerns degree 1 variable nodes in LDPC code ensembles. Since a

degree 1 variable node only sends its channel observation during message-passing decod-

ing, if it receives an erroneous channel observation then any check node it is connected

to is likely to pass on the erroneous message to its neighbouring variable nodes. It is

easy to see that if two or more degree 1 variable nodes are connected to the same check

node and a few receives erroneous channel observations, they will never be corrected

under message-passing decoding. Density evolution on a single-edge-type ensemble with

degree 1 variable nodes correctly gives a bit error probability bounded away from zero.

Chapter 2. Technical Background 27

For multi-edge-type code ensembles, the code designer can explicitly impose structure

on degree 1 variable nodes to eliminate the case where two or more are connected to the

same check node. With the extra structure, the bit error probability can be made to

go to zero for infinite block-length [29, pp. 394-397]. The inclusion of degree 1 variable

nodes brings many benefits such as lower error floor, improved decoding threshold, and

simpler implementation.

Technically, the main difference between multi-edge and single-edge LDPC code en-

sembles is that the edges between variable and check nodes are assigned to more than

one type. Refer to the single-edge-type Tanner graph example in Fig. 1.2, if we assign

the edges of all degree 1 variable nodes to the red edge-type, and the edges of all other

variable nodes to the blue edge-type, then we obtain the multi-edge-type Tanner graph

representation of Fig. 2.4.

Recall the useful concept of variable and check node sockets from Sec. 2.2. In multi-

edge-type ensembles, sockets are also assigned different edge-types. For example, in Fig.

2.4, the left-most check node has 1 red socket and 3 blue sockets, while the right-most

variable nodes has 0 red sockets and 3 blue sockets. Only sockets of the same type can

be connected by an edge of that type.

The tremendous range of code structure can be appreciated by considering both the

most general case of defining only one edge-type which is exactly the same as traditional

code ensembles, and the most specific case where each edge is assigned a different edge-

type, resulting in the definition of a single Tanner graph. The code structures between

these two extremes are of the highest interest in applications of multi-edge-type LDPC

+ + +

Figure 2.4: A possible multi-edge-type representation of the Tanner graph in Fig. 1.2

Chapter 2. Technical Background 28

codes.

In the rest of this section we overview the notation used to work with multi-edge-type

code ensembles, closely following the treatment in [29, pp. 382-397]. The emphasis will

be on the distinguishing features between single and multi-edge-type notations. It may

be helpful for the reader to review the single-edge-type notations in Sec. 2.2 to compare

with those introduced here.

The notation used to distinguish edge-types in multi-edge-type (MET) degree distri-

butions extends the placeholder variable x and node degree d to vectors x and d, where

each vector component refers to an edge-type. Whereas a single-edge-type check node of

degree 4 in the node-perspective is denoted by x4, a check node of 3 edge-types of degree

d = (2, 3, 4) is denoted in the node-perspective by x21x32x

43. In Fig. 2.4, the left-most

check node would be denoted by x11x32. A variable node in an MET ensemble has the

additional vector r specifying the channel output densities to which it is connected. A

variable node with 2 type 1 sockets and 2 type 2 sockets receiving the channel output

density 1 is denoted by x21x2xr1 while the same variable node receiving channel output

density 2 is denoted by x21x22r2. In addition to imposing structure, the ability to assign

different channel output densities to different variable nodes is another reason for using

MET ensembles in this work. Our goal is to exploit the different output densities of

bit-channels using inherent differences in LDPC codeword bit protection.

Throughout this thesis, the total number of edge-types is denoted by T and indexed

by k = (1, . . . , T ). The total number of distinct channel output densities is denoted by

S and indexed by s = (0, . . . , S). Note that there are S + 1 channel output densities,

however the s = 0 density corresponds to the channel output of a punctured variable node

which is not used in this work. All edge-type specific quantities such as the maximum

variable or check degree, or edge-perspective degree distribution, will be distinguished by

a superscript. For example d1v or λ1 are quantities of the edge-type 1. Finally, let ∂x be

a shorthand for the partial differentiation operator ∂∂x

.

Chapter 2. Technical Background 29

The most general node-perspective degree distribution pair for an MET ensemble is

given by

L(x, r) = L(x1, . . . , xT , r0, . . . , rS) =

d1v∑d1=1

· · ·dTv∑dT=1

S∑s=0

Ld1,...,dT ,s xd1

1 . . . xdT

T rs

R(x) = R(x1, . . . , xT ) =

d1c∑d1=1

· · ·dTc∑dT=1

Rd1,...,dT xd1

1 . . . xdT

T

(2.21)

To make sure the number of sockets of the each type is kept equal between variable and

check nodes, the degree distribution pair (2.21) must satisfy the socket-count constraints

∂xkL(1,1) = ∂xkR(1), ∀k = (1, . . . , T ). (2.22)

Futhermore, (2.21) must also maintain the correct fraction of distinct channel output

densities by satisfying the channel-ratio constraints

∂rsL(1,1) = πs, (2.23)

where πs is the fraction of channel output density s over all channel output densities.

The code rate is given by

r = L(1)−R(1). (2.24)

Note that all constraints and the code rate are linear in the coefficients of degree

distribution polynomials.

The edge-perspective degree distributions used by density evolution can be calculated

by taking partial derivatives with respect to each edge-type and normalizing

Chapter 2. Technical Background 30

(λ1(x1), λ

2(x2), . . . , λT (xT )

)=

(∂x1L(x, r)

∂x1L(1,1),∂x2L(x, r)

∂x2L(1,1), . . . ,

∂xTL(x, r)

∂xTL(1,1)

)(2.25)

(ρ1(x1), ρ

2(x2), . . . , ρT (xT )

)=

(∂x1R(x)

∂x1R(1),∂x2R(x)

∂x2R(1), . . . ,

∂xTR(x)

∂xTR(1)

). (2.26)

There are T edge-perspective variable (check) node degree distribution polynomials.

Practically, this means density evolution now tracks T message densities to determine the

infinite block-length ensemble threshold. Since EXIT chart analysis is a one-dimensional

approximation of density evolution, the extrinsic mutual information that must be con-

sidered in the MET EXIT chart is also expanded to a vector of T components. One

of the main contributions of this thesis is to develop an analytical and design technique

based on multi-dimensional EXIT vector fields for a specific MET ensemble defined for

LDPC coded modulation.

Chapter 3

Multi-edge LDPC Coded

Modulation

In this chapter we develop the main contributions of this thesis. In Sec. 3.1 the general

MET ensemble is reduced to a specific parameterization for LDPC coded modulation.

Thorough justifications are given for all simplifications. Sec. 3.2 develops the main

analytical tool for the specified MET ensemble: the multi-dimensional EXIT vector field.

Several properties of the vector field are proved. Code design using the multi-dimensional

EXIT vector field is accomplished after deriving the multi-edge-type convergence criterion

based on the fixed points of the iterated system.

3.1 Multi-edge Parameterization

We seek to leverage two important properties unique to multi-edge-type (MET) LDPC

framework in our coded modulation design. MET ensembles allow, as input, more than

one channel output density at variable nodes. This is precisely the desired property for

incorporating bit-level differences into the ensemble optimization process. Furthermore,

the expanded number of edge-types offers flexible control over the structure of the LDPC

ensemble. Structural features can be defined in the ensemble definition before optimiza-

31

Chapter 3. Multi-edge LDPC Coded Modulation 32

tion. Several reasons exist for imposing code structure, most common are to reduce

design and implementation complexity or to lower the error floor.

In this work, we specify a MET structure for complexity reduction. The number of free

coefficients in the general MET variable and check degree distribution polynomials (2.21)

grows exponentially with the number of edge-types. Taking into account the plausible

number of distinct channel output densities, for example 5 in the case of Gray-labelled

1024-QAM, optimizing the general degree distributions quickly becomes intractable.

We would like to simplify the parameterization to a manageable complexity without

sacrificing the desired properties of the MET framework. This can be achieved by first

assigning one edge-type to each distinct bit-channel output density. For example, in

Gray-labelled 16-QAM there are 4 bit-channels but only 2 distinct bit-channel output

densities, therefore only 2 edge-types are used; whereas for set-partition labelled 16-

QAM there are 4 distinct bit-channel output densities, requiring 4 edge-types in the

MET parameterization.

In addition, each variable node is restricted to have sockets of only one edge-type,

while different edge-type sockets are present at check nodes. We are inspired to make

this simplification by the MLC scheme, where a distinct LDPC code is optimized for each

bit-level in order to approach the capacity of the overall symbol channel. We extend the

idea by allowing variable node messages of different edge-types to interact at check nodes,

and more importantly, by optimizing the code across all bit-channels simultaneously.

With these two restrictions on the general MET ensemble, a flexible trade-off between

designing one code for an averaged channel (BICM) and designing distinct codes for dis-

tinct bit-channels (MLC) is achieved by our specific MET parameterization. The single,

optimized code under our MET parameterization will not only be properly matched to

each bit-channel, but also to the overall high-order modulation symbol channel. Fig. 3.1

(b) illustrates the specified MET parameterization for the case of 2 distinct bit-channels

such as Gray-labelled 16-QAM.

Chapter 3. Multi-edge LDPC Coded Modulation 33

σ21 σ2

1 σ21 σ2

1 σ21 σ2

1 σ21 σ2

1 σ21

π

(a) Single-edge-type LDPC Tanner graph

σ21 σ2

1 σ21 σ2

1 σ21

σ22 σ2

2 σ22 σ2

2

π1

π2

(b) Multi-edge-type LDPC Tanner graph

Figure 3.1: Tanner graph of the 2 edge-type specified MET parameterization, node

degrees are illustrative and not meant to be realistic.

