Comparison of the Achievable Rates in OFDM and Single Carrier Modulation with I.I.D. Inputs

Similar documents
On the Shamai-Laroia Approximation for the Information Rate of the ISI Channel

Lower Bounds and Approximations for the Information Rate of the ISI Channel

NUMERICAL COMPUTATION OF THE CAPACITY OF CONTINUOUS MEMORYLESS CHANNELS

Communication Limits with Low Precision Analog-to-Digital Conversion at the Receiver

Single-Carrier Block Transmission With Frequency-Domain Equalisation

EE-597 Notes Quantization

Approximately achieving the feedback interference channel capacity with point-to-point codes

Appendix B Information theory from first principles

a) Find the compact (i.e. smallest) basis set required to ensure sufficient statistics.

THE inverse tangent function is an elementary mathematical

Interactions of Information Theory and Estimation in Single- and Multi-user Communications

Channel Coding 1. Sportturm (SpT), Room: C3165

Decoupling of CDMA Multiuser Detection via the Replica Method

4184 IEEE TRANSACTIONS ON INFORMATION THEORY, VOL. 51, NO. 12, DECEMBER Pranav Dayal, Member, IEEE, and Mahesh K. Varanasi, Senior Member, IEEE

Decision-Point Signal to Noise Ratio (SNR)

392D: Coding for the AWGN Channel Wednesday, January 24, 2007 Stanford, Winter 2007 Handout #6. Problem Set 2 Solutions

Fast Near-Optimal Energy Allocation for Multimedia Loading on Multicarrier Systems

Minimum BER Linear Transceivers for Block. Communication Systems. Lecturer: Tom Luo

Trellis Coded Modulation

EE6604 Personal & Mobile Communications. Week 15. OFDM on AWGN and ISI Channels

Chapter 9 Fundamental Limits in Information Theory

Nearest Neighbor Decoding in MIMO Block-Fading Channels With Imperfect CSIR

An Alternative Proof of Channel Polarization for Channels with Arbitrary Input Alphabets

MMSE Dimension. snr. 1 We use the following asymptotic notation: f(x) = O (g(x)) if and only

Analysis of Receiver Quantization in Wireless Communication Systems

Chapter 6. Nonlinear Equations. 6.1 The Problem of Nonlinear Root-finding. 6.2 Rate of Convergence

Direct-Sequence Spread-Spectrum

ADDITIVE noise is most often represented by a fixed

Shannon meets Wiener II: On MMSE estimation in successive decoding schemes

Interactive Interference Alignment

MULTICARRIER code-division multiple access (MC-

ELEC546 Review of Information Theory

Supplementary Figure 1: Scheme of the RFT. (a) At first, we separate two quadratures of the field (denoted by and ); (b) then, each quadrature

Performance of DS-CDMA Systems With Optimal Hard-Decision Parallel Interference Cancellation

Information, Estimation, and Lookahead in the Gaussian channel

Degrees-of-Freedom for the 4-User SISO Interference Channel with Improper Signaling

Digital Band-pass Modulation PROF. MICHAEL TSAI 2011/11/10

Approximate Capacity of Fast Fading Interference Channels with no CSIT

arxiv:cs/ v1 [cs.it] 11 Sep 2006

Lecture 12. Block Diagram

Diversity Multiplexing Tradeoff in ISI Channels Leonard H. Grokop, Member, IEEE, and David N. C. Tse, Senior Member, IEEE

Problem Set 3 Due Oct, 5

Performance of small signal sets

Wideband Fading Channel Capacity with Training and Partial Feedback

On the Performance of. Golden Space-Time Trellis Coded Modulation over MIMO Block Fading Channels

M.S. Project Report. Efficient Failure Rate Prediction for SRAM Cells via Gibbs Sampling. Yamei Feng 12/15/2011

Revision of Lecture 5

Design of MMSE Multiuser Detectors using Random Matrix Techniques

Upper Bounds to Error Probability with Feedback

Taylor Series and Asymptotic Expansions

Binary Transmissions over Additive Gaussian Noise: A Closed-Form Expression for the Channel Capacity 1

ECE 587 / STA 563: Lecture 5 Lossless Compression

6196 IEEE TRANSACTIONS ON INFORMATION THEORY, VOL. 57, NO. 9, SEPTEMBER 2011

ECE 587 / STA 563: Lecture 5 Lossless Compression

On the Required Accuracy of Transmitter Channel State Information in Multiple Antenna Broadcast Channels

Asynchronous Physical-layer Network Coding

Computing Probability of Symbol Error

Lecture 4 Capacity of Wireless Channels

Lecture 4. Capacity of Fading Channels

Determining the Optimal Decision Delay Parameter for a Linear Equalizer

Data Detection for Controlled ISI. h(nt) = 1 for n=0,1 and zero otherwise.

Capacity of the Discrete-Time AWGN Channel Under Output Quantization

Computation of Bit-Error Rate of Coherent and Non-Coherent Detection M-Ary PSK With Gray Code in BFWA Systems

The Capacity of Finite Abelian Group Codes Over Symmetric Memoryless Channels Giacomo Como and Fabio Fagnani

IEEE TRANSACTIONS ON INFORMATION THEORY, VOL. 52, NO. 2, FEBRUARY Uplink Downlink Duality Via Minimax Duality. Wei Yu, Member, IEEE (1) (2)

Adaptive Bit-Interleaved Coded OFDM over Time-Varying Channels

ANALYSIS OF A PARTIAL DECORRELATOR IN A MULTI-CELL DS/CDMA SYSTEM

Diversity Performance of a Practical Non-Coherent Detect-and-Forward Receiver

Es e j4φ +4N n. 16 KE s /N 0. σ 2ˆφ4 1 γ s. p(φ e )= exp 1 ( 2πσ φ b cos N 2 φ e 0

On the complexity of maximizing the minimum Shannon capacity in wireless networks by joint channel assignment and power allocation

Schur-convexity of the Symbol Error Rate in Correlated MIMO Systems with Precoding and Space-time Coding

APROBLEM often encountered in transmitter design is that

Vector Channel Capacity with Quantized Feedback

PSK bit mappings with good minimax error probability

Capacity Penalty due to Ideal Zero-Forcing Decision-Feedback Equalization

Practical Polar Code Construction Using Generalised Generator Matrices

ECE 564/645 - Digital Communications, Spring 2018 Homework #2 Due: March 19 (In Lecture)

Particle Methods as Message Passing

RADIO SYSTEMS ETIN15. Lecture no: Equalization. Ove Edfors, Department of Electrical and Information Technology

LOW-density parity-check (LDPC) codes were invented

Notes 3: Stochastic channels and noisy coding theorem bound. 1 Model of information communication and noisy channel

Bounds on Mutual Information for Simple Codes Using Information Combining

IN this paper, we consider the capacity of sticky channels, a

ebay/google short course: Problem set 2

Chapter 2 Piecewise Bijective Functions and Continuous Inputs

Lecture 6 Channel Coding over Continuous Channels

Lecture 4 Noisy Channel Coding

Signals and Spectra - Review

Performance Analysis and Interleaver Structure Optimization for Short-Frame BICM-OFDM Systems

On the Concentration of the Crest Factor for OFDM Signals

Modulation & Coding for the Gaussian Channel

On the Low-SNR Capacity of Phase-Shift Keying with Hard-Decision Detection

Lower Bounds on the Graphical Complexity of Finite-Length LDPC Codes

WE study the capacity of peak-power limited, single-antenna,

Economics 205 Exercises

2318 IEEE TRANSACTIONS ON SIGNAL PROCESSING, VOL. 54, NO. 6, JUNE Mai Vu, Student Member, IEEE, and Arogyaswami Paulraj, Fellow, IEEE

Dirty Paper Coding vs. TDMA for MIMO Broadcast Channels

CDMA Systems in Fading Channels: Admissibility, Network Capacity, and Power Control

EE6604 Personal & Mobile Communications. Week 13. Multi-antenna Techniques

DS-GA 1002 Lecture notes 0 Fall Linear Algebra. These notes provide a review of basic concepts in linear algebra.

