This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Tight Upper Bounds on the Error Probability of Spinal Codes over Fading Channels

Xiaomeng Chen, Aimin Li and Shaohua Wu Aimin Li has contributed equally to this work. This work has been supported in part by the National Key Research and Development Program of China under Grant no. 2020YFB1806403, and in part by the National Natural Science Foundation of China under Grant nos. 61871147, 62071141, 61831008, 62027802, and in part by the Shenzhen Municipal Science and Technology Plan under Grant no. GXWD20201230155427003-20200730122528002, and in part by the Major Key Project of PCL under Grant PCL2021A03-1. (Corresponding author: Shaohua Wu.) Harbin Institute of Technology (Shenzhen)
518055 Shenzhen, China
[email protected], [email protected], [email protected]
Abstract

Spinal codes, a family of rateless codes introduced in 2011, have been proved to achieve Shannon capacity over both the additive white Gaussian noise (AWGN) channel and the binary symmetric channel (BSC). In this paper, we derive explicit tight upper bounds on the error probability of Spinal codes under maximum-likelihood (ML) decoding and perfect channel state information (CSI) over three typical fading channels, including the Rayleigh channel, the Nakagami-m channel and the Rician channel. Simulation results verify the tightness of the derived upper bounds.

Index Terms:
Spinal codes, decoding error probability, fading channels, ML decoding. upper bounds.

I Introduction

First proposed in 2011 [1], Spinal codes are a family of rateless codes that have been proved to achieve Shannon capacity over both the AWGN channel and the BSC [2]. Possessing the rateless capacity-achieving nature, Spinal codes demonstrate their superiority in bridging a reliable and high-efficiency information delivery pipeline between transceivers under highly dynamic channels. Specifically, the rateless nature allows Spinal codes to automatically adapt to time-varying channel conditions. Unlike fixed-rate codes, which require a specific code rate in advance, rateless codes work by a natural channel adaptation manner: the sender transmits a potentially limitless stream of encoded bits, and the receiver collects bits consecutively until the successful decoding process takes place. In [3], Spinal codes have shown advantages in terms of rate performance compared to other state-of-the-art rateless codes under different channel conditions and message sizes. Also, [3] notes the similarity between Spinal codes and Trellis Coded Modulation (TCM) [4, 5] is superficial because of their differences in nature and encoded popuse.

In coding theory, performance analysis is an intriguing topic. A closed-form expression for the error probability, in general, could not only enable more efficient performance evaluations but also shed light on coding optimization design. However, in most cases, it is intractable to obtain a closed-form expression for error probabilities. As an alternative, bounding techniques are usually used to approximate performance [6]. Along this avenue, there are already a lot of established bounds for some specific channel codes, such as the advanced tight bounds on Polar codes [7, 8], the upper and lower bounds on Raptor codes [9], and the performance bounds on the LT codes [10, 11]. And in [12], upper and lower bounds on the error probability of linear codes under ML decoding are surveyed. For Spinal codes, however, it is still early days to get tight explicit bounds over fading channels.

To date, there have been a few works that have preliminarily explored the performance analysis of Spinal codes. In [2], Balakrishnan et al. analyze the asymptotic rate performance of Spinal codes and theoretically prove that Spinal codes are capacity-achieving over both the AWGN channel and the BSC. In [13], two state-of-the-art results in the finite block length (FBL) regime, i.e., the Random Coding Union (RCU) bound [14] and a variant of Gallager result [15] are applied to analyze the error probability of Spinal codes over the AWGN channel and the BSC, respectively. In [16], the authors further tighten bounds by characterizing the error probability as the volume of a hypersphere divided by the volume of a hypercube, while the analysis is still performed over the AWGN channel. Until now, little has been done in the way of error probability analysis for Spinel codes over fading channels. An exception work is [17], which derived a probability-dependent convergent upper bound over Rayleigh fading channels by adopting the Chernoff bound [18]. However, ii) the derived bound over the Rayleigh fading channel in [17] is not strictly explicit, since the convergence of the upper bound is probability-dependent; iiii) Nakagami-m and Rician channels, both common in practical wireless communication scenarios, have not been considered in the available error probability analyses; iiiiii) There is not yet an upper bound that achieves uniform tight approximations over a wide range of signal-to-noise ratio (SNR), either over the fading channel or over the AWGN channel.

Motivated by the above, this paper aims to derive tight upper bounds over three typical fading channels, Rayleigh, Nakagami-m and Rician fading channels. Strictly explicit upper bounds are provided, which cover uniform tight error probability approximations over a wide range of SNR.

The paper is organized as follows. In Section II, the encoding process of Spinal codes is given as a priori knowledge. Section III derives upper bounds over Rayleigh, Nakagami-m and

Refer to caption
Figure 1: The encoding process of Spinal codes.

Rician fading channels, respectively. Numerical results comparing the bounds with Monte Carlo simulations are illustrated in Section IV. Section V presents the conclusions of our work.

II Encoding Process of Spinal Codes

This section briefly introduces the primary building blocks of Spinal codes, where the combination of a hash function and random number generator (RNG) functions is a key. Fig. 1 shows the encoding process of Spinal codes, which comprises five steps:

  1. 1.

    Segmentation: Divide an nn-bit message 𝐌\mathbf{M} into kk-bit segments 𝐦i{0,1}k\mathbf{m}_{i}\in\left\{0,1\right\}^{k}, where i=1,2,,n/ki=1,2,\cdots,n/k.

  2. 2.

    Sequentially Hashing111A discussion concerning the properties of the hash function can be found in Appendix A.: The hash function sequentially generates vv-bit spine values 𝐬i{0,1}v\mathbf{s}_{i}\in{\{0,1\}}^{v}, with

    𝐬i=h(𝐬i1,𝐦i),i=1,2,,n/k,𝐬0=𝟎v.\mathbf{s}_{i}=h(\mathbf{s}_{i-1},\mathbf{m}_{i}),~{}i=1,2,\cdots,n/k,~{}\mathbf{s}_{0}=\mathbf{0}^{v}\text{.} (1)
  3. 3.

    RNG: Each spine value 𝐬i\mathbf{s}_{i} is used to seed an RNG to generate a binary pseudo-random uniform-distributed sequence {𝐱i,j}j+\{\mathbf{x}_{i,j}\}_{j\in\mathbb{N}^{+}}:

    RNG:𝐬i{𝐱i,j},j=1,2,3,,\mathrm{RNG:}~{}\mathbf{s}_{i}\to\{\mathbf{x}_{i,j}\},~{}j=1,2,3,\cdots\text{,} (2)

    where 𝐱i,j{0,1}c\mathbf{x}_{i,j}\in{\{0,1\}}^{c}.

  4. 4.

    Constellation Mapping: The constellation mapper maps each cc-bit symbol 𝐱i,j\mathbf{x}_{i,j} to a channel input set Ψ\Psi:

    f:𝐱i,jΨ.\operatorname{f}:\mathbf{x}_{i,j}\rightarrow\Psi\text{.} (3)

    In this paper, f\operatorname{f} is a uniform constellation mapping function, i.e., it converts each cc-bit symbol 𝐱i,j\mathbf{x}_{i,j} to a decimal symbol f(𝐱i,j)\operatorname{f}(\mathbf{x}_{i,j}).

  5. 5.

    Rateless Transmission: The encoder continues to generate and transmit symbols pass by pass until the receiver successfully decodes the message.

III UPPER BOUNDS ON THE ERROR PROBABILITY

Error probability analysis is of primary importance in coding theory. Closed-form expressions for error probabilities, however, are intractable in most cases. To address this issue, a commonly adopted alternative approach is to introduce bounding techniques for the error probability approximation. In line with this idea, this section aims to derive tight upper bounds on the error probability for Spinal codes over Rayleigh, Nakagami-m and Rician fading channels, respectively.

III-A Upper Bounds on the Rayleigh Fading Channel

Theorem 1.

Consider Spinal codes with message length nn, segmentation parameter kk, modulation parameter cc, and sufficiently large hash parameter vv transmitted over a flat slow Rayleigh fading channel with mean square Ω\Omega and AWGN variance σ2\sigma^{2}, the average error probability given perfect channel state information (CSI) under ML decoding for Spinal codes can be upper bounded by

Pe1a=1n/k(1ϵa),\small P_{e}\leq 1-\prod_{a=1}^{n/k}\left(1-\epsilon_{a}\right)\text{,} (4)

with

ϵa=min{1,(2k1)2nakRayleigh(La,σ)},\epsilon_{a}=\mathrm{min}\left\{1,\left(2^{k}-1\right)2^{n-ak}\cdot\mathscr{F}_{\text{Rayleigh}}(L_{a},\sigma)\right\}\text{,} (5)
Rayleigh(La,σ)=r=1NbrRayleigh(θr;σ,La),\small\mathscr{F}_{\text{Rayleigh}}\left(L_{a}\text{,}\sigma\right)=\sum_{r=1}^{N}b_{r}\mathcal{F}_{\text{Rayleigh}}(\theta_{r}\text{;}\sigma\text{,}L_{a})\text{,} (6)
Rayleigh(θr;σ,La)=(iΨjΨ22c8σ2sin2θrΩ(ij)2+8σ2sin2θr)La,\small\mathcal{F}_{\text{Rayleigh}}(\theta_{r}\text{;}\sigma\text{,}L_{a})={\left(\sum_{i\in\Psi}\sum_{j\in\Psi}2^{-2c}\frac{8{\sigma}^{2}{\sin}^{2}{\theta}_{r}}{{\Omega}(i-j)^{2}+8{\sigma}^{2}{\sin}^{2}{\theta}_{r}}\right)}^{L_{a}}\text{,} (7)

where br=θrθr1πb_{r}=\frac{\theta_{r}-\theta_{r-1}}{\pi}, La=(n/ka+1)LL_{a}=(n/k-a+1)L, LL is the number of transmitted passes, θr\theta_{r} is arbitrarily chosen with θ0=0\theta_{0}=0, θN=π2\theta_{N}=\frac{\pi}{2} and 0<θ1<θ2<<θN1<π20<\theta_{1}<\theta_{2}<\cdots<\theta_{N-1}<\frac{\pi}{2}, and NN represents the number of θ\theta values which enables the adjustment of accuracy.