Chapter 3. Multi-edge LDPC Coded Modulation 34

The variable degree distribution for the specific MET parameterization is

L(x1, . . . , xT , r1, . . . , rT ) =T∑k=1

dkv∑i=1

Li,kxikrk. (3.1)

The index k serves the dual purposes of indexing the edge-types and channel densities,

since they are the same under our specification. The bit-channel density of index k = 0

is removed since puncturing is not considered in this work. From (3.1) and Fig. 3.1 it is

clear that the MET parameterization from the variable node perspective is identical to

a single-edge-type parameterization.

In fact, if the check node degree distributions are specified such that no mixing of

edge-types can occur, the specific MET parameterization degenerates to the MLC scheme.

However, we do allow different edge-types at the check nodes, which allows messages from

different bit-channels to mix. As a final simplification, we require the total check node

degree dc be concentrated to only one value. The total check node degree is the number

of all sockets at a check node, regardless of type. Under this simplification, (2.21) can be

written as

R(x1, . . . , xT ) =∑

{d1,...,dT |d1+···+dT=dc}

Rd1,...,dT xd11 . . . xdTT . (3.2)

Even after concentrating to one total check degree, the check degree distribution

remains overly complex. The additional complication is in choosing the assignment of

check node sockets to different edge-types, under the same total degree. For 2 edge-types

and a total check degree of dc there are dc + 1 possible edge-type assignments. For more

than 2 edge-types the number of possible assignments grows rapidly, and is related to

the partition function P [dc] from number theory [38]. In order to gain insight into this

problem and to explore the possibility of concentrating to only one check node edge-type

assignment, we undertook an empirical study using simple MET LDPC ensembles with

2 edge-types.

Chapter 3. Multi-edge LDPC Coded Modulation 35

3.1.1 Check degree edge-type assignment

The goal of this study is to justify further simplifying the check node degree distribution

to a single term, given by

R(x1, . . . , xT ) = Rd1c ,...,dTcxd1c1 . . . x

dTcT , (3.3)

where d1c + · · ·+ dTc = dc is a chosen check degree edge-type assignment. Note the direct

use of dkc to denote the number of check sockets of type k, since only one term of the sum

is present.

The first code under study is the regular (3,6) ensemble, where the notation corre-

sponds to (dv,dc), of rate 1/2. For Gray-labelled 16-QAM with 2 distinct bit-channels,

the degree 6 check node can be split into (d1c , d2c) = {(0, 6), (1, 5), . . . , (6, 0)} over the

2 edge-types. However, since the variable node parameterization imposes a constraint

(2.22) on the number of edges of each type, we focus on the case of pairs of check degree

edge-type splits where the distribution coefficients can be directly found. For example,

the check degree distribution polynomial for the pair of assignments (0,6), (4,2) is

R(x1, x1) = R1x62 +R2x

41x

22, (3.4)

where R1,R2 are obtained by substituting (3.4) into (2.22) and solving the system. The

thresholds for all pairs of check node splits are given in Table 3.1, the pairs that do not

satisfy the edge-count constraint are marked by “-”.

The check node split pairing with the highest density evolution threshold is the sin-

gle symmetrical (3,3) split. The next highest threshold belongs to the pair of nearly-

symmetrical (2,4), (4,2) splits. From this simple example it appears that when the check

degree polynomial is restricted to pairs of edge-type assignments as in (3.4), concentrat-

ing to a single symmetrical split gives the highest threshold.

To see if the same property can be observed for an odd total check degree, we repeated

Chapter 3. Multi-edge LDPC Coded Modulation 36

Table 3.1: Density evolution thresholds for the (3,6) regular LDPC check degree edge-

type split pairings under Gray-labelled 16-QAM.

σ∗ (0, 6) (1, 5) (2, 4) (3, 3) (4, 2) (5, 1) (6, 0)

(0, 6) -

(1, 5) - -

(2, 4) - - -

(3, 3) - - - 0.4808

(4, 2) 0.4711 0.4768 0.4793 - -

(5, 1) 0.4439 0.4737 0.4778 - - -

(6, 0) 0.3925 0.4742 0.4768 - - - -

the above study for the (3,9) rate 2/3 regular LDPC code. Since the total check degree

is odd, a symmetrical split is not possible. It is hypothesized that the near-symmetrical

pair of (4,5), (5,4) will give the highest threshold. Table 3.2 shows the thresholds for the

(3,9) code.

The simulated thresholds confirm the near-symmetrical pairing (4,5), (5,4) to be the

best check degree splits. Again note the drop in threshold as the edge-type splits move

away from symmetry.

Finally, our last empirical example uses an irregular LDPC ensemble defined in single-

edge-type by λ(x) = 1/3x1 + 2/3x3, ρ(x) = x5 of rate 1/2. The check degree edge-type

splits are exactly the same as the (3,6) regular case. The thresholds are given in Table

3.3.

The same pattern of threshold/split-pair correspondence as Table 3.1 can be ob-

served, with the only difference being a slightly higher threshold due to the variable node

irregularity.

The above empirical studies of check degree edge-type splits between 2 edge-types

point to the conclusion that a concentrated symmetric (or near-symmetric pairing in the

Chapter 3. Multi-edge LDPC Coded Modulation 37

Table 3.2: Density evolution thresholds for the (3,9) regular LDPC check degree edge-

type split pairings under Gray-labelled 16-QAM.

σ∗ (0, 9) (1, 8) (2, 7) (3, 6) (4, 5) (5, 4) (6, 3) (7, 2) (8, 1) (9, 0)

(0, 9) -

(1, 8) - -

(2, 7) - - -

(3, 6) - - - -

(4, 5) - - - - -

(5, 4) 0.3621 0.3628 0.3628 0.3634 0.3634 -

(6, 3) 0.3587 0.3607 0.3621 0.3628 0.3634 - -

(7, 2) 0.3525 0.3587 0.3607 0.3621 0.3634 - - -

(8, 1) 0.3340 0.3573 0.3601 0.3621 0.3634 - - - -

(9, 0) 0.3168 0.3525 0.3594 0.3614 0.3634 - - - - -

case of odd total check degree) split of total check degree is optimal for all pairs of splits.

The fact that the irregular ensemble also shows the same property is highly encouraging

in extending this “concentration to symmetrical edge-type assignments” observation to

more complicated irregular ensemble parameterizations.

We conjecture that by concentrating the check node edge-type assignment to a sin-

gle symmetric or near-symmetric split for Gray-labelled 16-QAM, the performance loss

from a more general linear combination of check degree splits is minimal. Therefore,

for subsequent code design of Gray-labelled 16-QAM, we will assume the concentrated

symmetrical edge-type split. For higher-order coded modulation design, an extension

technique will be introduced in Sec. 3.2.2 to alleviate the high complexity of optimizing

the check degree edge-type assignment for a concentrated total degree.

In summary, the specific MET parameterization for coded modulation assigns a dif-

ferent edge-type to each distinct bit-channel output density. A variable node can only

Chapter 3. Multi-edge LDPC Coded Modulation 38

Table 3.3: Density evolution thresholds for irregular LDPC check degree edge-type split

pairings under Gray-labelled 16-QAM.

σ∗ (0, 6) (1, 5) (2, 4) (3, 3) (4, 2) (5, 1) (6, 0)

(0, 6) -

(1, 5) - -

(2, 4) - - -

(3, 3) - - - 0.4946

(4, 2) 0.4854 0.4906 0.4926 - -

(5, 1) 0.4617 0.4879 0.4916 - - -

(6, 0) 0.4059 0.4886 0.4906 - - - -

belong to one edge-type and check nodes are concentrated to one total degree. Fur-

thermore, check node edge-type assignments are concentrated to one particular vector of

check degrees (d1c , . . . , dTc ). The multi-edge-type ensemble parameterization used in this

thesis is given by the pair of variable and check node degree distribution polynomials

(3.1) and (3.3).

3.2 Multi-edge Optimization

We seek to optimize the MET ensemble by using single-parameter based LDPC design.

Single-parameter LDPC design refers to all methods that approximates full density evolu-

tion by assuming symmetric Gaussian intermediate message densities, which can be fully

characterized using a single parameter such as the mean, variance, probability of error, or

extrinsic mutual information [35]. Recall from Sec. 2.2.2 that the EXIT chart technique

tracks the change of the average extrinsic mutual information (2.15) through variable

and check node updates during decoding iterations. The EXIT functions were scalar

valued thus analysis and design can be easily organized by plotting both variable and

Chapter 3. Multi-edge LDPC Coded Modulation 39

check transfer curves on one coordinate plane and solving a curve fitting problem [35,36].

3.2.1 Multi-dimensional EXIT vector field

In multi-edge-type ensembles, the MET density evolution as given by (2.25),(2.26) con-

tains as many distinct densities as the number of edge-types, T . Therefore, the single-

parameter EXIT approximation of MET density evolution uses a vector of mutual in-

formations to keep track of all edge-type message densities. In other words, the EXIT

chart is now multi-dimensional. The key contribution of this thesis is developing effi-

cient and accurate analysis and design methods for a specific MET ensemble based on

multi-dimensional EXIT charts.

For illustrative purposes, we focus on variable mutual information in the EXIT up-

date equations. A full iteration of the EXIT update equations maps the variable node

extrinsic mutual information in the previous iteration Ivl−1 to the output extrinsic mutual

information of the current iteration Ivl , while the check node update is implicitly nested

into the update as shown

Ivl = f v(f c(Ivl−1), Iv0 ). (3.5)

This expression can be fully determined if the check node degree distribution has

been given. This can be satisfied either by concentrating the check node to one degree,

as we have done in our parameterization, or by an iterative design procedure where one

of the check or variable node distributions is assumed to be fixed while the other is being

optimized [29, pp. 239-240]. It is not difficult to derive the check node mutual information

analogue of the analysis and design procedures. However, in our development we shall

only focus on the variable mutual informations. With this understanding, we drop the v

superscript to avoid excessive notation.