Transcription:

Comparison of the Achievable Rates in OFDM and Single Carrier Modulation with I.I.D. Inputs Yair Carmon, Shlomo Shamai and Tsachy Weissman arxiv:36.578v4 [cs.it] Nov 4 Abstract We compare the maimum achievable rates in single-carrier and OFDM modulation schemes, under the practical assumptions of i.i.d. finite alphabet inputs and linear ISI with additive Gaussian noise. We show that the Shamai- Laroia approimation serves as a bridge between the two rates: while it is well known that this approimation is often a lower bound on the single-carrier achievable rate, it is revealed to also essentially upper bound the OFDM achievable rate. We apply Information-Estimation relations in order to rigorously establish this result for both general input distributions and to sharpen it for commonly used PAM and QAM constellations. To this end, novel bounds on MMSE estimation of PAM inputs to a scalar Gaussian channel are derived, which may be of general interest. Our results show that, under reasonable assumptions, optimal single-carrier schemes may offer spectral efficiency significantly superior to that of OFDM, motivating further research of such systems. I. INTRODUCTION Intersymbol interference ISI) is a ubiquitous impairment in communication and data storage media []. Techniques of information transmission over ISI channels can be roughly divided into two types: single-carrier SC) modulation and multi-carrier modulation. In SC modulation, symbols are transmitted at a rate approimately equal to the available bandwidth, and are distorted by ISI. The distortion can be compensated for at the receiver using different equalization techniques such as maimum-likelihood sequence estimation [] and linear MMSE estimation, possibly with decision-feedback [3]. Equalization can also be combined with decoding, resulting in various iterative schemes [4]. In multi-carrier modulation, the available bandwidth is divided among several lower symbol-rate signals, each with a different carrier frequency, often referred to as subcarriers. Orthogonal frequency-division multipleing OFDM) [5], [6] is the most important multi-carrier modulation technique. In OFDM, the frequency spacing between subcarriers is chosen so that they are orthogonal on every signaling interval. Additionally, a portion of the end of each interval is copied to its beginning and is commonly referred to as the cyclic prefi. These two modifications allow for low-compleity optimal equalization using the FFT algorithm, which is the primary advantage of OFDM. Technion, Israel Institute of Technology. Emails: yairc@t.technion.ac.il, sshlomo@ee.technion.ac.il Stanford University. Email: tsachy@stanford.edu This paper was presented at the 4 International Zurich Seminar on Communications.

Today, OFDM is the predominant modulation technique in high-bandwidth communications over channels with significant ISI, and is featured in a large number of standards, including DSL [7], WiFi [8], WiMAX [9], DVB-T [] and the LTE downlink []. However, OFDM waveforms suffer from a higher peak to average power ratio PAPR) than SC waveforms. Due to their lower PAPR, and due to the introduction of efficient frequency-domain decision feedback equalization techniques [], SC schemes have become a viable alternative to OFDM in certain settings. In particular, small, cheap and power-efficient amplifiers require low PAPR input due to their limited dynamic range, making SC desirable when they are used. Two such eamples are the uplink air-interface of LTE, where a scheme called SC-FDMA is used [3], and the new 8.ad specification, which includes a SC option [4]. The purpose of this paper is to compare the maimum achievable rates of reliable communication in OFDM and SC modulations, when optimal equalization and channel coding is assumed. Since optimal OFDM equalization is trivial, modern OFDM systems are able to approach this maimum theoretical rate using advanced coding schemes such as turbo codes or LDPC [5]. In contrast, optimal equalization and decoding cannot be decoupled in SC schemes, and instead must be approimated by iterative turbo equalization techniques [4]. Such techniques incur a high computational cost which currently renders them infeasible in practice. Nonetheless, with ever-growing computation resources and steadily improving iterative receivers c.f. [6]), SC schemes able to approach the maimum achievable rate might soon become feasible. Therefore, comparison of the two maimum achievable rates is of practical as well as theoretical interest. When only average power constraints apply on the channel input, it is well known that Gaussian signaling achieves the same maimum rate in both OFDM and SC schemes [7]. However, we impose two practical restrictions, which rule out the classical Gaussian solution. The first restriction is that the channel inputs take values in a certain fied finite alphabet constellation), as is always the case in practice. The second restriction is that the inputs are i.i.d., or more precisely that an i.i.d. random coding distribution is used. This restriction is justified as long as we limit ourselves to channel coding schemes that were designed for memoryless channels, as is common in practice. For OFDM the latter restriction is taken to mean that all subcarriers use the same distribution this will simplify our calculations and it also accurately models wireless ISI channels, which most often change too rapidly for constellation loading to become practical. Under these constraints, there is no closed-form epression for the achievable rate of SC modulation. However, an epression introduced by Shamai and Laroia [8] is known to tightly approimate the SC achievable rate in virtually all scenarios. The Shamai-Laroia approimation is intimately connected to the performance of decision-feedback equalization of SC signals. In this paper, we show that in the sense of achievable rate and under the above-mentioned restrictions, SC modulation will often offer performance superior or equal to that of OFDM. In particular, we prove that for BPSK and QPSK inputs, the OFDM achievable rate is lower than the Shamai-Laroia approimation regardless of the ISI channel. A numerical study indicates this holds also for 4-PAM, 8-PSK, 6-QAM and 3-QAM inputs. For general finite-alphabet input distributions, we prove the same result in the low- and high-snr regimes. For PAM and square QAM inputs, we find an SNR threshold above which the high-snr regime is in effect. For square QAM inputs of order 56 and above, this threshold corresponds to a symbol error rate of more than 5% and is therefore reasonably

3 low. We provide an eact characterization of the maimum advantage OFDM may offer over the Shamai-Laroia approimation. Numerical evaluation shows that this advantage is very small for QAM constellations of orders up to 496 less than. bit). Complementing this result, we show that the advantage of the Shamai-Laroia approimation over the OFDM rate becomes arbitrarily large for some of ISI channels. Thus, SC schemes may offer a significant performance gain over OFDM in certain cases, but not vice versa. Finally, we study continuous uniform input, which is limiting case of increasingly high-order QAM inputs. We show that unlike the finite-alphabet cases, under uniform input the Shamai-Laroia approimation cannot significantly eceed the OFDM achievable rate. This indicates that the advantage of SC over OFDM will become small if a sufficiently dense input constellation is chosen. However, such inputs are not necessarily feasible. We provide a detailed discussion on practical scenarios in which SC is epected to offer a considerable performance gain over OFDM. Our results stem from the concavity properties of the input-output mutual information in a scalar Gaussian channel, as a function of a rescaled SNR variable that we call the log-snr. In order to investigate these properties, we make etensive use of Information-Estimation results that link derivatives of the mutual information function and estimation-theoretic quantities [9] []. We derive new bounds on MMSE estimation of PAM inputs to an additive Gaussian channel in the high-snr regime. Besides their use in proving some of our main results, we believe them to be of general interest. There is a considerable amount of literature that deals with comparison between SC and OFDM modulations. It is mostly concerned with comparison of specific schemes and quantities such as PAPR and bit or frame error rates c.f. [] [7]). Some works also compare fundamental limits. In [8] [3] achievable rates are considered, but Gaussian inputs are assumed in order to model adaptive constellation loading, leading to the epected conclusion that both methods offer the same rates. In [3] achievable rates are compared via simulation for binary input and two-taps ISI channels. The authors report that in these settings SC is superior, but make no general or theoretically-backed claims. In [3] it is conjectured that the SC achievable rate is always higher than the OFDM rate. This conjecture is supported by numerical evidence, but no theoretical analysis is performed. In [33] and [34] the cut-off rate is compared analytically, and the SC rate is shown to eceed the OFDM rate in several scenarios. In a recent report [35] the authors independently found that concavity with respect to log-snr yields an inequality between the OFDM achievable rate and the Shamai-Laroia approimation. Based on numerical evidence, they argue that concavity holds for QPSK and 6-QAM inputs and that while it does not always hold for higher order constellations, the maimum difference in favor of OFDM is small. However, the majority of our results remain eclusive to this work, including the analytic proofs of concavity, the study of the concave envelope and the application of Information-Estimation tools. The rest of this paper is organized as follows. Section II formulates the problem and presents the main results, briefly discussing their etension when linear precoding is allowed. Section III introduces the log-snr scale and studies the concavity of the mutual information function with respect to it, culminating in a proof to our main result for general finite-alphabet inputs. Section IV focuses on the nonlinear MMSE estimation of PAM random variables corrupted by additive Gaussian noise. In this section, a novel pointwise result is presented, bounding the conditional variance of the channel input given an observation of the channel output. This bound is then applied to