Proof.

Suppose a message 𝐌=(𝐦1,𝐦2,,𝐦n/k){0,1}n\mathbf{M}=\left(\mathbf{m}_{1},\mathbf{m}_{2},\cdots,\mathbf{m}_{n/k}\right)\in{\{0,1\}}^{n} is encoded to f(𝐱i,j(𝐌))\operatorname{f}\left(\mathbf{x}_{i,j}(\mathbf{M})\right) to be transmitted over a flat slow Rayleigh fading channel with an AWGN. The received signal can be expressed as:

yi,j=hi,jf(𝐱i,j(𝐌))+ni,j,y_{i,j}=h_{i,j}\operatorname{f}\left(\mathbf{x}_{i,j}(\mathbf{M})\right)+n_{i,j}\text{,} (8)

where yi,jy_{i,j} is the corresponding received signal at the receiver, hi,jh_{i,j} is the channel fading parameter following Rayleigh distribution with mean square Ω\Omega, and ni,jn_{i,j} is the AWGN with variance σ2\sigma^{2}.

ML decoding selects the one with the lowest decoding cost from the candidate sequence space {0,1}n{\{0,1\}}^{n}. Given received symbols yi,jy_{i,j} and perfect CSI h^i,j=hi,j\hat{h}_{i,j}=h_{i,j}, the ML rule for the Rayleigh fading channel with the AWGN is

𝐌^argmin𝐌¯{0,1}ni=1n/kj=1L(yi,jhi,jf(𝐱i,j(𝐌¯)))2,\small\begin{split}\hat{\mathbf{M}}&\in\underset{\mathbf{\bar{M}}\in{\{0,1\}}^{n}}{\mathrm{arg\,min}}\sum_{i=1}^{n/k}\sum_{j=1}^{L}{(y_{i,j}-h_{i,j}\operatorname{f}(\mathbf{x}_{i,j}(\bar{\mathbf{M}})))}^{2}\text{,}\end{split} (9)

where 𝐌^\hat{\mathbf{M}} is the decoding result, 𝐌¯\bar{\mathbf{M}} is the candidate sequence.

From this perspective, we classify the set of candidate sequences {0,1}n{\{0,1\}}^{n} into two subsets: ii) the correct decoding sequence 𝐌=(𝐦1,𝐦2,,𝐦n/k)\mathbf{M}=\left(\mathbf{m}_{1},\mathbf{m}_{2},\cdots,\mathbf{m}_{n/k}\right); iiii) wrong decoding sequences 𝐌=(𝐦1,𝐦2,,𝐦n/k)𝒲\mathbf{M}^{\prime}=\left(\mathbf{m}^{\prime}_{1},\mathbf{m}^{\prime}_{2},\cdots,\mathbf{m}^{\prime}_{n/k}\right)\in\mathcal{W}, with 𝒲{(𝐦1,𝐦2,,𝐦n/k):1in/k,𝐦i𝐦i}\mathcal{W}\triangleq\left\{\left(\mathbf{m}^{\prime}_{1},\mathbf{m}^{\prime}_{2},\cdots,\mathbf{m}^{\prime}_{n/k}\right):\exists 1\leq i\leq n/k,\mathbf{m}^{\prime}_{i}\neq\mathbf{m}_{i}\right\}. Denoting the cost of 𝐌\mathbf{M} as 𝒟(𝐌)\mathscr{D}(\mathbf{M}), it turns out that

𝒟(𝐌)i=1n/kj=1L(yi,jhi,jf(𝐱i,j(𝐌)))2=i=1n/kj=1Lni,j2.\displaystyle\small\begin{split}\mathscr{D}(\mathbf{M})&\triangleq\sum_{i=1}^{n/k}\sum_{j=1}^{L}{\left(y_{i,j}-h_{i,j}\operatorname{f}(\mathbf{x}_{i,j}(\mathbf{M}))\right)}^{2}=\sum_{i=1}^{n/k}\sum_{j=1}^{L}n_{i,j}^{2}\text{.}\end{split} (10)

Similarly, denote the cost of 𝐌\mathbf{M}^{\prime} as 𝒟(𝐌)\mathscr{D}(\mathbf{M}^{\prime}), given by:

𝒟(𝐌)i=1n/kj=1L(yi,jhi,jf(𝐱i,j(𝐌)))2.\displaystyle\mathscr{D}(\mathbf{M}^{\prime})\triangleq\sum_{i=1}^{n/k}\sum_{j=1}^{L}{\left(y_{i,j}-h_{i,j}\operatorname{f}(\mathbf{x}_{i,j}(\mathbf{M}^{\prime}))\right)}^{2}\text{.} (11)

In the sequal, we attempt to explicitly express the error probability of Spinal codes. Let a\mathcal{E}_{a} be the event that there exists an error in the atha^{th} segment, which implies that:

  1. 1.

    The atha^{th} segment is different, i.e., 𝐦a𝐦a\mathbf{m}_{a}\neq\mathbf{m}^{\prime}_{a}.

  2. 2.

    The cost of the wrong decoding sequence is less than the correct one, i.e., 𝒟(𝐌)𝒟(𝐌)\mathscr{D}(\mathbf{M}^{\prime})\leq\mathscr{D}(\mathbf{M}). In this case, the ML decoder will incorrectly choose a certain wrong sequence 𝐌𝒲\mathbf{M}^{\prime}\in\mathcal{W} as the decoding output.

Denote ¯a\overline{\mathcal{E}}_{a} as the complement of a\mathcal{E}_{a}. The error probability of Spinal codes can be expressed as:

Pe=Pr(a=1n/ka)=1Pr(a=1n/k¯a)=1a=1n/k[1Pr(a|i=1a1¯i)].\small\begin{split}P_{e}&=\mathrm{Pr}\left(\bigcup_{a=1}^{n/k}\mathcal{E}_{a}\right)=1-\mathrm{Pr}\left(\bigcap_{a=1}^{n/k}\overline{\mathcal{E}}_{a}\right)\\ &=1-\prod_{a=1}^{n/k}\left[1-\mathrm{Pr}\left(\mathcal{E}_{a}\bigg{|}\bigcap_{i=1}^{a-1}\overline{\mathcal{E}}_{i}\right)\right]\text{.}\end{split} (12)

Thus, to obtain the error probability PeP_{e}, the remaining issue is to calculate Pr(a|i=1a1¯i)\mathrm{Pr}\bigg{(}\mathcal{E}_{a}\bigg{|}\bigcap_{i=1}^{a-1}\overline{\mathcal{E}}_{i}\bigg{)}, which is interpreted as the probability that the atha^{th} segment is wrong while the previous (a1)(a-1) segments are correct. With the defination that 𝒲a{(𝐦1,,𝐦a):𝐦1=𝐦1,,𝐦a1=𝐦a1,𝐦a𝐦a}𝒲\mathcal{W}_{a}\triangleq\{\left(\mathbf{m}^{\prime}_{1},\cdots,\mathbf{m}^{\prime}_{a}\right)\text{:}\mathbf{m}^{\prime}_{1}=\mathbf{m}_{1},\cdots,\mathbf{m}^{\prime}_{a-1}=\mathbf{m}_{a-1},\mathbf{m}^{\prime}_{a}\neq\mathbf{m}_{a}\}\subseteq\mathcal{W}, the conditional probability is transformed as:

Pr(a|i=1a1¯i)=Pr(𝐌𝒲a:𝒟(𝐌)𝒟(𝐌)).\small\mathrm{Pr}\left(\mathcal{E}_{a}\bigg{|}\bigcap_{i=1}^{a-1}\overline{\mathcal{E}}_{i}\right)=\mathrm{Pr}\left(\exists\mathbf{M}^{\prime}\in\mathcal{W}_{a}:\mathscr{D}(\mathbf{M}^{\prime})\leq\mathscr{D}(\mathbf{M})\right)\text{.} (13)

Applying the union bound of probability [19] yields that

Pr(𝐌𝒲a:𝒟(𝐌)𝒟(𝐌))𝐌𝒲aPr(𝒟(𝐌)𝒟(𝐌)).\begin{split}&\mathrm{Pr}\left(\exists\mathbf{M}^{\prime}\in\mathcal{W}_{a}:\mathscr{D}(\mathbf{M}^{\prime})\leq\mathscr{D}(\mathbf{M})\right)\\ &\leq\sum_{\mathbf{M}^{\prime}\in\mathcal{W}_{a}}\mathrm{Pr}\left(\mathscr{D}(\mathbf{M}^{\prime})\leq\mathscr{D}(\mathbf{M})\right)\text{.}\end{split} (14)