In general, optimization based on the multi-dimensional EXIT chart is as difficult

as directly optimizing using MET density evolution. The mixing of different edge-type

Chapter 3. Multi-edge LDPC Coded Modulation 40

densities at both variable and check nodes complicates the EXIT chart and prohibits

an efficient optimization procedure. An additional edge-type exponentially increases

the number of EXIT functions in the optimization problem. This may be why prior

work on single-parameter analysis and design of MET ensembles has been scarce, where

as EXIT techniques for single-edge ensembles have flourished. A review of literature

revealed only [25,39] as attempts at EXIT-based MET ensemble optimization. Only [39]

explicitly defined multi-dimensional EXIT charts but fell short of providing an effective

optimization procedure.

Keeping design complexity low while retaining the desired properties of MET ensem-

bles has been a guiding principle throughout this work. It is this disciplined approach

that allows for the reduction in complexity of multi-dimensional EXIT charts to allow for

an efficient optimization procedure. The key simplification in the MET parameterization

of Sec. 3.1 is to restrict variable nodes to only one edge-type, determined by its assigned

bit-channel. Under this restriction, from the variable node perspective the EXIT charts

are exactly the same as the single-edge case, as shown by the variable node update equa-

tions of the multi-dimensional EXIT chart for the vector of variable mutual informations

(I1l , . . . , ITl )

Ikl =

dkv∑i=2

λki J[(i− 1)J−1(Ic,kl ) + σ2

k

](3.6)

where σ2k is the LLR variance of the bit-channel output assigned to edge-type k. Unless

specifically noted, all expressions in this section containing the index k are to be un-

derstood as the set of T expressions over all edge-types (1, . . . , T ), indexed by k. The

mutual information conversion functions J(σ2), J−1(I) denote the composite functions

J(√σ2) and [J−1(I)]2 respectively.

The check node mutual information update is slightly more complex since all edge-

types mix at check nodes. Given the concentrated check node edge-type assignment

vector (d1c , . . . , dTc ) the multi-dimensional EXIT check node update expression is

Chapter 3. Multi-edge LDPC Coded Modulation 41

Ic,kl = 1− J

(dkc − 1)J−1(1− Ikl−1) +T∑t=1t6=k

dtcJ−1(1− I tl−1)

(3.7)

where the coefficients {ρk} have been removed since the concentrated check degree edge-

type assignment in (3.3) means {ρk ≡ 1}. The check node update expression is similar to

the single-edge type version, with the main difference being the input mutual informations

now come in T types, and the output consists of T simultaneous mutual information

updates.

For a full iteration update of the variable vector mutual information, substitute the

appropriate edge-type output of (3.7) into (3.6) to obtain

Ikl =

dkv∑i=2

λki J

(i− 1)J−1

1− J

(dkc − 1)J−1(1− Ikl−1) +T∑t=1t6=k

dtcJ−1(1− I tl−1)

+ σ2

k

.(3.8)

To clarify the functional relationships between all input and output mutual infor-

mation components in (3.8), it is more convienent to encapsulate the expression within

function representations. Let Ic,kl = f c,k(I1l−1, . . . , ITl−1) denote the check node update

(3.7) for mutual information of edge-type k . Let Iv,kl = f v,k(Ic,kl , Ik0 ) denote the variable

node update (3.6) for mutual information of edge-type k, where Ik0 = J(σ2k) is the mu-

tual information of the LLR density from the bit-channel corresponding to edge-type k.

Expression (3.8) can now be written as

I1l = f v,1(f c,1(I1l−1, . . . , ITl−1), I

10 )

I2l = f v,2(f c,2(I1l−1, . . . , ITl−1), I

20 )

......

...

ITl = f v,T (f c,T (I1l−1, . . . , ITl−1), I

T0 )

, (3.9)

or in bold-font notation

Chapter 3. Multi-edge LDPC Coded Modulation 42

Il = f(Il−1). (3.10)

We see that the multi-dimensional EXIT update (3.10) is a vector field in T dimen-

sional space RT .

The domain D of f is the Cartesian product of all closed real intervals between bit-

channel output mutual informations Ik0 and 1 for each edge-type

D = [I10 , 1]× [I20 , 1]× · · · × [IT0 , 1]. (3.11)

Here we state a property of the vector field f .

Proposition 3.1. For any set of edge-perspective variable degree distributions {λki }dkvi=2,

f : D 7→ D

Proof. By definition all bit mutual informations are between 0 and 1. Looking at expres-

sion (3.6), the inner-most check node output Ic,kl = f c,k(Il) is greater or equal to 0 for all

Il ∈ D since it is a mutual information. Since the function J−1(Ic,kl ) returns the standard

deviation of the symmetric Gaussian density corresponding to the check output mutual in-

formation, it is also greater than 0, therefore (i−1)J−1(Ic,kl )+σ2k ≥ σ2

k. Using the fact that

J(σ2) is monotonically increasing in σ2 [40] we have J((i−1)J−1(Ic,kl ) +σ2k) ≥ J(σ2

k), ∀i.

Since the degree distribution coefficients {λki }dkvi=2 are probabilities thus greater than or

equal to 0, we have

Ik0 ≤dkv∑i=2

λki J[(i− 1)J−1(Ic,kl ) + σ2

k

]≤ 1 (3.12)

where J(σ2k) is replaced by Ik0 since they are equivalent. Since ∀Il ∈ D and each compo-

nent of the vector field update f(Il) is in [Ik0 , 1], we have f : D 7→ D.

Fig. 3.2 illustrates a 2 edge-type Gray-labelled 16-QAM multi-dimensional EXIT

vector field f for an optimized multi-edge ensemble at the threshold noise level. The

Chapter 3. Multi-edge LDPC Coded Modulation 43

two-component input mutual information vector is plotted on both subplots and labelled

as I1l−1 and I2l−1. The domain D is the x-y plane. The vector field expression for this

example is

f(Il−1) =

I1l = f v,1(f c,1(I1l−1, I2l−1), I

10 )

I2l = f v,2(f c,2(I1l−1, I2l−1), I

20 )

. (3.13)

The components of the vector field output have been separated into two subplots with

I1l corresponding to the left subplot and I2l corresponding to the right subplot.

Each subplot consists of two surfaces and an EXIT decoding path. Take for example

the left subplot which describes the mutual information transfer characteristic for every

point in D. At the initial point (I10 , I20 ) = (0.610, 0.905), the top surface is the EXIT

transfer function f v,1(f c,1(I1l−1, I2l−1), I

10 ), referred to here as F1. The bottom surface is

the “no-improvement” reference where the output type 1 mutual information is equal

to the input type 1 mutual information, referred to here as R1. At the beginning of

decoding, surface F1 is above R1, indicating that the mutual information of type 1 im-

proves with the initial EXIT update. Similarly, at the same point in the right subplot,

the EXIT transfer function f v,2(f c,2(I1l−1, I2l−1), I

20 ), denoted by F2, is also above the “no-

improvement” reference of type 2 mutual information R2. Since both edge-type mutual

informations improve with EXIT update, the decoding path can take a step forward as

shown by the staircase progression of the two blue decoding curves in both subplots.

Decoding progresses to a new mutual information input vector (I11 , I21 ).

The same reasoning applies in all subsequent EXIT vector field decoding iterations.

Right after iteration l − 1, the multi-dimensional EXIT vector field can be assumed to

have reached the point (I1l−1, I2l−1) somewhere in D. The position of the surfaces F1 and

F2 relative to their respective reference surfaces R1 and R2 determines whether or not

decoding can proceed. Intuitively, if F1 is above R1 and F2 is above R2 decoding will

proceed, and if F1 is equal or below R1 and F2 is equal or below R2 then decoding will not

Chapter 3. Multi-edge LDPC Coded Modulation 44

Figure 3.2: Multi-dimensional EXIT vector field for 2 edge-types at threshold σ∗ =

0.3665.

Chapter 3. Multi-edge LDPC Coded Modulation 45

proceed. To contrast with the successful decoding case of Fig. 3.2, Fig. 3.3 illustrates the

unsuccessful decoding case where the noise power is increased to just above the threshold

for the same optimized ensemble. The decoding path stops when it reaches a point in D

where either F1 is below R1 or F2 is below R2 or both. Recall that if decoding does not

reach the point (1, 1) in extrinsic mutual information, the probability of error is bounded

away from zero asymptotically in block-length and decoding iterations.

Analysis of MET ensembles

Figs. 3.2 and 3.3 visually illustrates the analytical power of the multi-dimensional EXIT

vector field. Even though no visual representations exist for more than 2 edge-types,

conceptually the analysis is straightforward. The procedure for analyzing multi-edge

coded modulation ensembles using the multi-dimensional EXIT vector field is as follows.

Given a set of edge-perspective variable degree distributions {λki }dkvi=2, the check degree

edge-type assignment vector (d1c , . . . , dTc ), and channel noise power σ2, iteratively evaluate

Il = f(Il−1), 1 ≤ l ≤ L, with I0 being the initial bit-channel mutual information vector

and L the maximum number of allowed iterations. If Il = 1 for some l then σ is below the

threshold σ∗, increment noise power and repeat. If l = L and Il < 1 then σ is considered

to be above the threshold σ∗, decrement and repeat. An outer binary search can be

used to efficiently determine the threshold, since the AWGN threshold is monotonic with

respect to channel degradation [29, pp. 224].

However, since multi-edge density evolution already exists as an efficient method to

determine thresholds of LDPC coded modulation ensembles, using the multi-dimensional

EXIT vector field is less accurate and redundant. The main reason for its development

is to design of multi-edge LDPC ensembles. For this reason we conclude the study of the

analytical uses of this technique and move on to code design.

Chapter 3. Multi-edge LDPC Coded Modulation 46

Figure 3.3: Multi-dimensional EXIT vector field for 2 edge-types at above threshold

σ = 0.3666.