4 derive a tight high-snr characterization of the MMSE function and its derivative. Section V uses insights from the two previous sections in order to prove our results pertaining PAM and square QAM inputs, as well as continuous uniform input. Section VI contains an in-depth discussion on the differences between the SC and OFDM achievable rates likely to occur in practice, and the implications of increasing the constellation order. Section VII concludes this paper. II. PRELIMINARIES AND MAIN RESULTS We consider a comple-valued, discrete-time ISI channel model, L y k = h i k i + n k, ) i= where is the channel input sequence, h L are arbitrary comple-valued ISI taps and n is an i.i.d. standard comple Gaussian sequence independent of the input. Here L denotes the length of the channel impulse response. Let H θ) = L k= h ke jkθ be the ISI channel transfer function. We assume throughout the paper that the input sequence has zero mean and unit average power, i.e. E i =. Since the input and noise are both normalized to unit variance, the quantity L k= h k = π signal-to-noise ratio SNR). π π Hθ) dθ, which is not normalized to unity, epresses the input A. Single carrier modulation model In our model for single-carrier modulation, the channel input sequence is assumed to be i.i.d., zero-mean and unit power, with every input symbol drawn from a finite comple-valued alphabet also known as a signal constellation. Conventional constellations are composed of m uniformly spaced symbols, each representing m data bits. Commonly used constellations include BPSK, QPSK, 8-PSK, 6-QAM and 64-QAM []. Unless specifically mentioned otherwise, our results apply for any finite alphabet) input distribution. However, when we refer to a certain input distribution by its constellation name e.g. BPSK input or 56-QAM input), a uniform distribution over the constellation points will be assumed. The maimum achievable rate for reliable communication under the assumptions of this model is given by the input-output Average Mutual Information [36]: I SC lim K K + I K K ; y K K ) = I ; y ) where I A ; B) denotes the mutual information between A and B, and I A ; B C) denotes the mutual information between A and B, conditioned on C c.f. [37]). When the input distribution is symmetric Gaussian, a closed-form epression for I SC can be easily derived cf. [7], [8]), I SC,Gaussian = ˆ π log + Hθ) ) dθ 3) π π We use the standard notation a N N for the sequence [a N, a N +,..., a N ] with the natural interpretation when N = and/or N =. A standard comple Gaussian variable is of the form n I + jn Q, where n I, n Q N, /) and n I n Q. )

5 Let I γ) I ; γ + n) 4) stand for the input-output mutual information in a scalar comple-valued Gaussian channel with unit-power input 3, SNR γ and standard comple Gaussian noise n, independent of. The Gaussian achievable rate can also be epressed as I SC,Gaussian = I Gaussian SNR MMSE-DFE-U ) 5) where { ˆ π SNR MMSE-DFE-U = ep log + Hθ) ) } dθ 6) π π is the output SNR of the unbiased MMSE linear estimator of given and y, known as the MMSE decision-feedback equalizer DFE), and I Gaussian γ) = log + γ) is the input-ouput mutual information for a comple-valued Gaussian channel with standard comple Gaussian input. A concise presentation of the MMSE descision-feedback equalizer, as well as a derivation of its output SNR can be found in [3]. When the input distribution in not Gaussian, no closed-form epression for I SC is known and it must be approimated either analytically [8], [38], [39] or by Monte-Carlo simulations [4] [4]. A simple and oftenused approimation for I SC was first proposed by Shamai and Laroia [8], I SC I SL I SNR MMSE-DFE-U ) 7) where is a random variable distributed as one of the input symbols to the ISI channel. This approimation can be derived by applying the MMSE DFE on the channel output sequence and replacing the residual ISI by independent Gaussian variables with equal power however, as eplained in [8], the central limit theorem cannot be used to rigorously justify this approimation, as the residual ISI coefficients do not meet its conditions. The Shamai-Laroia approimation was originally conjectured to be a lower bound on I SC. However, in [39] a countereample based on highly skewed binary input is constructed, showing that this conjecture does not hold for all input distributions. Nonetheless, etensive eperimentation has shown that when conventional input distributions are used, I SL is an etremely tight lower bound for I SC for any ISI channel and SNR [8], [38], [4]. In particular, there is no known countereample to I SC I SL that involves symmetrically distributed inputs as all conventional inputs are). Moreover, in [39] the lower bound I SC I SL is proven to hold for sufficiently high SNR, further establishing its validity. Whether I SC I SL can be proven to always hold for specific input distributions is a question open to future research. B. OFDM modulation model In OFDM, information is transmitted in blocks of N + N CP channel inputs, where the first N CP elements of each block are identical to its last N CP elements and thus constitute a cyclic prefi CP). With the CP discarded at 3 By Gaussian channel we mean a channel with additive Gaussian noise independent of the input, which is not necessarily Gaussian. The subscript will commonly be used to indicate that a general input distribution is discussed, and subscripts with indicative names will be employed when referring to specific input distributions, e.g. I BPSK γ) and I Gaussian γ).

6 the receiver, the ISI channel is transformed into a vector channel, y = H + n 8) The vectors y and represent blocks of N channel outputs and inputs, respectively, n is a standard comple Gaussian noise vector, and H is a matri representing the ISI. Practical OFDM schemes are designed so that the cyclic prefi is longer than the channel memory N CP > L). Moreover, we consider a sequence of schemes in which the block size N and the cyclic prefi size N CP grow to infinity, so that N CP is guaranteed to eceed L eventually. Assuming N CP > L, H is a circulant matri, with first row equal to [h,, h L,, h ]. Therefore, H is diagonalized by the the DFT matri of order N, H d = WHW 9) with H d a diagonal matri and W m,k = N e πjmk/n the DFT matri. Applying the input precoding = W and output transformation ỹ = Wy thus yields an equivalent diagonal vector channel, ỹ = H d + ñ ) We assume that the elements of are i.i.d. 4, zero mean and have unit average power. Since the channel model in ) is simply N parallel channels, the maimum achievable rate per channel input is given by I OFDM N) N + N CP I ; ỹ) = N + N CP N I H d i,i ) ) with I ) as defined in 4) and distributed as one of the elements of. The Toeplitz Distribution Theorem [43] allows us to take the limit of the large block size, I OFDM lim N I OFDM N) = π ˆ π π i= I Hθ) ) dθ ) where it is assumed that N CP grows as on), so that the rate overhead of the cyclic prefi vanishes. C. MMSE estimation in a scalar Gaussian channel Consider once more the scalar comple-valued Gaussian channel y = γ+n with input and standard comple Gaussian noise n independent of. The minimum mean square error MMSE) in estimating from y is given by mmse γ) = E E [ y] 3) For a standard comple Gaussian input we have mmse Gaussian γ) = / + γ) mmse γ), for any other unitpower input. 4 This is usually the case in wireless links, where the communication overhead of coordinating different powers and constellations for different subcarriers often makes doing so undesirable.