Given that 𝒟(𝐌)\mathscr{D}(\mathbf{M}) and 𝒟(𝐌)\mathscr{D}(\mathbf{M}^{\prime}) have been calculated in (10) and (11) respectively, the probability Pr(𝒟(𝐌)𝒟(𝐌))\mathrm{Pr}\left(\mathscr{D}(\mathbf{M}^{\prime})\leq\mathscr{D}(\mathbf{M})\right) is calculated as follows:

Pr(𝒟(𝐌)𝒟(𝐌))\displaystyle\mathrm{Pr}\left(\mathscr{D}(\mathbf{M}^{\prime})\leq\mathscr{D}(\mathbf{M})\right)
=Pr(i=1n/kj=1L(yi,jhi,jf(xi,j(𝐌)))2i=1n/kj=1Lni,j2)\displaystyle=\mathrm{Pr}\bigg{(}\sum_{i=1}^{n/k}\sum_{j=1}^{L}{\left(y_{i,j}-h_{i,j}\operatorname{f}(x_{i,j}(\mathbf{M^{\prime}}))\right)}^{2}\leq\sum_{i=1}^{n/k}\sum_{j=1}^{L}n_{i,j}^{2}\bigg{)}
=(a)Pr(i=an/kj=1L(yi,jhi,jf(xi,j(𝐌)))2i=an/kj=1Lni,j2)\displaystyle\overset{(a)}{=}\mathrm{Pr}\bigg{(}\sum_{i=a}^{n/k}\sum_{j=1}^{L}{\left(y_{i,j}-h_{i,j}\operatorname{f}(x_{i,j}(\mathbf{M^{\prime}}))\right)}^{2}\leq\sum_{i=a}^{n/k}\sum_{j=1}^{L}n_{i,j}^{2}\bigg{)}
=(b)Pr(i=an/kj=1L[hi,j(f(xi,j(𝐌))f(xi,j(𝐌)))+ni,j]2\displaystyle\overset{(b)}{=}\mathrm{Pr}\bigg{(}\sum_{i=a}^{n/k}\sum_{j=1}^{L}{\left[h_{i,j}(\operatorname{f}(x_{i,j}(\mathbf{M}))-f(x_{i,j}(\mathbf{M^{\prime}})))+n_{i,j}\right]}^{2}
i=an/kj=1Lni,j2),\displaystyle\qquad\qquad\qquad\qquad\leq\sum_{i=a}^{n/k}\sum_{j=1}^{L}n^{2}_{i,j}\bigg{)}, (15)

where (aa) establishes since f(𝐱i,j(𝐌))=f(𝐱i,j(𝐌))\operatorname{f}(\mathbf{x}_{i,j}(\mathbf{M}))=\operatorname{f}(\mathbf{x}_{i,j}(\mathbf{M^{\prime}})) for 1i<a1\leq i<a, which is proved in Appendix A by leveraging the property of hash function. (bb) is obtained by applying (8).

Define Ui,jf(𝐱i,j(𝐌))f(𝐱i,j(𝐌))U_{i,j}\triangleq\operatorname{f}(\mathbf{x}_{i,j}(\mathbf{M}))-\operatorname{f}(\mathbf{x}_{i,j}(\mathbf{M^{\prime}})) and Vi,jhi,jUi,jV_{i,j}\triangleq h_{i,j}U_{i,j}, and then (III-A) can be expanded as

Pr(i=an/kj=1L[hi,j(f(xi,j(𝐌))f(xi,j(𝐌)))Vi,j+ni,j]2i=an/kj=1Lni,j2)=Pr(i=an/kj=1LVi,j2+2i=an/kj=1LVi,jni,j0).\small\begin{split}&\mathrm{Pr}\bigg{(}\sum_{i=a}^{n/k}\sum_{j=1}^{L}{\bigg{[}\underbrace{h_{i,j}(\operatorname{f}(x_{i,j}(\mathbf{M}))-\operatorname{f}(x_{i,j}(\mathbf{M^{\prime}})))}_{V_{i,j}}+n_{i,j}\bigg{]}}^{2}\\ &\qquad\qquad\quad\quad\leq\sum_{i=a}^{n/k}\sum_{j=1}^{L}n^{2}_{i,j}\bigg{)}\\ &=\mathrm{Pr}\left(\sum_{i=a}^{n/k}\sum_{j=1}^{L}V^{2}_{i,j}+2\sum_{i=a}^{n/k}\sum_{j=1}^{L}V_{i,j}n_{i,j}\leq 0\right)\text{.}\end{split} (16)

Denoting 𝐕La\mathbf{V}^{L_{a}} as the random row vector composed of random variables Vi,jV_{i,j} with ain/k,1jLa\leq i\leq n/k,1\leq j\leq L, and 𝐍La\mathbf{N}^{L_{a}} as the random row vector composed of random variables ni,jn_{i,j} with ain/k,1jlia\leq i\leq n/k,1\leq j\leq l_{i}, (16) could be rewritten into a vector form, given as

Pr(i=an/kj=1LVi,j2+2i=an/kj=1LVi,jni,j0)=Pr(𝐕La(𝐕La+2𝐍La)T0),\small\begin{split}&\mathrm{Pr}\left(\sum_{i=a}^{n/k}\sum_{j=1}^{L}V^{2}_{i,j}+2\sum_{i=a}^{n/k}\sum_{j=1}^{L}V_{i,j}n_{i,j}\leq 0\right)\\ &=\mathrm{Pr}\left(\mathbf{V}^{L_{a}}{\left(\mathbf{V}^{L_{a}}+2{\mathbf{N}}^{L_{a}}\right)}^{\mathrm{T}}\leq 0\right),\end{split} (17)

which could be then expanded as

LaPr(𝐕La(𝐕La+2𝐍La)T0|𝐕La=𝐯La)\displaystyle\int_{\mathbb{R}^{L_{a}}}\mathrm{Pr}\left(\mathbf{V}^{L_{a}}{\left(\mathbf{V}^{L_{a}}+2{\mathbf{N}}^{L_{a}}\right)}^{\mathrm{T}}\leq 0\big{|}\mathbf{V}^{L_{a}}=\mathbf{v}^{L_{a}}\right)\cdot (18)
Pr(𝐕La=𝐯La)d𝐯La\displaystyle\quad\quad\quad\quad\quad\quad\quad\quad\quad\quad\quad\quad\mathrm{Pr}\left(\mathbf{V}^{L_{a}}=\mathbf{v}^{L_{a}}\right){~{}\mathrm{d}\mathbf{v}^{L_{a}}}
=LaPr(𝐯La(𝐯La+2𝐍La)T0)\displaystyle=\int_{\mathbb{R}^{L_{a}}}\mathrm{Pr}\left(\mathbf{v}^{L_{a}}{\left(\mathbf{v}^{L_{a}}+2{\mathbf{N}}^{L_{a}}\right)}^{\mathrm{T}}\leq 0\right)\cdot
Pr(𝐕La=𝐯La)d𝐯La.\displaystyle\quad\quad\quad\quad\quad\quad\quad\quad\quad\quad\quad\quad\mathrm{Pr}\left(\mathbf{V}^{L_{a}}=\mathbf{v}^{L_{a}}\right){~{}\mathrm{d}\mathbf{v}^{L_{a}}}.

With (18) in hand, the next problem is to explicitly solve Pr(𝐯La(𝐯La+2𝐍La)T0)\mathrm{Pr}\left(\mathbf{v}^{L_{a}}{\left(\mathbf{v}^{L_{a}}+2{\mathbf{N}}^{L_{a}}\right)}^{\mathrm{T}}\leq 0\right) and Pr(𝐕La=𝐯La)\mathrm{Pr}\left(\mathbf{V}^{L_{a}}=\mathbf{v}^{L_{a}}\right). First, we attempt to simplify Pr(𝐯La(𝐯La+2𝐍La)T0)\mathrm{Pr}\left(\mathbf{v}^{L_{a}}{\left(\mathbf{v}^{L_{a}}+2{\mathbf{N}}^{L_{a}}\right)}^{\mathrm{T}}\leq 0\right). By introducing two rotation matrices for LaL_{a}-dimensions hyperspace, we obtain Lemma 1 as follows:

Lemma 1.

Given that ni,jn_{i,j} is i.i.d AWGN with variance σ2\sigma^{2}, the probability in (18) can be simplified as

Pr(𝐯La(𝐯La+2𝐍La)T0)=Q(𝐯La2σ),\small\mathrm{Pr}\left(\mathbf{v}^{L_{a}}{\left(\mathbf{v}^{L_{a}}+2{\mathbf{N}}^{L_{a}}\right)}^{\mathrm{T}}\leq 0\right)=Q\left(\frac{\left\|\mathbf{v}^{L_{a}}\right\|}{2\sigma}\right), (19)

where Q()Q(\cdot) represents the QQ function, \left\|\cdot\right\| represents the 2\ell^{2} norm.

Proof.