Chapter 3. Multi-edge LDPC Coded Modulation 47

3.2.2 Design of multi-edge coded modulation

Code design using the multi-dimensional EXIT vector field is very similar to single-edge

EXIT design in all but one aspect. This is not surprising since the single-edge EXIT

chart is the special 1-dimensional instance of the multi-dimensional EXIT vector field.

The goal remains to optimize the code rate over variable node degree distributions, now

one distribution per edge-type. The solution must satisfy linear constraints on socket

count (2.22) and bit-channel ratio (2.23). Most importantly, the solution must allow for

successful decoding from the initial variable mutual information vector I0 to the point

1. It is this final convergence condition that differs between single-edge and multi-edge

EXIT design. Once we find the equivalent multi-edge convergence condition, the rest of

the optimization problem can be extended from single-edge EXIT charts.

The single-edge convergence condition (2.18) requires the variable node mutual in-

formation to increase at every point in the 1-dimensional domain D after one iteration

of variable and check node updates. This is a sufficient condition in 1-dimension since

the EXIT decoding path is also 1-dimensional. In multi-dimensional EXIT vector fields,

while the decoding path remains 1-dimensional, the domain D is not. A convergence

condition requiring improvement in all edge-type mutual informations such as

f v,k(f c,k(I), Ik0 ) > Ik, ∀I ∈ D (3.14)

is unnecessarily stringent. In the successful decoding example of Fig. 3.2, the EXIT

surface F1 is not above R1 for all of the domain D. The condition is only true in the

projection of the decoding path onto D, which is 1-dimensional. Looking along this

projected path, the EXIT vector field is again a 1-dimensional EXIT chart. This exam-

ple illustrates the difficulty in developing a convergence condition for multi-dimensional

EXIT vector fields. The decoding path changes with optimization parameters, so we do

not know which points in D belong to its projection. However, to impose the convergence

constraint (3.14) we must know the projected path. We solve this problem by taking a

Chapter 3. Multi-edge LDPC Coded Modulation 48

completely different view of the multi-dimensional EXIT vector field, focusing on the

fixed points of the iterated system.

Fixed points of multi-dimensional EXIT vector fields

A fixed point of the multi-dimensional EXIT vector field f is defined to be a point I∗ ∈ D

that satisfies

I∗ = f(I∗). (3.15)

By definition, the point 1 is a fixed point of all multi-dimensional EXIT vector fields.

In the following we provide a sufficient condition for the successful convergence of

multi-dimensional EXIT vector fields. The key to the proof lies in the monotonicity of

f with respect to I. Here we distinguish between two important notations. We make

explicit the dependence of f on bit-channel output mutual information vector I0, and

denote a single vector field update at the point I by fI0(I). We use fI0 to represent the

iterated vector field f ◦· · ·◦ f(I0). The distinguishing idea is that fI0(I) is an actual vector

field evaluation with a vector value, whereas fI0 is only a label.

Lemma 3.1. Let I2 � I1 denote the relation between two mutual information vectors

where Ik2 ≥ Ik1 ∀k ∈ (1, . . . , T ), and I2 � I1 if the components are strictly greater. Then

fI0(I2) � fI0(I1) if I2 � I1.

Proof. Recall the monotonicity property of the functions J(σ2), J−1(I) first used in

the proof of Proposition 3.1. Given this property and non-negativity of variable degree

distributions, the claim is equivalent to showing a stripped down version of the vector

field update equations (3.8)

(dkc − 1)J−1(1− Ik) +T∑t=1t6=k

dtcJ−1(1− I t) (3.16)

Chapter 3. Multi-edge LDPC Coded Modulation 49

monotonically decreases (or remains constant) for I2 � I1. Assume only one component

k′ increases between I2 and I1. Then since mutual information is upper-bounded by

1, J−1(1 − Ik′) decreases. For k = k′, since the minimum useful check degree is 2,

the term containing Ik′

in (3.16) decreases. The same argument applies for the k 6= k′

case. Therefore the claim is true for the increase of only one component of the mutual

information vector.

Since (3.16) is a sum of terms each containing a different component of the input

mutual information vector, the general case where I2 � I1 is true by the superposition

principle.

Now we are ready to prove a sufficient convergence condition for the multi-dimensional

EXIT vector field.

Theorem 3.1. Let D = D \ 1, the multi-dimensional EXIT iterated vector field fI0 will

successfully converge to 1 as l→∞, if it has no fixed points in D.

Proof. Given that fI0 has no fixed points in D and by Proposition 3.1 the vector field

only maps to D, we must have fI0(I0) � I0 otherwise I0 is a fixed point in D. Assuming

this holds for Il−1, for an iterated vector field, Il = fI0(Il−1), hence the assumption

implies Il � Il−1. From the monotonicity Lemma 3.1 we know that if Il � Il−1 then

fI0(Il) � fI0(Il−1). Since there are no fixed points in D and fI0(Il) is upper-bounded by

1, by induction we have fI0 converging to 1 as l→∞.

The backward-difference vector field

Theorem 3.1 suggests that to achieve successful convergence of the multi-dimensional

EXIT vector field, the design algorithm must avoid fixed points of fI0 in D. This can be

achieved in practice by using the backward-difference vector field ∇f defined as

Chapter 3. Multi-edge LDPC Coded Modulation 50

∇f(I) ≡ f(I)− I ≡

∇1f(I) = f v,1(f c,1(I), I10 )− I1

∇2f(I) = f v,2(f c,2(I), I20 )− I2...

......

∇T f(I) = f v,T (f c,T (I), IT0 )− IT

. (3.17)

The reason for using ∇f is straightforward since by definition of a fixed point I∗,

f(I∗) − I∗ = 0 therefore ∇f(I∗) = 0. Alternatively, one can consider the backward-

difference vector field as a difference-based representation of an EXIT decoding path.

Let {I′l} denote an EXIT decoding sequence in D where at the lth iteration, the output

of the EXIT update vector field is I′l. We can write this sequence using elements of ∇f

as

I′l = I0 +l−1∑i=1

∇f(I′i). (3.18)

If there exists i′ such that ∇f(I′i′) = 0 then decoding will not progress beyond I′i′ .

Using the backward-difference vector field, the convergence condition in Theorem 3.1 can

be written as

∇f(I) 6= 0, ∀I ∈ D. (3.19)

Fig. 3.4 shows the backward-difference vector field for the optimized 2 edge-type

Gray-labelled 16-QAM ensemble used in Fig. 3.2. The domain D has been quantized to

a 30 by 30 grid. The decoding path is represented by the black vector field streamline.

The arrows are coloured according to the basin of attraction of the fixed point to which

they belong. In the successful decoding case only the fixed point 1 exists, hence all arrows

are coloured blue to indicate that they belong to its basin of attraction.

To derive the operational convergence constraint to be used in an optimization prob-

lem, first write (3.17) explicitly showing the optimization parameters: the variable degree

distribution coefficients

Chapter 3. Multi-edge LDPC Coded Modulation 51

0.91 0.92 0.93 0.94 0.95 0.96 0.97 0.98 0.99 1

0.65

0.7

0.75

0.8

0.85

0.9

0.95

1

I2

I1

Figure 3.4: Backward-difference vector field of 2 edge-types at threshold σ∗ = 0.3665.

Chapter 3. Multi-edge LDPC Coded Modulation 52

f v,k(f c,k(I), Ik0 )− Ik =

dkv∑i=2

λki J[(i− 1)J−1[f c,k(I)] + σ2

k

]− Ik (3.20)

=

dkv∑i=2

λki J[(i− 1)J−1[f c,k(I)] + σ2

k

]− Ik

dkv∑i=2

λki (3.21)

=

dkv∑i=2

λki

(J[(i− 1)J−1[f c,k(I)] + σ2

k

]− Ik

)(3.22)

where (3.21) uses the fact that {λki }dkvi=2 is a probability.

Let gv,ki (f c,k(I), Ik0 ) denote J [(i− 1)J−1[f c,k(I)] + σ2k], we have

f v,k(f c,k(I), Ik0 )− Ik =

dkv∑i=2

λki

(gv,ki (f c,k(I), Ik0 )− Ik

)(3.23)

which is a linear combination of the backward-differences for each variable degree. Define

the elementary backward-difference to be

∇ki f(I) ≡ gv,ki (fc,k(I), Ik0 )− Ik. (3.24)

Finally, we can write ∇f(I) in terms of ∇ki f(I)

∇f(I) =

d1v∑i=2

λ1i∇1i f(I),

d2v∑i=2

λ2i∇2i f(I), . . . ,

dTv∑i=2

λTi ∇Ti f(I)

(3.25)

By using the backward-difference vector field, we recover the elementary EXIT func-

tion representation of the overall EXIT chart first given in (2.19) for single-edge EXIT

charts. Now we can give a practical convergence condition for use in the optimization

problem.

Multi-dimensional EXIT convergence condition

For the multi-dimensional EXIT vector field defined by the specialized multi-edge variable

and check node degree distributions (3.1), (3.3), with fixed maximum variable degrees

Chapter 3. Multi-edge LDPC Coded Modulation 53

dkv , check degree assignment vector (d1c , . . . , dTc ), and bit-channel noise variances σ2

k, a

sufficient condition for the convergence of the EXIT iterated vector field is

‖∇f(I)‖2 > 0, ∀I ∈ D (3.26)

or in expanded form

T∑k=1

dkv∑i=2

λki∇ki f(I)

2

> 0, ∀I ∈ D. (3.27)

Two details must be addressed in practice. The domain D must be quantized to a

finite number of points using some quantization method. We use uniform quantization

with the same resolution in each edge-type. For example a resolution of 30 points was

used in generating Fig. 3.4. Furthermore, the strict inequality in (3.26) cannot be used

as a constraint in optimization. We modify the condition slightly by introducing a small

positive tolerance ε to allow equality

‖∇f(I)‖2 ≥ ε, ∀I ∈ D. (3.28)

The tolerance ε has been observed to be highly dependent on the quantization, max-

imum variable degree dkv , bit-channel noise variance σ2k and to a lesser extent the check

degree assignment vector.