7 Assuming has zero mean and unit variance i.e. E = ), the connection between the mutual information 4) and the MMSE 3) is given by: I γ) = mmse γ) 4) Note that the above equation differs from the familiar Guo-Shamai-Verdú formula [9] by a factor of on the right-hand side. This is due to the fact that our channel model is comple-valued. The relation 4) can be easily derived from the vector version of the GSV theorem eq. ) in [9]), by considering a two-dimensional scalar channel matri. D. Statement of results Our main result provides a connection between I OFDM and I SL for finite-alphabet inputs. First, we show that an inequality of the form I OFDM I SL + always holds, where depends only on the input distribution and not on the ISI channel). Second, we characterize low and high SNR regions in which the strengthened inequality I OFDM I SL holds, even when. To this end, we introduce two pairs of SNR thresholds. The first pair, denoted γ and γ, constitutes low and high SNR thresholds with respect to SNR MMSE-DFE-U, defined in 6) when SNR MMSE-DFE-U is below γ or above γ, we have I OFDM I SL. The second threshold pair is denoted γ and γ, and relates to the channel s frequency response when Hθ) is bounded by γ from above or by γ from below, we are guaranteed once more to have I OFDM I SL. Like, these thresholds depend only on the input distribution. The eplicit construction of γ, γ, γ, γ and is deferred to Section III and Definitions, and 3 therein, as it relies on concepts and results developed there. These quantities are defined in terms of concavity properties of a certain function, and are straightforward to evaluate numerically. In particular, we provide an epression for in terms of a three-variable optimization problem which is easily solved numerically see 6) below. Formally stated, our main result is as follows, Theorem. For any ISI channel and any finite alphabet distribution, I OFDM I SL + 5) Where is given by log sup γ,γ,γ s.t. γ γ γ ) +γ +γ ) [I γ ) I γ)] log +γ +γ [I γ) I γ )] log [ + γ ] / [ + γ ]) 6) Additionally, if >, there eit < γ γ γ γ < that depend only on the input distribution, such that I OFDM I SL holds whenever the channel transfer function H θ) satisfies at least one of the following conditions: ) SNR MMSE-DFE-U [, γ ] [ γ, ) ) Hθ) γ for every θ π, π) 3) Hθ) γ for every θ π, π)

8 Net, we provide the following results, which sharpen Theorem for specific input distributions, Theorem. For BPSK and QPSK inputs, = and so I OFDM I SL for every ISI channel. Theorem 3. For M-PAM and square M -QAM inputs, d min /) γ, where d min is the minimum distance between input symbols, assuming unit input power. Combined with Theorem, Theorem 3 implies that for M-PAM and square M -QAM inputs, I OFDM I SL whenever the ISI channel is such that d min /) Hθ) for every θ π, π). Since the uncoded symbol error rate in OFDM subcarrier frequency θ is a function of d min /) Hθ ), the following corollary is immediate, Corollary. For a given ISI channel and square M -QAM inputs with M 6, if the uncoded symbol error rate is below 5% in all OFDM subcarriers, I OFDM I SL. Our last result deals with uniformly distributed input. This input distribution represents the limit of infinitely high-order QAM, and is therefore referred to also as. Since this input has an infinite alphabet, Theorem does not apply to it. Instead, we have Theorem 4. For uniformly distributed comple input and any ISI channel, ) I OFDM I SL ma H θ) θ π,π) 7) where.68 [bit], and γ), is a non-decreasing function that satisfies γ) = for every γ γ ) 8.76 [db], and lim γ) = log πe/6).59 [bit] 8) γ is the uniform input shaping loss with respect to Gaussian input. Moreover, I OFDM I SL when H θ) satisfies at least one of the following conditions: ) SNR MMSE-DFE-U γ ) 6.5 [db] ) Hθ) γ ) 8.76 [db] for every θ π, π) Note that Theorem 4 provides some of the guarantees of Theorem. In particular, we have I OFDM I SL + for any ISI channel, as well as I OFDM I SL for every channel that satisfies Hθ) γ ). Figure graphs γ), which is evaluated numerically based on the analysis carried out in subsection V-D. As seen in the figure, the convergence of γ) to is etremely slow. Thus, for any practical purpose we may select an SNR level γ that is much higher than any plausible value of Hθ), and use γ) instead of. Depending on the application, appropriate choices of γ are likely to yield values between 3 db).84 and 6 db).8. Table I summarizes a numerical study of the quantities that appear in Theorem. The table is consistent with Theorem and indicates it also etends to 4-PAM, 8-PSK, 6-QAM and 3-QAM inputs. It also reveals that while nonzero, 64-QAM is negligible, being of the order of a millionth of a bit. For higher order constellations is more significant, but remains quite small even in very high-order constellations such as 496-QAM. The limiting

9 Table I NUMERICAL EVALUATION OF THE QUANTITIES APPEARING IN THEOREM Input dmin ) [db] γ [db] γ [db] γ [db] γ [db] [bits] BPSK, 4-PAM, QPSK, 8-PSK, varies) - - - - 6-QAM, 3-QAM 64-QAM 6..7..7..86 6 56-QAM.3 5.9 9. 9... 4-QAM 8.3 3.63 8.8 5. 7.5.585 496-QAM 34.4.6 8.77 3. 33.6.987-8.76 - -.59 Note: All values are rounded to three significant digits. case of input is also included in the table. It is seen that 496-QAM has a value of γ quite close to that of, but that 496-QAM is still far from converging to. Eamining Figure, it is seen that for the various QAM inputs considered, is well-approimated by γ ). Finally, the table shows that the general bound provided by Theorem 3 is slack by an approimate factor of i.e., for higher-order QAM, d min /) γ /. Our results show that I OFDM may only eceed I SL by a small amount, but the opposite is not true. Indeed, in subsection VI-A we construct a family of channels for which I OFDM tends to zero while I SL tends to the input entropy and in subsection VI-B we discuss practical scenarios in which the SC achievable rate is significantly higher than the OFDM rate. However, Theorem 4 indicates that this difference can be made small by increasing the constellation order. Subsection VI-C further discusses this course of action. E. A note on linear precoding Our results etend straightforwardly to the following generalized problem setting. In the SC case, we add a linear precoding filter that is applied on the i.i.d. inputs prior to their transmission. In the OFDM case, we allow a different power allocation for each subcarrier. More concretely, the SC precoded transmitted symbols are precoded n = i c i n i 9) where is an i.i.d. input sequence and the precoder taps satisfy i c i =. The OFDM transmitted symbols are precoded i = P i i ) where N are the i.i.d. OFDM block inputs as in ), and the power allocations satisfy i P i =. Clearly, precoded SC modulation with taps c and ISI channel H θ) is equivalent to normal SC with channel