Please refer to Appendix B. ∎

Note that [20] has shown a transformation of Q()Q(\cdot), which is presented as an exponential form, given as

Q(x)=1π0π2exp(x22sin2θ)dθ.\small Q(x)=\frac{1}{\pi}\int_{0}^{\frac{\pi}{2}}\mathrm{exp}\left(\frac{-x^{2}}{2\mathrm{sin}^{2}\theta}\right)\mathrm{d}\theta\text{.} (20)

Adopting (20) into (19) yields that

Pr(𝐯La(𝐯La+2𝐍La)T0)\displaystyle\mathrm{Pr}\left(\mathbf{v}^{L_{a}}{\left(\mathbf{v}^{L_{a}}+2{\mathbf{N}}^{L_{a}}\right)}^{\mathrm{T}}\leq 0\right)
=Q(𝐯La2σ)=1π0π2exp(𝐯La28σ2sin2θ)dθ.\displaystyle=Q\left(\frac{\left\|\mathbf{v}^{L_{a}}\right\|}{2\sigma}\right)=\frac{1}{\pi}\int_{0}^{\frac{\pi}{2}}\mathrm{exp}\left(\frac{-{\left\|\mathbf{v}^{L_{a}}\right\|}^{2}}{8\sigma^{2}\mathrm{sin}^{2}\theta}\right)\mathrm{d}\theta\text{.} (21)

Applying (III-A) in (18) and swapping the integrates, we have

Pr(𝐕La(𝐕La+2𝐍La)T0)=1πLa0π2exp(𝐯La28σ2sin2θ)dθPr(𝐕La=𝐯La)d𝐯La=1π0π2Laexp(𝐯La28σ2sin2θ)Pr(𝐕La=𝐯La)d𝐯LaRayleigh(θ;σ,La)dθ.\small\begin{split}&\mathrm{Pr}\left(\mathbf{V}^{L_{a}}{\left(\mathbf{V}^{L_{a}}+2{\mathbf{N}}^{L_{a}}\right)}^{\mathrm{T}}\leq 0\right)\\ &=\frac{1}{\pi}\int_{\mathbb{R}^{L_{a}}}\int_{0}^{\frac{\pi}{2}}\mathrm{exp}\left(\frac{-{\left\|\mathbf{v}^{L_{a}}\right\|}^{2}}{8\sigma^{2}\mathrm{sin}^{2}\theta}\right)\mathrm{d}\theta\cdot\mathrm{Pr}\left(\mathbf{V}^{L_{a}}=\mathbf{v}^{L_{a}}\right)\mathrm{d}\mathbf{v}^{L_{a}}\\ &=\frac{1}{\pi}\int_{0}^{\frac{\pi}{2}}\underbrace{\int_{\mathbb{R}^{L_{a}}}\mathrm{exp}\left(\frac{-{\left\|\mathbf{v}^{L_{a}}\right\|}^{2}}{8\sigma^{2}\mathrm{sin}^{2}\theta}\right)\cdot\mathrm{Pr}\left(\mathbf{V}^{L_{a}}=\mathbf{v}^{L_{a}}\right)~{}\mathrm{d}\mathbf{v}^{L_{a}}}_{\mathcal{F}_{\text{Rayleigh}(\theta;\sigma,L_{a})}}\mathrm{d}\theta\text{.}\end{split} (22)

For the Rayleigh fading channel, we denote the multiple integrals with respect to 𝐯La\mathbf{v}^{L_{a}} as Rayleigh(θ;σ,La)\mathcal{F}_{\text{Rayleigh}}\left(\theta;\sigma,L_{a}\right). By adopting the i.i.d of Vi,jV_{i,j}, given as

Pr(𝐕La=𝐯La)=i=an/kj=1LfVi,j(vi,j),\small\Pr\left(\mathbf{V}^{L_{a}}=\mathbf{v}^{L_{a}}\right)=\prod_{i=a}^{n/k}\prod_{j=1}^{L}f_{V_{i,j}}\left(v_{i,j}\right), (23)

Rayleigh(θ;σ,La)\mathcal{F}_{\text{Rayleigh}}\left(\theta;\sigma,L_{a}\right) could be decomposed as

Laexp(𝐯La28σ2sin2θ)i=an/kj=1LfVi,j(vi,j)i=an/kj=1Ldvi,j=(a)i=an/kj=1Lexp(vi,j28σ2sin2θ)fVi,j(vi,j)dvi,j=(b)(exp(va,128σ2sin2θ)fVa,1(va,1)dva,1)La,\small\begin{split}&\underbrace{\int_{\mathbb{R}}\cdots\int_{\mathbb{R}}}_{L_{a}}\mathrm{exp}\left(\frac{-{\left\|\mathbf{v}^{L_{a}}\right\|}^{2}}{8\sigma^{2}\mathrm{sin}^{2}\theta}\right)\prod_{i=a}^{n/k}\prod_{j=1}^{L}f_{V_{i,j}}(v_{i,j})\prod_{i=a}^{n/k}\prod_{j=1}^{L}\mathrm{d}v_{i,j}\\ &\overset{(a)}{=}\prod_{i=a}^{n/k}\prod_{j=1}^{L}\int_{\mathbb{R}}\mathrm{exp}\left(\frac{-v_{i,j}^{2}}{8\sigma^{2}\mathrm{sin}^{2}\theta}\right)f_{V_{i,j}}(v_{i,j})~{}\mathrm{d}v_{i,j}\\ &\overset{(b)}{=}{\left(\int_{\mathbb{R}}{\mathrm{exp}}{\left(-\frac{v_{a,1}^{2}}{{8}{\sigma^{2}}{\sin^{2}}\theta}\right)}f_{V_{a,1}}(v_{a,1})\mathrm{d}v_{a,1}\right)}^{L_{a}}\text{,}\end{split} (24)

where (a) establishes by adopting

exp(𝐯La28σ2sin2θ)=exp(i=an/kj=1Lvi,j28σ2sin2θ)=i=an/kj=1Lexp(vi,j28σ2sin2θ),\small\begin{split}\exp\left(\frac{-\|\mathbf{v}^{L_{a}}\|^{2}}{8\sigma^{2}\sin^{2}\theta}\right)&=\exp\left(\frac{-\sum_{i=a}^{n/k}\sum_{j=1}^{L}v^{2}_{i,j}}{8\sigma^{2}\sin^{2}\theta}\right)\\ &=\prod_{i=a}^{n/k}\prod_{j=1}^{L}\exp\left(\frac{-v^{2}_{i,j}}{8\sigma^{2}\sin^{2}\theta}\right)\text{,}\end{split} (25)

and (b) holds for the i.i.d of the random variable Vi,jV_{i,j}. Recall that Vi,j=hi,jUi,jV_{i,j}=h_{i,j}U_{i,j} where hi,jh_{i,j} and Ui,jU_{i,j} are independent with each other, the integral in (24) with respect to va,1v_{a,1} could be then transformed to

exp(va,128σ2sin2θ)fVa,1(va,1)dva,1\xlongequalva,1=huupU(u)exp(h2u28σ2sin2θ)g1(h)dh,\small\begin{split}&\int_{\mathbb{R}}{\mathrm{exp}}{\left(-\frac{v_{a,1}^{2}}{{8}{\sigma^{2}}{\sin^{2}}\theta}\right)}f_{{V}_{a,1}}(v_{a,1})\mathrm{d}v_{a,1}\\ &\xlongequal{v_{a,1}=hu}\sum_{u}p_{{}_{U}}(u)\int_{\mathbb{R}}{\mathrm{exp}}{\left(-\frac{{h^{2}}{u^{2}}}{{8}{\sigma^{2}}{\sin^{2}}\theta}\right)}g_{{}_{1}}(h)\mathrm{d}h\text{,}\end{split} (26)

where pU(u)p_{{}_{U}}(u) is the probability mass function (PMF) of Ua,1U_{a,1} and g1(h)g_{{}_{1}}(h) is the probability density function (PDF) of hh. At this point, the right-hand side (RHS) of (26) could be explicitly calculated by the following lemma.

Lemma 2.

For Rayleigh distribution whose PDF is g1(h)g_{{}_{1}}(h), the RHS of (26) can be calculated by

iΨjΨ22c8σ2sin2θΩ(ij)2+8σ2sin2θ,\small\sum_{i\in\Psi}\sum_{j\in\Psi}2^{-2c}\frac{8{\sigma}^{2}{\sin}^{2}{\theta}}{{\Omega}(i-j)^{2}+8{\sigma}^{2}{\sin}^{2}{\theta}}\text{,} (27)

where Ψ\Psi is the channel input set.

Proof.