Fig. 3.5 shows the backward-difference vector field for the same optimized ensemble

as Fig. 3.4, except noise power has been increased to above threshold. There are now

two stable fixed points in the domain D at 1 and near the point where the decoding path

terminates. An unstable fixed point lies somewhere between the stable fixed points near

the separation between the red and blue arrows. Again, the arrow colours denote the

basin of attraction of fixed points. Here, the blue arrows belong to the basin of attraction

of the fixed point 1, and the red arrows belong to the basin of attraction of the lower left

fixed point.

Chapter 3. Multi-edge LDPC Coded Modulation 54

In support of the practical convergence condition (3.28), the point marked by the pink

square is where ‖∇f(I)‖2 is at its minimum value. The correspondence between where

the multi-dimensional EXIT vector field stops improving and the estimate given by the

practical convergence condition is clear. A lower resolution was used in Fig. 3.5 for

illustrative purposes. At sufficiently high resolutions, (3.28) is a reliable approximation

of the theoretical convergence condition of Theorem 3.1.

Multi-edge coded modulation optimization

Our parameterization of the multi-edge ensemble for high-order coded modulation uses

variable node-perspective degree distribution (3.1) and check node-perspective degree

distribution (3.3). In Sec. 3.1.1 we concluded that concentrating the check degree to

a single total degree with edge-type assignment vector (d1c , . . . , dTc ) does not degrade

performance significantly over general check degree distributions. Therefore, the check

node edge-type assignment vector is assumed to be given at design time. The AWGN

channel noise power σ2 is also assumed to be fixed at design time. The bit-channel

mutual informations can then be empirically determined. Lastly, the maximum variable

node degrees dkv are chosen at design time.

Under these assumptions, an optimization problem can be setup to optimize the code

rate over the variable node degree distributions{λki }dkvi=2. We derive the constraints in the

following.

Since the optimization parameters {λki }dkvi=2 are normalized edge-perspective variable

degree distributions (2.25), they are constrained to be probabilities

dkv∑i=2

λki = 1 (3.29)

λki ≥ 0 (3.30)

λki ≤ 1. (3.31)

Chapter 3. Multi-edge LDPC Coded Modulation 55

0.9 0.91 0.92 0.93 0.94 0.95 0.96 0.97 0.98 0.99 10.6

0.65

0.7

0.75

0.8

0.85

0.9

0.95

1

I2

I1

Figure 3.5: Backward-difference vector field of 2 edge-types at above threshold σ =

0.3666.

Chapter 3. Multi-edge LDPC Coded Modulation 56

Recall from Sec. 2.3 that all multi-edge degree distribution pairs (L(x, r), R(x)) must

satisfy the socket-count (2.22) and the channel-ratio (2.23) constraints. A slight compli-

cation occurs due to the constraints being stated in node-perspective degree distributions

while the optimization parameters are edge-perspective degree distributions. Assuming

(3.29) holds, there exists a conversion between the coefficients of the two degree distri-

butions

Li,k =1

T

λki /i∑i λ

ki /i

. (3.32)

The reader may verify the conversion by plugging in (3.32) into (2.25). The 1/T

factor is used to satisfy the channel-ratio constraint as explained below.

From the socket-count constraint, on the variable side

∂xkL(1,1) =∑i

iLi,k

=∑i

i1

T

λki /i∑j λ

kj/j

=∑i

λkiT∑

j λkj/j

=1

T∑

j λkj/j

(3.33)

where the last step assumes (3.29) is satisfied. On the check side

∂xkR(1) = dkcR (3.34)

together they form the constraint on edge-type k

1

Tdkc∑

j λkj/j

= R. (3.35)

For T different edge-types, there are T different expressions each containing {λki }dkvi=2

Chapter 3. Multi-edge LDPC Coded Modulation 57

which all equal R. After cancelling T and simple algebraic manipulations, the socket-

count constraint produces(T2

)linear constraints

d1c∑

iλ1ii

= d2c∑

iλ2ii

d1c∑

iλ1ii

= d3c∑

iλ3ii

......

...

d1c∑

iλ1ii

= dTc∑

iλTii

d2c∑

iλ2ii

= d3c∑

iλ3ii

......

...

dT−1c

∑iλT−1i

i= dTc

∑iλTii.

(3.36)

The channel-ratio constraint is satisfied by the inclusion of the factor 1/T

∂rkL(1,1) =∑i

Li,k

=∑i

1

T

λki /i∑j λ

kj/j

=1

T

∑i λ

ki /i∑

j λkj/j

=1

T(3.37)

and does not impose additional constraints on the optimization problem.

Lastly, we have the inequality convergence constraint (3.28) at every point of a quan-

tized version of D. Note that it is quadratic in {λki }dkvi=2.

For T edge-types, there are a total of T linear equality constraints from (3.29),(T2

)linear equality constraints from (3.36), 2T linear inequality constraints from (3.30) and

(3.31), and most importantly quadratic inequality constraints (3.28), the number of which

depends on the quantization resolution of D.

The code rate objective function is linear in {λki }dkvi=2

Chapter 3. Multi-edge LDPC Coded Modulation 58

maxλki

{L(1,1)−R(1)} = maxλki

{∑k

∑i

1

T

λki /i∑j λ

kj/j−R

}

= maxλki

{∑k

1

T− 1

Td1c∑

i λ1i /i

}

= maxλki

{1− 1

Td1c∑

i λ1i /i

}= min

λki

{1

Td1c∑

i λ1i /i

}= max

λki

{∑i

λ1i /i

}. (3.38)

The variable degree distribution for edge-type 1 was arbitrarily used to derive (3.38),

any edge-type will suffice since they were all constrained to R by (3.36).

In summary, the multi-edge LDPC coded modulation optimization problem consists of

maximizing (3.38) over parameters {λki }dkvi=2 subject to linear constraints (3.29),(3.30),(3.31),

(3.36), and quadratic constraint (3.28). We call this optimization problem the multi-edge

LDPC coded modulation optimization (ME-LCM-OPT).

Higher-order extension

For high-order modulation with T distinct bit-channels, solving ME-LCM-OPT for T

sets of variable degree distributions {λki }dkvi=2 is a valid design method. An easier method,

however, is to exploit the nested structure inherent in coded modulation. For example, in

the decomposition of the symbol channel into bit-channels in MLC, upper bit-channels

depended on lower bit-channels. We will leverage this characteristic to simplify code

design for high-order modulation where the number of edge-types is large.

Moreover, in implementation it is often desirable to use a single code at the transmit-

ter and receiver which can quickly adapt to different code rates and constellation sizes

according to varying channel conditions. Solving ME-LCM-OPT for a rate 3/4 16-QAM

code and rate 5/6 64-QAM code is likely to result in two completely different codes.

Chapter 3. Multi-edge LDPC Coded Modulation 59

Switching between these two codes in practice requires high implementation complexity.

Alternatively, if we the design the rate 5/6 code by extending the available rate 3/4 16-

QAM code, then changing code rates only involves appending or removing one additional

edge-type. The popular protograph LDPC codes [23] is one code family with such an

extension property, and has seen success in high-order modulation.

Finally, the complexity of ME-LCM-OPT increases significantly as the number of

edge-types increases. For example, for T edge-types, the number of edge-count con-

straints is(Tn

), while the number of quadratic convergence constraints is the product

of the quantization grid points for each dimension. Directly applying ME-LCM-OPT

results in an extremely large optimization problem, leading to numerical problems, long

run-times and poor results. If code design proceeds by extending lower-level codes, then

many of the constraints can be eliminated to keep the optimization complexity low. We

have observed much better convergence and results from an extension based design of

codes for high-order modulation.

To introduce the design-by-extension technique, observe the nesting of bit-channel

capacities for Gray-labelled 2n-QAM at rates (n− 1)/n for n = (4, 6, 8, 10) in Table 3.4.

Table 3.4: Bit-channel capacities for Gray-labelled 2n-QAM at rate (n− 1)/n

2n rate bit 1 bit 2 bit 3 bit 4 bit 5

16 3/4 0.661 0.831

64 5/6 0.715 0.858 0.930

256 7/8 0.733 0.867 0.933 0.967

1024 9/10 0.745 0.869 0.935 0.969 0.984

First use ME-LCM-OPT to design a rate 3/4 code for 16-QAM using 2 edge-types

for bit-channels of capacity (0.661, 0.831). The outputs are variable degree distributions

{λki }d1vi=2, {λki }

d2vi=2, and check degree assignment vector (d1c , d

2c).

Given these available code parameters, we formulate the extension-based ME-LCM-

Chapter 3. Multi-edge LDPC Coded Modulation 60

OPT by eliminating all constraints involving only λ1i or λ2i . In addition, ∇1f(I) and

∇2f(I) are now fully determined thus simplifying constraint (3.28). With these mod-

ifications, for rate 5/6, 64-QAM code design, the extension-based ME-LCM-OPT now

optimizes the code rate over {λ3i }d3vi=2 only.

Extension-based ME-LCM-OPT also resolves the issue of choosing the check degree

assignment vector when many edge-types are present. If full ME-LCM-OPT is used to

design for 3 edge-types the number of potential check degree assignment vectors for a

total check degree of dc is very large. By extending the available 16-QAM check node

assignment, the only free variable for 64-QAM code design is d3c . In a few attempts the

designer can find the best value and proceed with the optimization.

The extension to 256 and 1024-QAM follows in the same manner. At each extension

step, a new component is appended to the check degree assignment vector, while the

variable degree distribution adds a new set of nodes belonging to the highest bit-level.