.7.6 Information [bits].5.4.3. γ). 5 5 γ [db] Figure. Numerical evaluation of γ). C θ) H θ) where C θ) = k c ke ikθ. Moreover, OFDM with non-uniform power allocation is equivalent to normal OFDM with channel P θ) H θ) where P πi/n) = P i. Equating C θ) with P θ) we conclude that for any SC linear precoder there eist an OFDM power allocation such that both yield the same equivalent ISI channel. Hence, given a degree of freedom in choosing any SC linear precoder and any OFDM power allocation policy, our results are still applicable, revealing that SC has significant advantages over OFDM in this case as well. The introduction of linear precoding lends additional viability to our assumption of i.i.d. input, as the capacityachieving SC scheme can be viewed as i.i.d. Gaussian inputs linearly precoded with an optimal Waterfilling filter. Thus, it is reasonable to assume that when combined with a suitably chosen linear precoder, statistically independent symbols will be close to optimal even when input alphabet constraints prohibit Gaussian signaling. For OFDM with independent non-gaussian inputs, an optimal power allocation policy called Mercury/Waterfilling was proposed in [44]. However, the Mercury/Waterfilling spectrum does not necessarily describe the optimal linear precoder for the i.i.d. non-gaussian single carrier case. Using the methods described in [45], it should be possible to find the taps of this optimal precoder. A. Log-SNR scale III. CONCAVITY OF MUTUAL INFORMATION WITH RESPECT TO LOG-SNR For a given SNR γ, define the log-snr as ζ = log + γ), and let I log ζ) I e ζ ) )

Information [bits] 9 8 7 6 5 4 BPSK 4-PAM QPSK 8-PSK 6-QAM 3-QAM 64-QAM 8-QAM 56-QAM Gaussian 3 4 6 8 ζ [bits] Figure. I log ζ) for some common input distributions. be the input output mutual information, as a function of log-snr, for a scalar comple Gaussian channel with zeromean, unit-variance input. Since ζ is identical to I log Gaussian ζ), it is naturally measured in units of information. Moreover, we have I log ζ) I log Gaussian ζ) = ζ for all inputs. Figure shows Ilog ζ) for some common input distributions. As can be seen in the figure, I log ζ) is nearly linear for low ζ and, for finite alphabet inputs, it is nearly constant for high ζ, with the shoulder occurring at around the input entropy. The main results of this paper hinge on the concavity properties of I log ζ). In this section we study these properties for a general input distribution. We begin by showing that I log ζ) is concave for sufficiently low and sufficiently high ζ. Net, we consider the concave envelope of I log ζ) and show that it must equal I log ζ) for sufficiently low and sufficiently high ζ. Finally, we apply these conclusions to prove Theorem. B. Asymptotic concavity results Proposition. For every input distribution, there eists < ζ such that I log ζ) is concave for every ζ [, ζ ].

Proof: Setting γ = e ζ and differentiating I log twice, we find that I log ζ) = e ζ I e ζ ) + e ζ I e ζ ) = + γ) I γ) + + γ) I γ) = + γ) [mmse γ) + + γ) mmse γ)] ) = + γ) d dγ [ + γ) mmse γ)] = + γ) r γ) 3) where the transition to ) is due to the I-MMSE relation 4). The function r γ) + γ) mmse γ) denotes the ratio between the MMSE s of the non-linear and linear optimal estimators of in the scalar comple Gaussian channel with SNR γ. Clearly, r γ), and r ) =. Therefore, by continuity there must be a neighborhood of, denoted by [, γ ], in which r is decreasing. Hence, by 3) we find that I log ζ) is concave in [, ζ ], with ζ = log + γ ). Proposition. For every input distribution over a finite alphabet, there eists ζ < such that I log ζ) is concave for every ζ [ζ, ]. Proof: Let d min denote the minimum distance between any two symbols in the input alphabet. By the standard probability of error upper bound c.f. Appendi C in [44]), we have mmse γ) D e dmin/) γ 4) for some D >. Moreover in Appendi A it is shown that mmse γ) C e dmin/) γ γ 5) for sufficiently large γ and some C >. Therefore, denoting again γ = e ζ and substituting the above bounds in ), we find that I log ζ) + γ) D + γ ) C e dmin/) γ γ for some C > and sufficiently large γ. Clearly, this implies I log ζ) < for sufficiently large ζ. Definition. Let < ζ be the maimal ζ for which I log ζ) is concave for every ζ [, ζ ] and similarly let ζ < be the minimal ζ for which I log ζ) is concave for every ζ [ζ, ). Let γ = e ζ and γ = e ζ denote the SNR s corresponding to ζ and ζ, respectively. Notice that ζ < ζ if and only if I log ζ) is not a concave function of ζ, in which case I log 6) ζ ) = I log ζ ) = 7)

3 C. Concave envelope Let Îlog also given by, Clearly, Îlog is real-analytic, Î log following, ζ) denote the concave envelope [46] of I log, i.e. the smallest concave function that upper bounds I log, Î log ζ) = sup ζ,ζ s.t. ζ ζ ζ [ ζ ζ ) I log ζ ) + ζ ζ) I log ] ζ ) ζ ζ eists and is upper bounded by ζ, which is a concave functions that upper bounds I log ζ). Since I log ζ) is continuous and has a continuous derivative. Moreover, our previous results allow the Proposition 3. For every input distribution with finite alphabet, There eist ζ > and ζ < such that Îlog ζ) = I log ζ) for every ζ [, ζ ] [ζ, ). Proof: At any point ζ, the concave envelope Îlog ζ, such that the concave envelope is equal to I log set of disjoint intervals {[ζ,i, ζ,i ]} i S such that Moreover, since Îlog for every ζ [ζ,i, ζ,i ]. Î log ζ) = 8) ζ) is either equal to I log ζ) or linear in an interval containing at the edges of the intervals. Put in other words, there eists a ζ ζ,i ζ,i ζ,i I log ζ,i ) + ζ,i ζ ζ,i ζ,i I log ζ,i ) ζ [ζ,i, ζ,i ] I log ζ) otherwise ζ) is also continuous, the above statement implies that Îlog 9) ζ) = I log ζ,i ) = I log ζ,i ) Suppose by contradiction that there eists i such that ζ,i =. Denoting γ = e ζ, By the I-MMSE relationship we have Îlog ζ) = r ζ) = + γ) mmse γ). Since the input is normalized to unit power, we have I log ) = r ) =, and so there must eist ζ,ii > such that I log ζ,i ) = I log ζ,i ) =. However, since the input is not Gaussian it has finite alphabet), and since mmse ) =, the single-crossing property [] implies that mmse γ) < / + γ) for every γ > and therefore I log ζ) < for every ζ >, contradicting I log ζ,i ) =. Hence, setting ζ = min i S ζ i,, we find that I log ζ) = Îlog ζ) for any ζ [, ζ ]. Since the input has finite alphabet, Îlog ζ) converges to a finite value the input entropy). By the data-processing inequality, mutual information is an increasing function of SNR, and since the mapping ζ = log + γ) is monotonic and increasing, it follows that I log ζ) is also increasing in ζ. Therefore, I log ζ) > for any ζ < and lim ζ I log ζ) =. Moreover, by Proposition we know that there eist ζ < such that I log ζ) is monotonically decreasing for every ζ > ζ. By the above observations, there must eist ζ ζ < such that ζ a ζ ζ b implies I log ζ a ) > I log ζ ) I log ζ b ). Clearly ζ,i ζ for any i S, as otherwise the equality I log ζ,i ) = I log ζ,i ) contradicts I log for any ζ [ζ, ), concluding our proof. ζ,i ) > I log ζ ) I log Definition. Let < ζ be the maimal ζ for which Îlog let ζ < be the minimal ζ for which Îlog ζ) = I log γ = e ζ denote the SNR s corresponding to ζ and ζ, respectively. ζ,i ). Therefore, we have I log ζ) = Îlog ζ) ζ) = I log ζ) for every ζ [, ζ ] and similarly ζ) for every ζ [ζ, ). Let γ = e ζ and