Please refer to Appendix C for the detailed derivation. ∎

As such, substituting (27) back into (24) turns out that

Rayleigh(θ;σ,La)=(iΨjΨ22c8σ2sin2θΩ(ij)2+8σ2sin2θ)La.\small\mathcal{F}_{\text{Rayleigh}}(\theta\text{;}\sigma\text{,}L_{a})={\left(\sum_{i\in\Psi}\sum_{j\in\Psi}2^{-2c}\frac{8{\sigma}^{2}{\sin}^{2}{\theta}}{{\Omega}(i-j)^{2}+8{\sigma}^{2}{\sin}^{2}{\theta}}\right)}^{L_{a}}\text{.} (28)

With (28) in hand, the RHS of (LABEL:eq12) is transformed as

𝐌𝒲aPr(𝒟(𝐌)𝒟(𝐌))=1π𝐌𝒲a0π2Rayleigh(θ;σ,La)dθ.\small\begin{split}&\sum_{\mathbf{M}^{\prime}\in\mathcal{W}_{a}}\mathrm{Pr}\left(\mathscr{D}\left(\mathbf{M}^{\prime}\right)\leq\mathscr{D}\left(\mathbf{M}\right)\right)\\ &=\frac{1}{\pi}\sum_{\mathbf{M}^{\prime}\in\mathcal{W}_{a}}\int_{0}^{\frac{\pi}{2}}\mathcal{F}_{\text{Rayleigh}}(\theta\text{;}\sigma\text{,}L_{a})\mathrm{d}\theta\text{.}\end{split} (29)

Note that Rayleigh(θ;σ,La)\mathcal{F}_{\text{Rayleigh}}(\theta\text{;}\sigma\text{,}L_{a}) is inreasing with θ\theta for 0θπ20\leq\theta\leq\frac{\pi}{2} (see Appendix D for the detailed proof), so we can arbitrarily choose N+1N+1 values of θ\theta such that θ0=0\theta_{0}=0, θN=π2\theta_{N}=\frac{\pi}{2} and 0<θ1<θ2<<θN1<π20<\theta_{1}<\theta_{2}<\cdots<\theta_{N-1}<\frac{\pi}{2} to explicitly upper bound the error probability. Combining these values back into (29), we get the inequality as

𝐌𝒲aPr(𝒟(M)𝒟(M))=1π𝐌𝒲a0π2Rayleigh(θ;σ,La)dθ|𝒲a|πr=1Nθr1θrRayleigh(θr;σ,La)dθ=|𝒲a|r=1NbrRayleigh(θr;σ,La),\small\begin{split}&\sum_{\mathbf{M^{\prime}}\in\mathcal{W}_{a}}\mathrm{Pr}\left(\mathscr{D}\left(M^{\prime}\right)\leq\mathscr{D}\left(M\right)\right)\\ &=\frac{1}{\pi}\cdot\sum_{\mathbf{M}^{\prime}\in\mathcal{W}_{a}}\int_{0}^{\frac{\pi}{2}}\mathcal{F}_{\text{Rayleigh}}(\theta\text{;}\sigma\text{,}L_{a})\mathrm{d}\theta\\ &\leq\frac{\left|\mathcal{W}_{a}\right|}{\pi}\cdot\sum_{r=1}^{N}\int_{\theta_{r-1}}^{\theta_{r}}\mathcal{F}_{\text{Rayleigh}}(\theta_{r}\text{;}\sigma\text{,}L_{a})\mathrm{d}\theta\\ &=\left|\mathcal{W}_{a}\right|\cdot\sum_{r=1}^{N}b_{r}\mathcal{F}_{\text{Rayleigh}}(\theta_{r}\text{;}\sigma\text{,}L_{a})\text{,}\end{split} (30)

where

br=θrθr1π,|𝒲a|=(2k1)2nak.\small b_{r}=\frac{\theta_{r}-\theta_{r-1}}{\pi}\text{,}\quad\left|\mathcal{W}_{a}\right|=(2^{k}-1)2^{n-ak}. (31)

Denoting

Rayleigh(La,σ)=r=1NbrRayleigh(θr;σ,La),\small\mathscr{F}_{\text{Rayleigh}}\left(L_{a}\text{,}\sigma\right)=\sum_{r=1}^{N}b_{r}\mathcal{F}_{\text{Rayleigh}}(\theta_{r}\text{;}\sigma\text{,}L_{a})\text{,} (32)

and applying (32) to (LABEL:eq12) results in the explicit bound. ∎

III-B Bounds on the Nakagami-m Fading Channel

Theorem 2.

Consider Spinal codes with message length nn, segmentation parameter kk, modulation parameter cc and sufficiently large hash parameter vv transmitted over a flat slow Nakagami-m fading channel with mean square Ω\Omega, shape parameter mm and AWGN variance σ2\sigma^{2}. The average error probability given perfect CSI under ML decoding for Spinal codes can be upper bounded by

Pe1a=1n/k(1ϵa),\displaystyle\small P_{e}\leq 1-\prod_{a=1}^{n/k}\left(1-\epsilon_{a}\right)\text{,} (33)

with

ϵa=min{1,(2k1)2nakNakagami(La,σ)},\epsilon_{a}=\mathrm{min}\left\{1,\left(2^{k}-1\right)2^{n-ak}\cdot\mathscr{F}_{\text{Nakagami}}(L_{a},\sigma)\right\}\text{,} (34)
Nakagami(La,σ)=r=1NbrNakagami(θr;σ,La),\mathscr{F}_{\text{Nakagami}}\left(L_{a}\text{,}\sigma\right)=\sum_{r=1}^{N}b_{r}\mathcal{F}_{\text{Nakagami}}(\theta_{r}\text{;}\sigma\text{,}L_{a})\text{,} (35)
Nakagami(θr;σ,La)=(iΨjΨ22c(8mσ2sin2θrΩ(ij)2+8mσ2sin2θr)m)La,\small\begin{split}&\mathcal{F}_{\text{Nakagami}}(\theta_{r}\text{;}\sigma\text{,}L_{a})\\ &={\left(\sum_{i\in\Psi}\sum_{j\in\Psi}2^{-2c}{\left(\frac{8m{\sigma}^{2}{\sin}^{2}{\theta}_{r}}{{\Omega}(i-j)^{2}+8m{\sigma}^{2}{\sin}^{2}{\theta}_{r}}\right)}^{m}\right)}^{L_{a}}\text{,}\end{split} (36)

where br=θrθr1πb_{r}=\frac{\theta_{r}-\theta_{r-1}}{\pi}, La=(n/ka+1)LL_{a}=(n/k-a+1)L, LL is the number of transmitted passes, θr\theta_{r} is arbitrarily chosen with θ0=0\theta_{0}=0, θN=π2\theta_{N}=\frac{\pi}{2} and 0<θ1<θ2<<θN1<π20<\theta_{1}<\theta_{2}<\cdots<\theta_{N-1}<\frac{\pi}{2}, and N represents the number of θ\theta values which enables the adjustment of accuracy.

Proof.

Please refer to Appendix E. ∎

III-C Bounds on the Rician Fading Channel

Theorem 3.

Consider Spinal codes with message length nn, segmentation parameter kk, modulation parameter cc and sufficiently large hash parameter vv transmitted over a flat slow Rician fading channel with mean square Ω\Omega ,shape parameter KK and AWGN variance σ2\sigma^{2}. The average error probability given perfect CSI under ML decoding for Spinal codes can be upper bounded by

Pe1a=1n/k(1ϵa),\small P_{e}\leq 1-\prod_{a=1}^{n/k}\left(1-\epsilon_{a}\right)\text{,} (37)

with

ϵa=min{1,(2k1)2nakRician(La,σ)},\epsilon_{a}=\mathrm{min}\left\{1,\left(2^{k}-1\right)2^{n-ak}\cdot\mathscr{F}_{\text{Rician}}(L_{a},\sigma)\right\}\text{,} (38)
Rician(La,σ)=r=1NbrRician(θr;σ,La),\mathscr{F}_{\text{Rician}}\left(L_{a}\text{,}\sigma\right)=\sum_{r=1}^{N}b_{r}\mathcal{F}_{\text{Rician}}(\theta_{r}\text{;}\sigma\text{,}L_{a})\text{,} (39)
Rician(θr;σ,La)=(iΨjΨ22c8(K+1)σ2sin2θrΩ(ij)2+8(K+1)σ2sin2θrexp(8K(K+1)σ2sin2θrΩ(ij)2+8(K+1)σ2sin2θrK))La,\small\begin{split}&\mathcal{F}_{\text{Rician}}(\theta_{r}\text{;}\sigma\text{,}L_{a})\\ &=\Bigg{(}\sum_{i\in\Psi}\sum_{j\in\Psi}2^{-2c}\frac{8(K+1){\sigma}^{2}{\sin}^{2}{\theta}_{r}}{{\Omega}(i-j)^{2}+8(K+1){\sigma}^{2}{\sin}^{2}{\theta}_{r}}\\ &\qquad\cdot{\mathrm{exp}\left(\frac{8K(K+1){\sigma}^{2}{\sin}^{2}{\theta}_{r}}{{\Omega}(i-j)^{2}+8(K+1){\sigma}^{2}{\sin}^{2}{\theta}_{r}}-K\right)\Bigg{)}}^{L_{a}}\text{,}\end{split} (40)

where br=θrθr1πb_{r}=\frac{\theta_{r}-\theta_{r-1}}{\pi}, La=(n/ka+1)LL_{a}=(n/k-a+1)L, LL is the number of transmitted passes, θr\theta_{r} is arbitrarily chosen with θ0=0\theta_{0}=0, θN=π2\theta_{N}=\frac{\pi}{2} and 0<θ1<θ2<<θN1<π20<\theta_{1}<\theta_{2}<\cdots<\theta_{N-1}<\frac{\pi}{2}, and N represents the number of θ\theta values which enables the adjustment of accuracy.

Proof.

Please refer to Appendix F. ∎

IV Simulation Result

In this section, we conduct Monte Carlo simulations to illustrate the accuracy of the upper bounds derived in Section III. Since the exponential nature of the ML-decoding complexity, the input message bits nn is selected as low as n=8n=8 and the number of pass is set as L=6L=6 here for the easy ML-decoding simulation setup. The parameter vv is set to v=32v=32 for the implementation and experiments, demonstrated by the property 2 in Appendix A that the hash collision will occur only once per 2322^{32} hash function invocations on average. Furthermore, to improve the accuracy of approximation for upper bounds, we set N=20N=20 and the sample size of Monte Carlo simulations to be 10610^{6}.