Chapter 4

Results

The results of multi-edge LDPC ensembles optimized for high-order coded modulation

are presented. Sec. 4.1 provides the optimized variable degree distributions and check

node assignment vectors with their resulting ensemble thresholds. Sec. 4.2 provides

the probability of error performance of finite block-length realizations of these degree

distributions over the complex AWGN noise channel. Comparisons in both threshold

and finite-length performance are made with state-of-the-art designs whenever data is

available. The performance of our code designs match best reported results, with much

lower design and implementation complexity.

4.1 Threshold

We designed code ensembles for Gray-labelled 2n-QAM at rate (n − 1)/n where n =

(4, 6, 8, 10). The results were found by solving ME-LCM-OPT with high-order extension

using the fmincon solver in MATLAB. The resolution used to quantize D is 50 points

per dimension. The quadratic convergence constraint tolerance used in 16-QAM was

ε = 2.0 × 10−6, while in 64-QAM and higher the tolerance was ε = 4.3 × 10−6. The

maximum variable degree for all edge-types is 15. The check degree edge-type assignment

vector for 16-QAM was empirically found to be optimal at (d1c , d2c) = (9, 8). Higher-order

61

Chapter 4. Results 62

components of the assignment vector such as d3c were found by hand in a reduced search

space due to the extension procedure.

The ensemble parameters and thresholds are shown in Table 4.1. The thresholds were

evaluated using discrete density evolution [15] modified for multi-edge-type ensembles.

The “Gap (dB)” in Table. 4.1 is the gap to capacity of the ensemble threshold. Recall

from Sec. 2.1.2 that the capacity of the system under study is not the ultimate Shannon

limit of the channel. It is the capacity after constraining the channel input to a finite

constellation with Gray-labelling and uniform input distribution.

4.1.1 Discussion

At a maximum variable degree of 15, the threshold of the optimal rate 1/2 BICM ensemble

in [22] for 4-PAM has a gap to capacity of 0.199 dB. Since 16-QAM is two independent

4-PAM constellations and there is no published threshold for rate 3/4 16-QAM, we shall

use this gap to capacity as reference. In comparison, the threshold of ME-LCM-OPT

designed rate 3/4 16-QAM code ensemble is as close to capacity as the best available

code. No reference exists for other code rates or higher-order modulations. However,

note that the 64 to 1024 QAM ensemble thresholds are all within 0.214 dB of capacity,

likely constrained by the choice of maximum variable degree.

The main goal of the ME-LCM-OPT design method is to match variable node degrees

with the level of error protection required by each distinct high-order modulation bit-

channel. In Table. 4.1 the bolded entries mark the most significant degree distribution

weights for each bit-level. Recall that bit-channel quality increases from level 1 thru 5

according to Table 3.4. At the lowest quality bit-channel 1, degree distribution {λ1i }d1vi=2

contains significant weight at degree 15 variable nodes which provide the best error

protection, while balancing out their low rates by using degree 2 and 3 variable nodes.

At bit-level 2 where the bit-channel capacity at 0.831 is significantly higher than bit-level

1, the resultant {λ2i }d2vi=2 moves away from the strong error correction of degree 15 variable

Chapter 4. Results 63

Table 4.1: ME-LCM-OPT optimized ensembles for 2n-QAM rate (n− 1)/n codes

2n 16 64 256 1024

rate 3/4 5/6 7/8 9/10

dkc 9 8 7 7 7

i λ1i λ2i λ3i λ4i λ5i

2 0.1631 0.2192 – 0.5047 0.5047

3 0.3066 0.0002 0.4245 – –

4 0.0017 0.0618 0.5755 – –

5 0.0034 0.1542 – – –

6 0.0076 0.5575 – – –

7 0.0119 0.0028 – – –

8 0.0027 0.0014 – – –

9 0.0024 0.0008 – – –

10 0.0020 0.0005 – – –

11 0.0014 0.0004 – – –

12 0.0019 0.0004 – – –

13 0.0021 0.0003 – – –

14 0.0052 0.0003 – – –

15 0.4880 0.0002 – 0.4953 0.4953

σ∗ 0.3340 0.1520 0.0738 0.03638

Es/N0(dB) 9.525 16.363 22.639 28.783

Capacity (dB) 9.326 16.149 22.466 28.610

Gap (dB) 0.199 0.214 0.173 0.173

Chapter 4. Results 64

2 3 4 5 6 7 8 9 10 11 12 13 14 150

0.1

0.2

0.3

0.4

0.5

Variable Degree

Coe

ffici

ent

λ1

λ2

λ3

λ4

λ5

Figure 4.1: ME-LCM-OPT optimized ensembles for 2n-QAM rate (n− 1)/n codes.

nodes to use degree 5 and 6 variable nodes instead. This shift is even more pronounced

in bit-level 3 where only degree 3 and 4 variables nodes are used. Fig. 4.1 provides a

visual illustration of this key characteristic of ME-LCM-OPT design ensembles.

The resulting degree distributions for bit-channels 4 and 5 do not follow the “shift

towards lower degrees” pattern of the lower bit-channels. This is due to the extremely

high capacities of these bit-channels at 0.969 and 0.984. Such high input mutual infor-

mations cause the optimization solver to settle on the first feasible solution, resulting in

the use of the highest and lowest node degrees. Even though one can still evaluate the

threshold under these degree distributions using density evolution, the high fraction of

degree 2 nodes will cause high error floors to be present in finite-length realizations of

these designs. Further work on optimizing these highest bit-channels remains to be done

Chapter 4. Results 65

to make sure the code designs will be practically useful.

4.1.2 High code rate designs

To verify the capability of ME-LCM-OPT to design across code rates, code ensembles

were optimized for Gray-labelled 16-QAM with a maximum variable degree of 15, with

targeted code rates of 1/2, 2/3, 3/4, 5/6, 7/8, and 9/10. Only rates higher than 1/2 were

used because bandwidth-efficient coded modulation only works with high rate codes. The

resulting variable degree distributions and thresholds are given in Tables 4.2 and 4.3.

The mean gap to capacity across all 6 code rates is 0.216 with a maximum gap of

0.236. These values are consistent with the performance reported in [22] for a maximum

variable node of degree 15, and also with the designs in Table 4.1. The difference in

degree distributions between edge-types 1 and 2 remains, as degree 15 is nearly always

used in edge-type 1 and never used in edge-type 2. These results help to confirm the

ME-LCM-OPT technique as an effective method for designing high rate codes used in

LDPC coded modulation.

4.2 Finite-length

For finite-length realization of the code ensembles, in order to obtain good error floors

the parity-check matrices were generated using the improved progressive-edge-growth

(PEG) algorithm [41, 42] modified for multi-edge-type codes. One key modification is

to interleave variable nodes from different edge-types to their proper bit-level positions

according to bit-channel quality, before applying the PEG algorithm. Another modifica-

tion is to impose strict check degree edge counts for all edge-types, whereas check degrees

are allowed to vary in the original PEG algorithm.

Simulations to estimate probability of bit and frame errors were written in C and

compiled using GNU gcc version 4.2.1. The implemented system was shown in Fig. 2.1.

Chapter 4. Results 66

Table 4.2: ME-LCM-OPT optimized ensembles of various code rates for 16-QAM

rate 1/2 2/3 3/4

(d1c , d2c) (4, 3) (6, 6) (9, 8)

i λ1i λ2i λ1i λ2i λ1i λ2i

2 0.2448 0.3362 0.1970 0.2684 0.1631 0.2192

3 0.1572 0.3950 0.3289 0.0536 0.3066 0.0002

4 0.0663 0.0020 0.0022 0.0172 0.0017 0.0618

5 0.1224 0.0013 0.0056 0.0093 0.0034 0.1542

6 0.0048 0.0006 0.0562 0.3224 0.0076 0.5575

7 0.0043 0.0025 0.0237 0.0705 0.0119 0.0028

8 0.0061 0.2150 0.0068 0.0850 0.0027 0.0014

9 0.0012 0.0455 0.0055 0.0241 0.0024 0.0008

10 0.0012 0.0000 0.0049 0.0861 0.0020 0.0005

11 0.0022 0.0005 0.0049 0.0463 0.0014 0.0004

12 0.3124 0.0006 0.0033 0.0043 0.0019 0.0004

13 0.0050 0.0003 0.0104 0.0085 0.0021 0.0003

14 0.0212 0.0004 0.0031 0.0022 0.0052 0.0003

15 0.0509 0.0002 0.3476 0.0022 0.4880 0.0002

σ∗ 0.5310 0.3910 0.3340

Es/N0 (dB) 5.498 8.156 9.525

Capacity (dB) 5.273 7.920 9.339

Gap (dB) 0.225 0.236 0.186

Chapter 4. Results 67

Table 4.3: ME-LCM-OPT optimized ensembles of various code rates for 16-QAM (con-

tinued)

rate 5/6 7/8 9/10

(d1c , d2c) (13, 13) (18, 18) (23, 23)

i λ1i λ2i λ1i λ2i λ1i λ2i

2 0.1549 0.1404 0.0958 0.2717 0.1243 0.0005

3 0.3336 0.1677 0.3938 0.0000 0.3306 0.3332

4 0.0001 0.0010 0.0133 0.0000 0.0008 0.0471

5 0.0002 0.0037 0.0001 0.0004 0.0016 0.0746

6 0.0003 0.2293 0.0000 0.0142 0.0088 0.1724

7 0.0008 0.4527 0.0000 0.4416 0.0629 0.3076

8 0.0024 0.0024 0.0000 0.0000 0.0104 0.0378

9 0.1263 0.0009 0.0000 0.0004 0.0403 0.0109

10 0.0208 0.0005 0.0001 0.0002 0.0137 0.0074

11 0.0307 0.0004 0.1347 0.0216 0.0000 0.0031

12 0.0105 0.0003 0.0000 0.0002 0.0154 0.0001

13 0.0179 0.0002 0.3620 0.2481 0.0088 0.0026

14 0.0156 0.0002 0.0000 0.0014 0.0247 0.0002

15 0.2857 0.0002 0.0001 0.0001 0.3576 0.0026

σ∗ 0.2820 0.2580 0.2410

Es/N0 (dB) 10.995 11.768 12.360

Capacity (dB) 10.767 11.579 12.128

Gap (dB) 0.228 0.189 0.232

Chapter 4. Results 68

IID channel adapters described in Sec. 2.1.1 were included in the simulation. Since the

iid channel adapters result in output symmetric bit-channels, we only transmitted the

all-zeros codeword to estimate the equivalent error performance for all codewords.