4 9.95 ζ ζ 8 7 I log ζ) [bit] Î log ζ) [bit].9 ζ ζ 6 ζ.85.8.75 I log ζ) [bit/bit] Î log ζ) [bit/bit].7 4 6 8 ζ [bits] 5 4 3 4.54 4.5 4.5 log ζ 4.48 4.46 4.44 4.74 4.75 4.76 4 6 8 ζ [bits] Figure 3. I log and Îlog right), and their derivatives with respect to ζ left), with ζ, ζ, ζ, ζ and highlighted, for 56-QAM input. Notice that in light of the above definition, the optimization in the definition of the concave envelope 8) can be limited to ζ and ζ in the interval [ζ, ζ ]. Definition 3. Let denote that maimum difference between I log ζ) and its concave envelope. By the above definition and 8) we have, Using 9), it is seen that [Îlog = ma ζ log = sup γ,γ,γ s.t. γ γ γ = Îlog = [ I log ζ m ) I log ζ) I log ) +γ +γ ] ζ) for some i S and ζ m [ζ,i, ζ,i ] that satisfies I log Clearly, = if and only if I log [I γ ) I γ)] log +γ +γ ζ m ) = ζ m ζ,i ) I log log [ + γ ] / [ + γ ]) ) [I γ) I γ )] ζ m ) [ I log ζ m ) I log ζ,i ) ] ζ,i ) I log ζ m ) ] ζ,i ζ m ) I log ζ m ) 3) is not concave. Figure 3 illustrates Îlog on a 56-QAM constellation two 6-PAM constellations in quadrature). 3) 3) ζ m ) = I log ζ,i ) = I log ζ,i ). is concave in R +. As seen from Table I, is quite small, even when I log, ζ, ζ, ζ, ζ and as defined above, for an input uniformly distributed

5 D. Proof of Theorem Proof: The inequality 5) is immediate from the definitions of I SL, I OFDM, I log I OFDM = π = π π Îlog ˆ π π ˆ π π ˆ π π π, Îlog and : I Hθ) ) dθ 33) I log log + Hθ) )) dθ 34) Î log log + Hθ) )) dθ 35) ˆ π π log + Hθ) ) ) dθ 36) = Îlog log + SNR MMSE-DFE-U )) 37) I SNR MMSE-DFE-U ) + = I SL + 38) where in 36) the concavity of Îlog was used to invoke Jensen s inequality. Choosing γ and γ to be as defined in Definition, it is clear that if condition holds, then Î log log + SNR MMSE-DFE-U )) = I log log + SNR MMSE-DFE-U )) = I SL 39) and I OFDM I SL is therefore valid. Choosing γ and γ according to Definition, if either condition or conditions 3 hold, then I log is a concave function for every value of Hθ), and we may therefore echange Îlog in 36), yielding I OFDM I SL once more. with I log IV. INTERLUDE MMSE BOUNDS FOR PAM SIGNALING In order to prove Theorems and 3, we first need to derive a tight high-snr characterization of the MMSE function and its derivative, for PAM inputs and Gaussian noise. In this section, we first present a novel pointwise bound on the MMSE of PAM signaling conditioned on the channel output. The bound is then applied to derive inequalities that characterize the MMSE and its derivative in the high-snr regime, as required. Finally, some bounds on the MMSE and its derivative are presented for the special case of BPSK input. Besides their use in proving Theorems and 3, the results presented in this section may be of general interest. A. A pointwise MMSE inequality Let X be a real-valued random variable and define 5 Y γ = X + γ N with N N, ) and independent of X. Y γ is the Gaussian-noise contaminated version of X, at SNR γ. Let ] φ X y; γ) = E X [ X E [X Y γ = y] Y γ = y 4) 5 We have chosen here to let γ scale N and not X, as opposed to the convention in the I-MMSE literature. This definition ensures that Y γ and X are on the same scale, which simplifies many of the derivations that follow. We have also set the noise variance to be / when γ =, in the purpose of making these results easily applicable in the comple setting of the rest of the paper.

6 denote the point-wise conditional variance of X given a noisy channel observation. Clearly, mmse X γ) = E Yγ φ X Y γ ; γ). Moreover, eploration of Information-Estimation relations [] has revealed that, mmse Xγ) = E Yγ φ X Y γ ; γ) 4) Hence, intimate understanding of φ X y; γ) is epected to provide insights on both the MMSE function and its derivative. For the case of PAM input distribution, this understanding is presented in the form of the following, Theorem 5. Let X = {,..., M } be the alphabet of an M-ary PAM constellation such that m+ m = d for every m < M. Let X be uniformly distributed in X. Fi y R and choose J < M such that J, J+ are the nearest values to y in X. Let B J be uniformly distributed on { J, J+ }. Then, ) d ) d φ X y; γ) φ BJ y; γ) D γ) 4) with Proof: Appendi B. Note that φ BJ y; γ) = D γ) = 4 k= ) d φ BPSK k + ) e 4γk 6e 4γ e 4γ ) 3 43) d ) [ y ] ) J + J+ d ; γ) where φ BPSK y; γ) = tanh γy) is the conditional variance function for BPSK input. Put in words, Theorem 5 means that for PAM input and given channel output, the epected value of the MMSE is lower-bounded by the epected MMSE given the same channel output and assuming an input equally distributed on the two PAM symbols nearest to it. Figure 4 illustrates the bounds in 4) for 4-PAM input. As the figure indicates, the lower bound is reasonably tight for d/) γ as low as db, and both bounds are very tight for d/) γ = 3 db and above. 44) B. High-SNR characterization of the MMSE Let mmse d,m-pam γ) stand for the mmse X γ), with X uniformly distributed on an M-PAM constellation with distance d between adjacent points. We show that mmse d,m-pam γ) M M ) d mmse BPSK d ) γ) in the sense that the difference between the terms tends to zero with a faster eponential rate than mmse M P AM γ), where mmse BPSK γ) mmse, P AM γ). Moreover, we use similar techniques in order to show that this 45)

7..8.6.4. Φ X y;γ) Φ BJ y;γ) Φ. BJ y;γ)+d/) Dd/) γ) X.8.6.4. 5 4 3 y 3 4 5 a) 5 4 3 y 3 4 5 b) Figure 4. Illustration of Theorem 5 for 4-PAM input with d = and two different SNR s: a) d/) γ = and b) d/) γ =. approimation also applies to the derivative of the MMSE with respect to γ, i.e., mmse d,m-pam γ) M ) 4 d ) d mmse BPSK γ) M 46) It should be noted that 45) can be seen as a special case of the high-snr MMSE characterization for general discrete inputs that was recently presented in [47]. However, the following analysis provides two important advantages. First, it enables us to also establish 46) a characterization of the derivative of the MMSE. Second, it allows for eplicit and tight bounds on the difference between the eact M-PAM quantities and their BPSK approimations. Both of these features are crucial for establishing the bound in Theorem 3. Letting Q ) ˆ denote the standard error function, the result 45) is stated formally as follows, π e t / dt 47) Theorem 6. The following bounds hold for every M, d and γ : mmse d,m-pam γ) M ) [ d ) ) d ) d mmse BPSK γ + γ)] M B 48) with and ) B γ) = 6Q 8γ + 4 mmse d,m-pam γ) M M k= k + ) Q k ) 8γ ) [ d ) ) d mmse BPSK γ B d ) γ)] 49) 5)