We nomalize the mean square values of all fading channels, i.e., Ω=1\Omega=1. Besides the setting of Ω\Omega, for the Nakagami-m fading channel, we also set the shape parameter as m=2m=2. And for Rician fading channels, the shape parameter KK, defined as the ratio of the power contributions by line-of-sight path to the remaining multipaths, is set as K=0.5K=0.5 and K=1K=1, respectively.

Refer to caption
Figure 2: Upper bounds on the error probability of Spinal codes with n=8,v=32,L=6,c=8n=8,v=32,L=6,c=8 and k=2k=2 over several fading channels with Ω=1\Omega=1.

Fig.2 shows examples of the error probability for Rayleigh, Nakagami-m and Rician fading channels, respectively. All approximations are close to simulated values. We could observe that the FER curve for K=1K=1 is lower than the one for K=0.5K=0.5 in Rician fading channels, which is due to KK is the ratio of the energy in the specular path to the energy in the scattered paths, i.e., the larger KK, the more deterministic the channel.

V Conclusion

This paper analyzes the error probability of Spinal codes and derives upper bounds on the error probability under ML decoding over several fading channels, including the Rayleigh fading channel, the Nakagami-m fading channel and the Rician fading channel. Additionally, we conduct simulations for different fading channels and parameters. Our experimental examples show that the upper bounds we derived have a good performence on the estimation of the average error probablity.

The work in this paper may also inspire innovation about further research and efforts in related topics. The derived upper bounds can provide theoretical support and guidance for designing other high-efficiency coding-associated techniques, such as unequal error protection and concatenation with outer codes.

References

  • [1] J. Perry, H. Balakrishnan, and D. Shah, “Rateless Spinal Codes,” Proceedings of the 10th ACM Workshop on Hot Topics in Networks, 2011.
  • [2] H. Balakrishnan, P. Iannucci, J. Perry, and D. Shah, “De-randomizing shannon: The design and analysis of a capacity-achieving rateless code,” arXiv preprint arXiv:1206.0418, 2012.
  • [3] J. Perry, P. A. Iannucci, K. E. Fleming, H. Balakrishnan, and D. Shah, “Spinal Codes,” Proceedings of the ACM SIGCOMM 2012 Conference, pp. 49––60, 2012.
  • [4] G. Ungerboeck, “Channel coding with multilevel/phase signals,” IEEE Transactions on Information Theory, vol. 28, no. 1, pp. 55–67, 1982.
  • [5] G. Ungerboeck and I. Csajka, “On Improving data-link performance by increasing the channel alphabet and introducing sequence coding,” IEEE International Symposium on Information Theory, 1976.
  • [6] S. Yousefi and A. Khandani, “A new upper bound on the ML decoding error probability of linear binary block codes in AWGN interference,” IEEE Transactions on Information Theory, vol. 50, no. 12, pp. 3026–3036, 2004.
  • [7] D. Goldin and D. Burshtein, “Performance Bounds of Concatenated Polar Coding Schemes,” IEEE Transactions on Information Theory, vol. 65, no. 11, pp. 7131–7148, 2019.
  • [8] B. Shuval and I. Tal, “A Lower Bound on the Probability of Error of Polar Codes over BMS Channels,” IEEE Transactions on Information Theory, vol. 65, no. 4, pp. 2021–2045, 2019.
  • [9] F. Lázaro, G. Liva, G. Bauch, and E. Paolini, “Bounds on the Error Probability of Raptor Codes Under Maximum Likelihood Decoding,” IEEE Transactions on Information Theory, vol. 67, no. 3, pp. 1537–1558, 2021.
  • [10] B. Schotsch, G. Garrammone, and P. Vary, “Analysis of LT Codes over Finite Fields under Optimal Erasure Decoding,” IEEE Communications Letters, vol. 17, no. 9, pp. 1826–1829, 2013.
  • [11] B. Schotsch, Rateless coding in the finite length regime.   Hochschulbibliothek der Rheinisch-Westfälischen Technischen Hochschule Aachen, 2014.
  • [12] I. Sason and S. Shamai, “Performance Analysis of Linear Codes under Maximum-Likelihood Decoding: A Tutorial,” Foundations and Trends in Communications and Information Theory, vol. 3, no. 1–2, pp. 1–222, 2006.
  • [13] X. Yu, Y. Li, W. Yang, and Y. Sun, “Design and Analysis of Unequal Error Protection Rateless Spinal Codes,” IEEE Transactions on Communications, vol. 64, no. 11, pp. 4461–4473, 2016.
  • [14] Y. Polyanskiy, H. V. Poor, and S. Verdú, “Channel coding rate in the finite blocklength regime,” IEEE Transactions on Information Theory, vol. 56, no. 5, pp. 2307–2359, 2010.
  • [15] R. G. Gallager, Information theory and reliable communication.   Wiley, 1968, vol. 588.
  • [16] A. Li, S. Wu, J. Jiao, N. Zhang, and Q. Zhang, “New Upper Bounds on the Error Probability under ML Decoding for Spinal Codes and the Joint Transmission-Decoding System Design,” arXiv preprint arXiv:2101.07953, 2021.
  • [17] A. Li, S. Wu, J. Jiao, N. Zhang, and Q. Zhang, “Spinal Codes Over Fading Channel: Error Probability Analysis and Encoding Structure Improvement,” IEEE Transactions on Wireless Communications, vol. 20, no. 12, pp. 8288–8300, 2021.
  • [18] H. Chernoff, “A Measure of Asymptotic Efficiency for Tests of a Hypothesis Based on the sum of Observations,” The Annals of Mathematical Statistics, pp. 493–507, 1952.
  • [19] G. Boole, The mathematical analysis of logic.   Philosophical Library, 1847.
  • [20] U. Erez, M. D. Trott, and G. W. Wornell, “Rateless Coding for Gaussian Channels,” IEEE Transactions on Information Theory, vol. 58, no. 2, pp. 530–547, 2012.
  • [21] M. Mitzenmacher and E. Upfal, Probability and computing: Randomized algorithms and probabilistic analysis.   Cambridge University Press, 2005.
  • [22] W. Kaplan, Advanced calculus.   Pearson Education India, 1952.

Appendix A The properties of hash and related inferences

The core idea of Spinal codes is to apply a hash function sequentially to generate a pseudo-random mapping from message bits to coded bits. The hash function is expressed as

h:{0,1}v×{0,1}k{0,1}v,h:\{0,1\}^{v}\times\{0,1\}^{k}\rightarrow\{0,1\}^{v}, (41)

where vv and kk are both integers. In Spinal codes, hh is chosen from a pairwise independent family of hash functions \mathscr{H}[21] by uniformly using a random seed.

Property 1.

As Perry et al. indicate [3], the hash function employed by Spinal codes should have pairwise independent property:

Pr{h(𝐬,𝐦)=𝐱,h(𝐬,𝐦)=𝐱}=Pr{h(𝐬,𝐦)=𝐱}Pr{h(𝐬,𝐦)=𝐱}=22v,\begin{split}&\mathrm{Pr}\{h(\mathbf{s},\mathbf{m})=\mathbf{x},h(\mathbf{s}^{\prime},\mathbf{m}^{\prime})=\mathbf{x}^{\prime}\}\\ &=\mathrm{Pr}\{h(\mathbf{s},\mathbf{m})=\mathbf{x}\}\cdot\mathrm{Pr}\{h(\mathbf{s}^{\prime},\mathbf{m}^{\prime})=\mathbf{x}^{\prime}\}\\ &=2^{-2v}\text{,}\end{split} (42)

where (𝐬,𝐦)(\mathbf{s},\mathbf{m}) and (𝐬,𝐦)(\mathbf{s}^{\prime},\mathbf{m}^{\prime}) are arbitrarily chosen, (𝐬,𝐦)(𝐬,𝐦)(\mathbf{s},\mathbf{m})\neq(\mathbf{s}^{\prime},\mathbf{m}^{\prime}).

Property 2.

A sufficiently large v will lead to a zero-approaching hash collision probability, with [3]

Pr{h(𝐬,𝐦)=h(𝐬,𝐦)}=𝐱{0,1}vPr{h(𝐬,𝐦)=𝐱,h(𝐬,𝐦)=𝐱}Property1=2v22v=2v0,iffv0,\begin{split}&\mathrm{Pr}\{h(\mathbf{s},\mathbf{m})=h(\mathbf{s}^{\prime},\mathbf{m}^{\prime})\}\\ &=\sum_{\mathbf{x}\in{\{0,1\}}^{v}}\underbrace{\mathrm{Pr}\{h(\mathbf{s},\mathbf{m})=\mathbf{x},h(\mathbf{s}^{\prime},\mathbf{m}^{\prime})=\mathbf{x}^{\prime}\}}_{\mathrm{Property}\ref{property1}}\\ &=2^{v}\cdot 2^{-2v}\\ &=2^{-v}\approx 0,\ \text{iff}\ v\gg 0\text{,}\end{split} (43)

where (𝐬,𝐦)(\mathbf{s},\mathbf{m}) and (𝐬,𝐦)(\mathbf{s}^{\prime},\mathbf{m}^{\prime}) are arbitrarily chosen, (𝐬,𝐦)(𝐬,𝐦)(\mathbf{s},\mathbf{m})\neq(\mathbf{s}^{\prime},\mathbf{m}^{\prime}).