4.2.1 Rate 3/4 Gray-labelled 16-QAM

Figs. 4.2, 4.3 and 4.4 show the probability of bit and frame errors (BER, FER) with

respect to the per-bit signal-to-noise ratio Eb/N0 in dB for codeword lengths of 4096 (4K),

8192 (8K) and 16384 (16K) respectively. The code ensemble used is the rate 3/4 16-QAM

code with a threshold of 9.525 dB Es/N0 or 4.754 dB Eb/N0. The ensemble threshold is

indicated by the black vertical line in the figures. The red vertical line indicates capacity.

The coloured dots indicate the ±3σ error bars on the BER and FER curves, with some

of the lower error bars missing due to the logarithmic scale used.

The Eb/N0 axis of the three figures are fixed for comparison. Choosing an arbitrary

point of reference on the BER curve, for example the 10−6 point, we see that the SNRs

required are approximately 5.78 dB, 5.49 dB, and 5.27 dB for 4K, 8K and 16K. The 10−6

SNR for 16K is 0.52 dB from threshold and 0.72 dB from capacity.

The figures show an FER error floor at approximately 10−4. A quick calculation gives

the number of error bits per erroneous frame to be 20, 17 and 10 bits for 4K, 8K and 16K.

An efficient concatenated code solution can reduce the error floor to negligible levels and

has been utilized in the DVB-S2 standard [8]. For example, we can use the (255, 239)

rate 0.937 double error correcting Bose-Chaudhuri-Hocquenghem (BCH) code [43] with

an interleaver that disperses the 10 bit errors among 16384 codeword bits to ensure less

than 2 errors will occur per BCH codeword. If the slight rate loss is not acceptable, then

the inner LDPC code can be designed with the proper rate compensation to obtain the

correct concatenated code rate.

Chapter 4. Results 69

4.4 4.6 4.8 5 5.2 5.4 5.6 5.8 610

−8

10−7

10−6

10−5

10−4

10−3

10−2

10−1

100

Eb/N

0 (dB)

BE

R/F

ER

BERFER

Figure 4.2: Probability of bit and errors for n = 4096 rate 3/4 code and Gray-labelled

16-QAM.

Chapter 4. Results 70

4.4 4.6 4.8 5 5.2 5.4 5.6 5.8 610

−8

10−7

10−6

10−5

10−4

10−3

10−2

10−1

100

Eb/N

0 (dB)

BE

R/F

ER

BERFER

Figure 4.3: Probability of bit and errors for n = 8192 rate 3/4 code and Gray-labelled

16-QAM.

Chapter 4. Results 71

4.4 4.6 4.8 5 5.2 5.4 5.6 5.8 610

−8

10−7

10−6

10−5

10−4

10−3

10−2

10−1

100

Eb/N

0 (dB)

BE

R/F

ER

BERFER

Figure 4.4: Probability of bit and errors for n = 16384 rate 3/4 code and Gray-labelled

16-QAM.

Chapter 4. Results 72

4.2.2 Rate 1/2 Gray-labelled 16-QAM

Fig. 4.5 shows the BER and FER results of the rate 1/2 16-QAM code given in Table

4.2, at a codeword length of 16200 bits. At an BER of 10−6 the code requires an Eb/N0

of 3.27 dB, which is 0.78 dB from threshold and 1.0 dB from capacity. An FER error

floor is present at an FER of 5 × 10−5. In comparison, as reported in [24], a code with

the same rate and block-length requires an Eb/N0 of 3.17 dB, a difference of 0.1 dB.

Furthermore, the reference [22] code of rate 1/2 at a codeword length of 16200 has the

identical performance as our code [24]. These comparisons indicate that the ME-LCM-

OPT design procedure produces codes that are equivalent to best known codes not only

in terms of threshold, but also in finite-length performance.

4.2.3 Fixed BER performance comparison

For a final summary comparison, we plot the Es/N0 (dB) required by our code designs

to achieve a BER of 10−5 with several published results at the same spectral efficiency.

Efforts were made to select references that provide the fairest possible comparison in

terms of block-length and maximum variable degree. Fig. 4.6(a) plots the Es/N0 required

at a spectral efficiency of 3 bits per channel use for Gray-labelled 16-QAM, while Fig.

4.6(b) shows the comparison at 2 bits per channel use. Blue markers are our ME-LCM-

OPT code designs and the black markers are results from published literature.

Legends for the labelled data points in Fig. 4.6 are given in Table. 4.4. Note that

some descriptions use the number of informations k to define the code, instead of the

overall block-length n.

In Fig. 4.6(a) we compare the n = 4096 (data point 1) and k = 4096 (data point 4)

codes of our design with the AR4JA k = 4096 (data point 5) code. We see that our code

achieves the required BER at the shorter block length of 4096. At the same information

bit length, our code provides a 0.1 dB improvement. Part of this difference is because

Chapter 4. Results 73

2.2 2.4 2.6 2.8 3 3.2 3.4 3.6 3.810

−9

10−8

10−7

10−6

10−5

10−4

10−3

10−2

10−1

100

Eb/N

0 (dB)

BE

R/F

ER

BERFER

Figure 4.5: Probability of bit and errors for n = 16200 rate 1/2 code and Gray-labelled

16-QAM.

Chapter 4. Results 74

8 8.5 9 9.5 10 10.5 112.9

2.92

2.94

2.96

2.98

3

3.02

3.04

3.06

3.08

3.1

Es/N

0 (dB)

bits

/cha

nnel

use

123 4

5

16−QAM Gray capacityShannon Limit

(a) 3 bits per channel use

4 4.5 5 5.5 6 6.5 71.9

1.92

1.94

1.96

1.98

2

2.02

2.04

2.06

2.08

2.1

Es/N

0 (dB)

bits

/cha

nnel

use

1 2

3 4 5

16−QAM Gray capacityShannon Limit

(b) 2 bits per channel use

Figure 4.6: Comparison of Es/N0 required to achieve BER of 10−5 at different spectral

efficiencies for Gray-labelled 16-QAM.

Chapter 4. Results 75

Table 4.4: Legend for data points in Fig. 4.6

(a) (b)

1 n = 4096 1 n = 16200

2 n = 8192 2 k = 4096

3 n = 16384 3 n = 20000 [22]

4 k = 4096 4 n = 10000 [25]

5 k = 4096 AR4JA [23] 5 k = 4096 AR4JA [23]

AR4JA uses as maximum variable degree of 6 while our design uses degree 15. However,

given that we did not use puncturing or degree 1 variable nodes and the consensus of

AR4JA as one of the best performing codes for high-order modulation, this improvement

is significant.

In Fig. 4.6(b), comparing the n = 16200 (data point 1) code of our design and

n = 20000 (data point 3) code from [22] we see a performance difference of 0.1 dB. This

is not surprising given the difference in block lengths. The small magnitude of the gap

supports our conclusion that our code design matches the best available designs in finite-

length performance. A second comparison can be made between the n = 8192 (data point

2) code of our design and the k = 4096 AR4JA code. Again, we see an improvement of

our code over AR4JA. Finally, note that the last reference (data point 4), a n = 10000

eIRA code from [25] designed using multi-edge-type concepts, falls between our codes

of block length 16200 and 8192. In developing our ME-LCM-OPT technique, we have

successfully unified and generalized over several previous attempts at designing LDPC

coded modulation using the multi-edge-type framework [24,25,39].

Chapter 5

Conclusion

In this thesis, we identified a point of improvement to the currently popular LDPC-

BICM scheme for bandwidth-efficient high-order coded modulation. We sought to match

bit-channel quality differences with the inherent differences in error protection in LDPC

codeword bits. A specific multi-edge-type parameterization was developed in order to

impose structure to reduce design complexity while retaining the key property of incor-

porating different bit-channel output densities into the code ensemble.

A new analysis technique using the multi-dimensional EXIT vector field was devel-

oped as an efficient and accurate way of determining the threshold of the multi-edge

parameterization for coded modulation. In order to apply the technique to code design,

a new decoding convergence condition was derived. Using the convergence condition,

the ME-LCM-OPT problem was setup in order to optimize the code ensembles for the

highest code rate at a given channel noise power. We designed several code ensembles

for 16 to 1024 QAM at different code rates, with thresholds as close to capacity as best

available results. Finite-length probability of error performance also matched current

best, with lower design and implementation complexity.

In conclusion, this thesis provided one of the first methods of LDPC code design

for high-order modulation without decomposing the symbol channel into individual bit-

76

Chapter 5. Conclusion 77

channels. In the process, an innovative analysis and design technique for multi-edge-type

ensembles based on EXIT functions was developed, which was shown to be efficient and

accurate when compared to density evolution.

We believe that the multi-edge EXIT-based optimization technique is not limited to

high-order modulation code design, but can also be applied to any application where a

single LDPC code is desired for several different channel output densities. Seeking out

such applications and applying the technique to code design is one of the major topics

of future work. Another possible direction is to extend the optimization procedure to a

variable-check iterative optimization process. While one of the two degree distributions

is being optimized, the other is assumed to be fixed. The roles are reversed in the next

optimization iteration. In doing so, the concentration to a single check degree edge-type

assignment used in this thesis may be relaxed, perhaps with improved results. Lastly,

the well known protograph-based coded modulation scheme using AR4JA protograph

codes [23] has shown excellent performance likely due to the presence of punctured and

degree 1 variable nodes. Future work to include both into the code ensemble may also

provide performance gains.