8 with ) B γ) = 4Q 8γ e 4γ 5) πγ Proof: Appendices C-A and C-D. Similarly, 46) has the following formal form, Theorem 7. The following bounds hold for every M, d and γ : mmse d,m-pam γ) M ) [ 4 d ) d ) d mmse BPSK γ) + γ)] M C 5) with ) C γ) = 3e 8γ Q 3γ 4 e 8γ 53) πγ and mmse d,m-pam γ) M M ) [ 4 d mmse BPSK d ) d ) γ) C γ)] 54) with [ ) ) ) C γ) = 4 k + ) e 8 ] 4γk k + ) e 4γk + + Q 8γ 55) k= k= Proof: Appendices C-C and C-B. [ Figure 5 illustrates the high SNR behavior of mmse d,m-pam γ) / M d ) ] M and mmse d,m-pam [ γ) / M d ) 4 ] M for different values of M. It is seen that the above bounds become tight at d/) γ values of around 3 db. C. Bounds for BPSK inputs We present some upper and lower bound on the MMSE function and its derivative for the case of BPSK inputs. These bounds will be of use in proving analytically that = for BPSK and QPSK inputs, as claimed in Theorem. Theorem 8. The following bounds on mmse BPSK γ) hold π ) π e γ mmse BPSK γ) γ 8 γ π γ e γ 56) e γ + γ mmse BPSK γ) e γ 57)

9.5.4 M mmse M PAM γ)/ [ M M d/)] mmse BPSK d/) γ)+ Bd/) γ) mmse BPSK d/) γ) Bd/) γ).3.5 M.. 3 3 4 5 d/) γ db mmse M PAM γ)/[ M M d/)4] mmse BPSK d/) γ)+ Cd/) γ) mmse BPSK d/) γ) Cd/) γ).5 3 3 4 5 d/) γ db Figure 5. Illustration of Theorems 6 and 7 for M values of 3, 4, 5, 8 and 6. similarly, the following bounds on mmse BPSK γ) hold )) π π γ 8 π e γ mmse γ BPSK γ) e γ 58) γ e γ mmse + 6γ BPSK γ) e γ 59) Proof: Appendi D. We note that the bounds in 56) are composed of the first two terms in the asymptotic high SNR epansion of mmse BPSK γ) derived in [44]. Our contribution in this case is the proof that these approimations are upper and lower bounds. Our proof of the bounds also allows for a simpler derivation of the series epansion than the one in [44], as well as etension of these bounds to mmse BPSK γ). It is worth noting that while not asymptotically tight as γ, the lower bound e γ / + γ is a good approimation for mmse BPSK γ) for all values of γ, with a maimum slackness of less than.. V. RESULTS FOR PAM AND SQUARE QAM INPUTS In this section we utilize the results from Sections III and IV in order to prove Theorems 3, and 4. A. Proof of Theorem We recall the definition of I log ) and prove the following, Lemma. I log BPSK ζ) and Ilog QPSK ζ) are concave in ζ.

8, Proof: Let γ = e ζ. As shown in ), I log ζ) = + γ) [mmse γ) + + γ) mmse γ)]. By Theorem [ ] mmse BPSK γ) + + γ) mmse BPSKγ) e γ + γ) < 6) + 6γ and so I log BPSK ζ) < for every ζ, meaning I log BPSK is concave in ζ. Since, γ ) mmse QPSK γ) = mmse BPSK 6) we have, [ mmse QPSK γ) + + γ) mmse QPSKγ) e γ/ + γ ] + 3γ 6) and it is easily verified that + γ) / + 3γ for every γ. For lower SNR s, we use the Gaussian upper bound mmse QPSK γ) + γ) as well as e γ/ γ/ to show that [ ] mmse QPSK γ) + + γ) mmse QPSKγ) + γ) γ/) + γ + 3γ and once more it is easily verified that + γ) γ/) / + 3γ for every γ. We thus conclude that ζ) < for every ζ, and so I log is concave in ζ. I log QPSK Clearly, if I log is concave then I log QPSK ζ) = Îlog 63) ζ) for every ζ and by its definition 3), =. Thus, Lemma proves Theorem. While numeric investigation shows that I log 4-PAM, Ilog 8-PSK, Ilog 6-QAM and Ilog 3-QAM are also concave, no analytical proof of concavity has been found for these cases. B. Proof of Theorem 3 Proof: Let γ = e ζ and let ρ = d PAM min /) γ where d PAM min = M is the minimum distance between symbols of a unit-power M-PAM input. Using ) and bounds provided in Theorems 6 and 7, we find that I log M-PAM ζ) = + γ) [mmsem-pam γ) + + γ) mmse M-PAMγ)] K [ ρmmse BPSK ρ) + mmse BPSK ρ) + B ρ) + ρ C ρ) ] 64) with B ρ) and C ρ) given in 49) and 53), respectively, and K = + γ) M M d PAM min /). Evaluating numerically the epression in square brackets, it is found that I log M-PAM ζ) < for ρ, see Figure 6. Consequently, d PAM min /) γ for M-PAM inputs, as required. An M -ary square QAM constellation is composed of two M-ary PAM constellations in quadrature. Letting d QAM min = 6 M denote the minimum distance between symbols of a unit-power M -QAM input, we have mmse M -QAM γ) = mmse d QAM min,m-pam γ) 65) and mmse M -QAM γ) = mmse γ) 66) d QAM min,m-pam

..5. Ρ.5 Figure 6. Evaluation of ρe ρ ) [ ρmmse BPSK ρ) + mmse BPSK ρ) + B ρ) + ρ C ρ) ]. where mmse d,m-pam γ) denotes the MMSE for an M-PAM input with distance d between adjacent symbols and comple-valued additive Gaussian noise with power /γ, as defined in subsection IV-B. Applying Theorems 6 ) and 7 to the above equations, we find that I log M -QAM ζ) is also bounded from above by 64), with ρ = d QAM min / γ. ) Therefore, we have d QAM min / γ for M -QAM inputs, and the proof is complete. Eamining Figure 6 more closely, we find that the term 64) becomes negative for ρ values around.95, and the bound on d min /) γ might be slightly tightened accordingly. However, as remarked on Table I, numerical evaluation of γ for square QAM inputs indicate that d min /) γ is closer to.5. Therefore, reducing the bound by.5 does not significantly improve its tightness. C. Proof of Corollary Proof: In the large block size limit, the input SNR at the k th OFDM subcarrier is given by γ k = H θ k ) where θ k = πk/n is the subcarrier frequency and k is its inde spanning from to N ). Consider a unit-power square M -QAM input and Gaussian noise at SNR γ, and set q = Q d min /) γ, with the error function Q ) as given in 47). For M-PAM input with spacing d min, the probability of a symbol error is q for the M inner constellation points, and q for the outer points, i.e. P M-PAM err = M M q + M q = M M q 67) For M -QAM input, a symbol error event is the union of two independent error events along the in-phase and