Recall that 𝐌\mathbf{M} is the correct decoding sequence and 𝐌𝒲\mathbf{M}^{\prime}\in\mathcal{W} is a certain wrong decoding sequence. We denote the spine values of 𝐌\mathbf{M} as s(𝐌)=(𝐬1,𝐬2,,𝐬n/k)s(\mathbf{M})=(\mathbf{s}_{1},\mathbf{s}_{2},\cdots,\mathbf{s}_{n/k}) and the spine values of 𝐌\mathbf{M}^{\prime} as s(𝐌)=(𝐬1,𝐬2,,𝐬n/k)s(\mathbf{M}^{\prime})=(\mathbf{s}^{\prime}_{1},\mathbf{s}^{\prime}_{2},\cdots,\mathbf{s}^{\prime}_{n/k}).

In (1), the process of generating 𝐬i\mathbf{s}_{i} has given that 𝐬i=h(𝐬i1,𝐦i)\mathbf{s}_{i}=h(\mathbf{s}_{i-1},\mathbf{m}_{i}), 𝐬0=𝟎v\mathbf{s}_{0}=\mathbf{0}^{v}. So it is natural to get that

Pr(𝐬i=𝐬i|𝐦i=𝐦i,𝐬i1=𝐬i1)=1,\mathrm{Pr}(\mathbf{s}_{i}=\mathbf{s}^{\prime}_{i}|\mathbf{m}_{i}=\mathbf{m}^{\prime}_{i},\mathbf{s}_{i-1}=\mathbf{s}^{\prime}_{i-1})=1\text{,} (44)

for 1in/k1\leq i\leq n/k.

Conversely, if 𝐦i𝐦i\mathbf{m}_{i}\neq\mathbf{m}^{\prime}_{i} or 𝐬i1𝐬i1\mathbf{s}_{i-1}\neq\mathbf{s}^{\prime}_{i-1} with a sufficiently large vv, from Property 2 we can obtain that 𝐬i𝐬i\mathbf{s}_{i}\neq\mathbf{s}^{\prime}_{i} for 1in/k1\leq i\leq n/k.

When 𝐌𝒲a\mathbf{M}^{\prime}\in\mathcal{W}_{a}, i.e.i.e., 𝐦i=𝐦i\mathbf{m}_{i}=\mathbf{m}^{\prime}_{i} for 1i<a1\leq i<a and 𝐦a𝐦a\mathbf{m}_{a}\neq\mathbf{m}^{\prime}_{a}, we can iterate to obtain that 𝐬i=𝐬i\mathbf{s}_{i}=\mathbf{s}^{\prime}_{i} for 1i<a1\leq i<a and 𝐬i𝐬i\mathbf{s}_{i}\neq\mathbf{s}^{\prime}_{i} for ain/ka\leq i\leq n/k. Hence after RNGs and uniform constellation mapping, the coded symbols satisfy

f(𝐱i,j(𝐌))=f(𝐱i,j(𝐌)),\operatorname{f}(\mathbf{x}_{i,j}(\mathbf{M}))=\operatorname{f}(\mathbf{x}_{i,j}(\mathbf{M^{\prime}}))\text{,} (45)

where 1i<a1\leq i<a.

Attributed to 𝐬i𝐬i\mathbf{s}_{i}\neq\mathbf{s}^{\prime}_{i} for ain/ka\leq i\leq n/k, f(𝐱i,j(𝐌))\operatorname{f}(\mathbf{x}_{i,j}(\mathbf{M})) and f(𝐱i,j(𝐌))\operatorname{f}(\mathbf{x}_{i,j}(\mathbf{M^{\prime}})) follow the uniform distribution over [0,2c1][0,2^{c}-1], i.d.d (independent and identically distribution).

Appendix B Proof of lemma 1

To simplify the probability Pr(𝐯La(𝐯La+2𝐍La)T0)\mathrm{Pr}\left(\mathbf{v}^{L_{a}}{\left(\mathbf{v}^{L_{a}}+2{\mathbf{N}}^{L_{a}}\right)}^{\mathrm{T}}\leq 0\right) in (18), we introduce a rotation matrix for the LaL_{a}-dimensions hyperspace. The rotation matrix is denoted as

𝐀=[𝐀1𝐀La]La×La,\mathbf{A}=\begin{bmatrix}\mathbf{A}_{1}\\ \vdots\\ \mathbf{A}_{L_{a}}\end{bmatrix}\in\mathbb{R}^{L_{a}\times L_{a}}\text{,} (46)

which satisfies

𝐀[𝐯La]T=[𝐯La,0,,0La1]T.\mathbf{A}{\left[\mathbf{v}^{L_{a}}\right]}^{\mathrm{T}}={\bigg{[}\left\|\mathbf{v}^{L_{a}}\right\|,\underbrace{0,\cdots,0}_{L_{a}-1}\bigg{]}}^{\mathrm{T}}\text{.} (47)

As a rotation matrix, 𝐀\mathbf{A} owns the property that 𝐀T𝐀=𝐈La{\mathbf{A}}^{\mathrm{T}}\mathbf{A}=\mathbf{I}_{L_{a}}, which can be used to rewrite the probability as

Pr(𝐯La(𝐯La+2𝐍La)T0)=Pr(𝐯La𝐈La(𝐯La+2𝐍La)T0)=Pr(𝐯La𝐀T𝐀(𝐯La+2𝐍La)T0)=Pr([𝐀[𝐯La]T]T(𝐀[𝐯La]T+2𝐀[𝐍La]T)0).\begin{split}&\mathrm{Pr}\left(\mathbf{v}^{L_{a}}{\left(\mathbf{v}^{L_{a}}+2{\mathbf{N}}^{L_{a}}\right)}^{\mathrm{T}}\leq 0\right)\\ &=\mathrm{Pr}\left(\mathbf{v}^{L_{a}}\mathbf{I}_{L_{a}}{\left(\mathbf{v}^{L_{a}}+2{\mathbf{N}}^{L_{a}}\right)}^{\mathrm{T}}\leq 0\right)\\ &=\mathrm{Pr}\left(\mathbf{v}^{L_{a}}{\mathbf{A}}^{\mathrm{T}}\mathbf{A}{\left(\mathbf{v}^{L_{a}}+2{\mathbf{N}}^{L_{a}}\right)}^{\mathrm{T}}\leq 0\right)\\ &=\mathrm{Pr}\left({\left[\mathbf{A}{\left[\mathbf{v}^{L_{a}}\right]}^{\mathrm{T}}\right]}^{\mathrm{T}}\left(\mathbf{A}{\left[\mathbf{v}^{L_{a}}\right]}^{\mathrm{T}}+2\mathbf{A}{\left[\mathbf{N}^{L_{a}}\right]}^{\mathrm{T}}\right)\leq 0\right)\text{.}\end{split} (48)

Applying (46) and (47) in (48) results in

Pr(𝐯La2+2𝐯La𝐀1[𝐍La]T0)=Pr(𝐀1[𝐍La]T𝐯La2),\begin{split}&\mathrm{Pr}\left({\left\|\mathbf{v}^{L_{a}}\right\|}^{2}+2\left\|\mathbf{v}^{L_{a}}\right\|\cdot\mathbf{A}_{1}{\left[\mathbf{N}^{L_{a}}\right]}^{\mathrm{T}}\leq 0\right)\\ &=\mathrm{Pr}\left(\mathbf{A}_{1}{\left[\mathbf{N}^{L_{a}}\right]}^{\mathrm{T}}\leq-\frac{\left\|\mathbf{v}^{L_{a}}\right\|}{2}\right)\text{,}\end{split} (49)

the RHS of which is given by the following equivalent form:

Pr(𝐀1[𝐍La]T𝐯La2)=𝐀1[𝐍La]T𝐯La2La1(2πσ)Lae𝐍La22σ2i=an/kj=1Ldni,j.\begin{split}&\mathrm{Pr}\left(\mathbf{A}_{1}{\left[\mathbf{N}^{L_{a}}\right]}^{\mathrm{T}}\leq-\frac{\left\|\mathbf{v}^{L_{a}}\right\|}{2}\right)\\ &=\overbrace{\int\cdots\int}^{L_{a}}_{{}_{\mathbf{A}_{1}{\left[\mathbf{N}^{L_{a}}\right]}^{\mathrm{T}}\leq-\frac{\left\|\mathbf{v}^{L_{a}}\right\|}{2}}}\frac{1}{{\left(\sqrt{2\pi}\sigma\right)}^{L_{a}}}\mathrm{e}^{-\frac{{\left\|\mathbf{N}^{L_{a}}\right\|}^{2}}{2\sigma^{2}}}\prod_{i=a}^{n/k}\prod_{j=1}^{L}\mathrm{d}n_{i,j}\text{.}\end{split} (50)

In order to further simplify the LaL_{a}-fold integral above, we introduce another rotation matrix 𝐁\mathbf{B}, which is denoted as

𝐁=[𝐁1𝐁La]La×La,\mathbf{B}=\begin{bmatrix}\mathbf{B}_{1}\\ \vdots\\ \mathbf{B}_{L_{a}}\end{bmatrix}\in\mathbb{R}^{L_{a}\times L_{a}}\text{,} (51)

such that

𝐁[𝐍La]T=[𝐀1[𝐍La]T,0,,0La1]T,\mathbf{B}{\left[\mathbf{N}^{L_{a}}\right]}^{\mathrm{T}}={\bigg{[}\mathbf{A}_{1}{\left[\mathbf{N}^{L_{a}}\right]}^{\mathrm{T}},\underbrace{0,\cdots,0}_{L_{a}-1}\bigg{]}}^{\mathrm{T}}\text{,} (52)

and that

[𝐍La]T=𝐁T[𝐧La]T,{\left[\mathbf{N}^{L_{a}}\right]}^{\mathrm{T}}=\mathbf{B}^{\mathrm{T}}{\left[\mathbf{n}^{L_{a}}\right]}^{\mathrm{T}}\text{,} (53)

where 𝐧La\mathbf{n}^{L_{a}} is the transformed vector composed of ni,jn^{\prime}_{i,j} with ain/k,1jLa\leq i\leq n/k,1\leq j\leq L.