Bibliography

[1] G. Ungerboeck, “Channel coding with multilevel/phase signals,” IEEE Trans. Inf.

Theory, vol. 28, no. 1, pp. 55 – 67, Jan. 1982.

[2] ——, “Trellis-coded modulation with redundant signal sets Part I: Introduction,”

IEEE Commun. Mag., vol. 25, no. 2, pp. 5–11, Feb. 1987.

[3] C. Berrou, A. Glavieux, and P. Thitimajshima, “Near Shannon limit error-correcting

coding and decoding: Turbo-codes,” in IEEE International Conf. on Comm., vol. 2,

Geneva, Switzerland, May 1993, pp. 1064–1070.

[4] R. Gallager, “Low-density parity-check codes,” IRE Trans. Inf. Theory, vol. 8, no. 1,

pp. 21–28, Jan. 1962.

[5] G. Caire, G. Taricco, and E. Biglieri, “Bit-interleaved coded modulation,” IEEE

Trans. Inf. Theory, vol. 44, no. 3, pp. 927–946, May 1998.

[6] H. Imai and S. Hirakawa, “A new multilevel coding method using error-correcting

codes,” IEEE Trans. Inf. Theory, vol. 23, no. 3, pp. 371–377, May 1977.

[7] U. Wachsmann, R. Fischer, and J. B. Huber, “Multilevel codes: theoretical concepts

and practical design rules,” IEEE Trans. Inf. Theory, vol. 45, no. 5, pp. 1361–1391,

Jul. 1999.

[8] Digital Video Broadcasting (DVB); Second generation framing structure, channel

coding and modulation systems for broadcasting, interactive services, news gathering

78

Bibliography 79

and other broadband satellite applications, European Telecommunications Standards

Institute Std., Aug. 2009.

[9] Digital Video Broadcasting (DVB); Frame structure channel coding and modulation

for a second generation digital transmission system for cable systems, European

Telecommunications Standards Institute Std., Jun. 2010.

[10] IEEE Standard for local and metropolitan area networks, Part 16: Air interface for

broadband wireless access systems, Institute of Electrical and Electronic Engineers

Std., May 2009.

[11] Low density parity check codes for use in near-earth and deep space applications,

The Consultive Commitee for Space Data Systems Std., Sep. 2007.

[12] H. Jin, A. Khandekar, and R. McEliece, “Irregular repeat-accumulate codes,” in

Proc. 2nd Int. Symp. Turbo Codes Related Topics, Brest, France, Sep. 2000.

[13] R. Tanner, “A recursive approach to low complexity codes,” IEEE Trans. Inf. The-

ory, vol. 27, no. 5, pp. 533 – 547, Sep. 1981.

[14] T. J. Richardson, M. A. Shokrollahi, and R. L. Urbanke, “Design of capacity-

approaching irregular low-density parity-check codes,” IEEE Trans. Inf. Theory,

vol. 47, no. 2, pp. 619–637, Feb. 2001.

[15] S.-Y. Chung, G. D. Forney, Jr., T. J. Richardson, and R. L. Urbanke, “On the

design of low-density parity-check codes within 0.0045 dB of the Shannon limit,”

IEEE Commun. Lett., vol. 5, no. 2, pp. 58–60, Feb. 2001.

[16] F. Kschischang, B. Frey, and H.-A. Loeliger, “Factor graphs and the sum-product

algorithm,” IEEE Trans. Inf. Theory, vol. 47, no. 2, pp. 498–519, Feb. 2001.

Bibliography 80

[17] M. Luby, M. Mitzenmacher, M. Shokrollahi, and D. Spielman, “Improved low-

density parity-check codes using irregular graphs,” IEEE Trans. Inf. Theory, vol. 47,

no. 2, pp. 585–598, Feb. 2001.

[18] T. Richardson and R. Urbanke, “The capacity of low-density parity-check codes

under message-passing decoding,” IEEE Trans. Inf. Theory, vol. 47, no. 2, pp. 599–

618, Feb. 2001.

[19] Y. Li and W. E. Ryan, “Bit-reliability mapping in LDPC-coded modulation sys-

tems,” IEEE Commun. Lett., vol. 9, no. 1, pp. 1–3, Jan. 2005.

[20] R. Maddock and A. Banihashemi, “Reliability-based coded modulation with low-

density parity-check codes,” IEEE Trans. Commun., vol. 54, no. 3, pp. 403–406,

Mar. 2006.

[21] J. Lei and W. Gao, “Matching graph connectivity of LDPC codes to high-order mod-

ulation by bit interleaving,” in 46th Annual Allerton Conference on Communication,

Control and Computing, Sep. 2008, pp. 1059–1064.

[22] J. Hou, P. H. Siegel, L. B. Milstein, and H. D. Pfister, “Capacity-approaching

bandwidth-efficient coded modulation schemes based on low-density parity-check

codes,” IEEE Trans. Inf. Theory, vol. 49, no. 9, pp. 2141–2155, Sep. 2003.

[23] D. Divsalar and C. Jones, “Protograph based low error floor LDPC coded mod-

ulation,” in IEEE Mil. Comm. Conf., Atlantic City, New Jersey, Oct. 2005, pp.

378–385.

[24] H. Sankar, N. Sindhushayana, and K. R. Narayanan, “Design of low-density parity

check LDPC codes for high order constellations,” in IEEE Global Telecom. Conf.,

vol. 5, Nov. 2004, pp. 3113–3117.

Bibliography 81

[25] G. Durisi, L. Dinoi, and S. Benedetto, “eIRA codes for coded modulation systems,”

in IEEE International Conf. on Comm., vol. 3, Istanbul, Turkey, Jun. 2006, pp.

1125–1130.

[26] S. Sandberg and N. Von Deetzen, “Design of bandwidth-efficient unequal error pro-

tection LDPC codes,” IEEE Trans. Commun., vol. 58, no. 3, pp. 802–811, Mar.

2010.

[27] L. Yan and W. E. Ryan, “Design of LDPC-coded modulation schemes,” un-

published. [Online]. Available: http://www2.engr.arizona.edu/∼ryan/publications/

turbo-symp-03.pdf

[28] M. Ardakani, T. Esmailian, and F. R. Kschischang, “Near-capacity coding in multi-

carrier modulation systems,” IEEE Trans. Commun., vol. 52, no. 11, pp. 1880–1889,

Nov. 2004.

[29] T. Richardson and R. Urbanke, Modern Coding Theory. Cambridge University

Press, 2008.

[30] G. D. Forney, Jr., “Trellis shaping,” IEEE Trans. Inf. Theory, vol. 38, no. 2, pp.

281–300, Mar. 1992.

[31] F. R. Kschischang and S. Pasupathy, “Optimal nonuniform signaling for Gaussian

channels,” IEEE Trans. Inf. Theory, vol. 39, no. 3, pp. 913–929, Mar. 1993.

[32] R. Storn and K. Price, “Differential evolution, a simple and efficient heuristic for

global optimization over continuous spaces,” Journal of Global Optimization, vol. 11,

pp. 341–359, 1997.

[33] A. Papoulis and S. U. Pillai, Probability, random variables and stochastic processes,

4th ed. McGraw-Hill Higher Education, 2002.

Bibliography 82

[34] S. Y. Chung, T. Richardson, and R. Urbanke, “Analysis of sum-product decoding of

low-density parity-check codes using a Gaussian approximation,” IEEE Trans. Inf.

Theory, vol. 47, no. 2, pp. 657–670, Feb. 2001.

[35] M. Ardakani and F. R. Kschischang, “A more accurate one-dimensional analysis and

design of irregular LDPC codes,” IEEE Trans. Commun., vol. 52, no. 12, pp. 2106

– 2114, Dec. 2004.

[36] S. ten Brink, G. Kramer, and A. Ashikhmin, “Design of low-density parity-check

codes for modulation and detection,” IEEE Trans. Commun., vol. 52, no. 4, pp.

670–678, Apr. 2004.

[37] A. Ashikhmin, G. Kramer, and S. ten Brink, “Extrinsic information transfer func-

tions: model and erasure channel properties,” IEEE Trans. Inf. Theory, vol. 50,

no. 11, pp. 2657–2673, Nov. 2004.

[38] E. W. Weisstein. (2011, May) Partition function P. MathWorld - A Wolfram Web

Resource. [Online]. Available: http://mathworld.wolfram.com/PartitionFunctionP.

html

[39] H. V. Neto, W. Henkel, and V. C. da Rocha, Jr., “Multi-edge type unequal error

protection LDPC codes,” Jan. 2011, unpublished. ArXiv:1101.5985v1. [Online].

Available: http://arxiv.org/abs/1101.5985v1

[40] S. ten Brink, “Convergence behaviour of iteratively decoded parallel concatenated

codes,” IEEE Trans. Commun., vol. 49, no. 10, pp. 1727–1737, Oct. 2001.

[41] H. Xiao and A. H. Banihashemi, “Improved progressive-edge-growth (PEG) con-

struction of irregular LDPC codes,” Communications Letters, IEEE, vol. 8, no. 12,

pp. 715 – 717, Dec. 2004.

Bibliography 83

[42] X. Y. Hu, E. Eleftheriou, and D. M. Arnold, “Regular and irregular progressive

edge-growth Tanner graphs,” IEEE Trans. Inf. Theory, vol. 51, no. 1, pp. 386–398,

Jan. 2005.

[43] J. Proakis and M. Salehi, Digital Communications, 5th ed. McGraw-Hill Sci-

ence/Engineering/Math, 2007.