.98.96 8 6.65.6 ζ m.55.6.7.8 ζ.94.9 ζ m ζ.9 4 6 8 ζ [bits] ζ Î log ζ) [bit/bit] I log ζ) [bit/bit] Ǐ log ζ) [bit/bit] 4 ζ ζ) [bit] I log ζ) [bit] Ǐ log ζ) [bit] 4 6 8 ζ [bits] Î log Figure 7. I log with its concave and conve envelopes right), and their derivatives with respect to ζ left), with ζ, ζ, and highlighted, for input. quadrature directions, each being M-ary PAM error events. Therefore, P M -QAM err = P M-PAM err P M-PAM err ) M = 4 M q M M q ) 68) It can thus be seen that for M 6, P M -QAM err < 5% implies d min /) γ > and hence γ > γ. Assuming this holds for all subcarriers and assuming large enough OFDM block size, we find that H θ) > γ for all values of θ, thus satisfying condition 3 of Theorem and proving the corollary. D. Analysis of uniform input [ ] [ ] 3 We consider a unit-power input distributed uniformly on the square, 3 3, 3, also referred to as input. As usual we let I γ) denote the mutual information between such input and its comple Gaussian noise corrupted version at SNR γ, and we let I log ζ) = I e ζ ) be the mutual information with respect to log SNR. Figure 7 illustrates I log and its derivative, as well as the quantities to be defined in the following paragraphs. For high SNR, it well known [48] that and therefore ) 6 I γ) log πe γ ) lim ζ I log πe ) ζ) = log.59 [bit] 7) ζ 6 as ζ is also the mutual information for Gaussian input at log-snr ζ, the above limit represent the loss of using 69)

3 uniform input rather than Gaussian input at high SNR, and is commonly referred to as the shaping gain. Proposition applies to the case of uniform input, and so we know there must a constant ζ such I log ζ) is concave for every ζ ζ. However, the high-snr behavior of I log ζ) is quite different than the finite-alphabet case, and is characterized as follows, Proposition 4. I log ζ) is concave for every ζ ζ and conve for every ζ ζ, where ζ 3.9 bits. Proof: Appendi E. Even though it becomes conve in high SNR s, I log still has a concave envelope. Additionally, it is of interest to study the conve envelope of I log, i.e. the maimum conve function that lower bounds Ilog, which we denote by Ǐlog. Moreover, we are interested in the concave envelope of Ilog when limited to the interval [, ζ], i.e. the minimum concave function that upper bounds I log for every ζ [, ζ ]. We will denote this function by Îlog;[, ζ] ζ). These concave and conve envelopes are characterized as follows, Proposition 5. The concave envelope of I log is given by Î log ζ) = lim Î log;[, ζ] ζ) = ζ 7) ζ and satisfies Îlog = sup ζ ) πe ) ζ) Ilog ζ) = log 6 7) Limited to the interval [, ζ ], where ζ > ζ, the concave envelope of I log is given by Î log;[, ζ] ζ) = I log ζ ) + ζ ζ )Ilog ζ ) ζ ζ 73) I log ζ) otherwise where ζ < ζ depends on ζ and is determined by the condition Îlog;[, ζ] ζ) = I log ζ). The function γ) is given by, ) e ζ sup ζ ζ Î log;[, ζ] ) ζ) Ilog ζ) = Îlog;[, ζ] ζ m) I log ζ m) 74) where ζ m > ζ depends on ζ and satisfies I log ζ m ) = I log ζ ). For ζ ζ, I log interval [, ζ ], so that Îlog;[, ζ] ζ) = Ilog ζ) and ) e ζ =. The conve envelope of I log is given by Ǐ log ζ) = is concave on the ζi log ζ ) ζ ζ 75) I log ζ) otherwise with ζ 5.5 [bits] determined by the continuity condition Ǐlog ζ ) = I log ζ ). The constant is

4 given by Proof: Appendi F. [ ] sup I log ζ) Ǐlog ζ).68 [bit] 76) ζ We let γ = e ζ 8.76 db and γ = e ζ 6.5 db. Armed with the above results, the proof of Theorem 4 is straightforward, Proof of Theorem 4: Letting γ = ma θ π,π) H θ) and ζ = log + γ), the proof that I OFDM I SL + γ) is identical to the proof of Theorem, with Îlog;[, ζ] ζ) replacing Îlog ζ) and γ) replacing. To show that I SL I OFDM + we reverse the direction of the derivation: I OFDM = π = π π where in 8) the conveity of Ǐlog ˆ π π ˆ π π ˆ π Ǐ log π Ǐlog the function I log and obtain I OFDM I SL. When SNR MMSE-DFE-U γ, I Hθ) ) dθ 77) I log log + Hθ) )) dθ 78) log + Hθ) )) dθ 79) ˆ π log + Hθ) ) ) dθ π 8) π = Ǐlog log + SNR MMSE-DFE-U)) 8) I SNR MMSE-DFE-U ) = I SL 8) was used to invoke Jensen s inequality. When Hθ) γ for all θ π, π), is conve for all values of log + Hθ) ), and we may therefore replace Ǐlog with Ilog Ǐ log log + SNR MMSE-DFE-U)) = I SNR MMSE-DFE-U ) 83) and hence the introduction of is unnecessary, resulting once more in I OFDM I SL. VI. DISCUSSION In this section we use the results obtained in the paper to draw insight on the differences between the maimum achievable rates of SC and OFDM that epected in practical scenarios. Additionally, we consider how increasing the constellation order affects these difference and discuss the implications of doing so. A. Maimum and minimum difference I OFDM and I SL Our analysis enables us to characterize the ISI channels for which I SL I OFDM will be smallest, and the channels for which it will be the largest. If the input distribution is such that = then clearly every memoryless flat fading) channel achieves the minimum difference I SL I OFDM =. If >, the minimum difference is obtained

5 by a channel with two-level transfer function, H θ) γ = γ ) ) θ π log +γ +γ m / log +γ +γ otherwise 84) with γ m = e ζm and ζ m as defined in 3). Clearly for this channel I SL I OFDM =, which is the minimum possible difference according to Theorem. Consider the channel, H θ) e Γ θ π/γ = otherwise where Γ > is an arbitrary sharpness parameter. As Γ) increases, the channel s frequency response becomes narrower and steeper. For this channel, I OFDM = I e Γ /Γ and I SL = I e Γ ). For any finite-alphabet input, we therefore have I SL I OFDM H ) as Γ, where H ) denotes the input entropy. Clearly, H ) is the maimum possible difference between I SL and I OFDM, as it upper bounds both quantities. Now consider uniform input, characterized in Theorem 4 and analyzed in subsection V-D. Since I γ) log γ) log πe/6) at high SNR, for the above channel we will have I OFDM I SL log πe/6) = as Γ. Thus, the etreme case that maimizes I SL I OFDM for finite-alphabet inputs also minimizes it for uniform inputs. However, is approached only for highly impractical channels. For eample, in order for I OFDM I SL to reach 9% of, we need Γ, which yields H θ) 43 db! Indeed, can only be approached as H θ) becomes eceedingly large this is proven by the very slow rate of convergence of γ) Figure ). 85) B. Difference between I OFDM and I SL in practical settings In OFDM wireless communication systems, the constellation and error correcting code will usually be chosen so that the code rate is between / and 5/6 [8] []. Assuming the system is efficient enough to have performance close to the maimum achievable rate, this means we would have H ) I OFDM = π ˆ π π I log ζ θ ) dθ 5 6 H ) 86) where H ) is the input entropy which equals the number of uncoded bits per input symbol for equiprobably inputs, and ζ θ = log + H θ) ) is the log-snr at subcarrier frequency θ. Let ζ OFDM be the log-snr in a memoryless channel with achievable rate of I OFDM, i.e. I log ζ OFDM ) = I OFDM. Under this notation, the single carrier achievable rate satisfies, I SC I SL = I SNR MMSE-DFE-U ) = I log ˆ π ) ζ θ dθ π π Eamining Figures and 3 while keeping 86) in mind, we are able to estimate the performance gain of SC over OFDM for different ISI channels. Clearly, for channels with little ISI nearly constant H θ) ), there will 87)