From (52) we can get that

𝐁1[𝐍La]T=𝐀1[𝐍La]T.\mathbf{B}_{1}{\left[\mathbf{N}^{L_{a}}\right]}^{\mathrm{T}}=\mathbf{A}_{1}{\left[\mathbf{N}^{L_{a}}\right]}^{\mathrm{T}}\text{.} (54)

Substitute (53) into (54), we have that

𝐀1[𝐍La]T=𝐁1[𝐍La]T\displaystyle\mathbf{A}_{1}{\left[\mathbf{N}^{L_{a}}\right]}^{\mathrm{T}}=\mathbf{B}_{1}{\left[\mathbf{N}^{L_{a}}\right]}^{\mathrm{T}}
=𝐁1𝐁T[𝐧La]T\displaystyle=\mathbf{B}_{1}\mathbf{B}^{\mathrm{T}}{\left[\mathbf{n}^{L_{a}}\right]}^{\mathrm{T}}
=[𝐁1𝐁1T𝐁1𝐁2T𝐁1𝐁LaT][𝐧La]T\displaystyle=\begin{bmatrix}\mathbf{B}_{1}\mathbf{B}_{1}^{\mathrm{T}}&\mathbf{B}_{1}\mathbf{B}_{2}^{\mathrm{T}}&\cdots&\mathbf{B}_{1}\mathbf{B}_{L_{a}}^{\mathrm{T}}\end{bmatrix}{\left[\mathbf{n}^{L_{a}}\right]}^{\mathrm{T}}
=(c)[1,0,,0La1][𝐧La]T=na,1,\displaystyle\overset{(c)}{=}\left[1,\underbrace{0,\cdots,0}_{L_{a}-1}\right]{\left[\mathbf{n}^{L_{a}}\right]}^{\mathrm{T}}=n^{\prime}_{a,1}\text{,} (55)

where (c) is obtained from the property that rotation matrix fulfills 𝐁𝐁T=𝐈La\mathbf{B}{\mathbf{B}}^{\mathrm{T}}=\mathbf{I}_{L_{a}}.

From (53) we can also derive that

𝐍La2=[𝐍La]T𝐍La=𝐧La𝐁𝐁T[𝐧La]T=(c)𝐧La2,\begin{split}{\left\|\mathbf{N}^{L_{a}}\right\|}^{2}&={\left[\mathbf{N}^{L_{a}}\right]}^{\mathrm{T}}\mathbf{N}^{L_{a}}\\ &=\mathbf{n}^{L_{a}}\mathbf{B}\mathbf{B}^{\mathrm{T}}{\left[\mathbf{n}^{L_{a}}\right]}^{\mathrm{T}}\overset{(c)}{=}{\left\|\mathbf{n}^{L_{a}}\right\|}^{2}\text{,}\end{split} (56)

where (c) establishes by adopting 𝐁𝐁T=𝐈La\mathbf{B}{\mathbf{B}}^{\mathrm{T}}=\mathbf{I}_{L_{a}}.

Together with (B) and (56), (50) can be rewritten as [22]

𝐀1[𝐍La]T𝐯La2La1(2πσ)Lae𝐍La22σ2i=an/kj=1Ldni,j=na,1𝐯La2La1(2πσ)Lae𝐧La22σ2|𝐉(𝐧La)|i=an/kj=1Ldni,j,\small\begin{split}&\overbrace{\int\cdots\int}^{L_{a}}_{\mathbf{A}_{1}{\left[\mathbf{N}^{L_{a}}\right]}^{\mathrm{T}}\leq-\frac{\left\|\mathbf{v}^{L_{a}}\right\|}{2}}\frac{1}{{\left(\sqrt{2\pi}\sigma\right)}^{L_{a}}}\mathrm{e}^{-\frac{{\left\|\mathbf{N}^{L_{a}}\right\|}^{2}}{2\sigma^{2}}}\prod_{i=a}^{n/k}\prod_{j=1}^{L}\mathrm{d}n_{i,j}\\ &=\overbrace{\int\cdots\int}^{L_{a}}_{n^{\prime}_{a,1}\leq-\frac{\left\|\mathbf{v}^{L_{a}}\right\|}{2}}\frac{1}{{\left(\sqrt{2\pi}\sigma\right)}^{L_{a}}}\mathrm{e}^{-\frac{{\left\|\mathbf{n}^{L_{a}}\right\|}^{2}}{2\sigma^{2}}}\left|\mathbf{J}\left(\mathbf{n}^{L_{a}}\right)\right|\prod_{i=a}^{n/k}\prod_{j=1}^{L}\mathrm{d}n^{\prime}_{i,j}\text{,}\end{split} (57)

in which 𝐉(𝐧La)\mathbf{J}\left(\mathbf{n}^{L_{a}}\right) is the Jacobi matrix, behaving as

𝐉(𝐧La)=[na,1na,1na,1nn/k,Lnn/k,Lna,1nn/k,Lnn/k,L].\mathbf{J}\left(\mathbf{n}^{L_{a}}\right)=\begin{bmatrix}\frac{\partial n^{\prime}_{a,1}}{n_{a,1}}&\cdots&\frac{\partial n^{\prime}_{a,1}}{n_{n/k,L}}\\ \vdots&\ddots&\vdots\\ \frac{\partial n^{\prime}_{n/k,L}}{n_{a,1}}&\cdots&\frac{\partial n^{\prime}_{n/k,L}}{n_{n/k,L}}\end{bmatrix}\text{.} (58)

Since [𝐍La]T=𝐁T[𝐧La]T{\left[\mathbf{N}^{L_{a}}\right]}^{\mathrm{T}}=\mathbf{B}^{\mathrm{T}}{\left[\mathbf{n}^{L_{a}}\right]}^{\mathrm{T}}, we get [𝐧La]T=𝐁[𝐍La]T{\left[\mathbf{n}^{L_{a}}\right]}^{\mathrm{T}}=\mathbf{B}{\left[\mathbf{N}^{L_{a}}\right]}^{\mathrm{T}}, therefore the Jacobi matrix behaves as

𝐉(𝐧La)=𝐁.\mathbf{J}\left(\mathbf{n}^{L_{a}}\right)=\mathbf{B}\text{.} (59)

Recalling that 𝐁𝐁T=𝐈La\mathbf{B}{\mathbf{B}}^{\mathrm{T}}=\mathbf{I}_{L_{a}}, we have

|𝐉(𝐧La)|=|𝐉(𝐧La)|2=|𝐁|2=|𝐁𝐁T|=1.\begin{split}\left|\mathbf{J}\left(\mathbf{n}^{L_{a}}\right)\right|&=\sqrt{{\left|\mathbf{J}\left(\mathbf{n}^{L_{a}}\right)\right|}^{2}}\\ &=\sqrt{{\left|\mathbf{B}\right|}^{2}}=\sqrt{\left|\mathbf{B}{\mathbf{B}}^{\mathrm{T}}\right|}=1\text{.}\end{split} (60)

Then, by substituting (60) into (57), we could verify that

na,1𝐯La2La1(2πσ)Lae𝐧La22σ2|𝐉(𝐧La)|i=an/kj=1Ldni,j=na,1𝐯La2La1(2πσ)Lae𝐧La22σ2i=an/kj=1Ldni,j=𝐯La212πσe(na,1)22σ2dna,1=Q(𝐯La2σ).\small\begin{split}&\overbrace{\int\cdots\int}^{L_{a}}_{n^{\prime}_{a,1}\leq-\frac{\left\|\mathbf{v}^{L_{a}}\right\|}{2}}\frac{1}{{\left(\sqrt{2\pi}\sigma\right)}^{L_{a}}}\mathrm{e}^{-\frac{{\left\|\mathbf{n}^{L_{a}}\right\|}^{2}}{2\sigma^{2}}}\left|\mathbf{J}\left(\mathbf{n}^{L_{a}}\right)\right|\prod_{i=a}^{n/k}\prod_{j=1}^{L}\mathrm{d}n^{\prime}_{i,j}\\ &=\overbrace{\int\cdots\int}^{L_{a}}_{n^{\prime}_{a,1}\leq-\frac{\left\|\mathbf{v}^{L_{a}}\right\|}{2}}\frac{1}{{\left(\sqrt{2\pi}\sigma\right)}^{L_{a}}}\mathrm{e}^{-\frac{{\left\|\mathbf{n}^{L_{a}}\right\|}^{2}}{2\sigma^{2}}}\prod_{i=a}^{n/k}\prod_{j=1}^{L}\mathrm{d}n^{\prime}_{i,j}\\ &=\int_{-\infty}^{-\frac{\left\|\mathbf{v}^{L_{a}}\right\|}{2}}\frac{1}{{\sqrt{2\pi}\sigma}}\mathrm{e}^{-\frac{{\left(n^{\prime}_{a,1}\right)}^{2}}{2\sigma^{2}}}\mathrm{d}n^{\prime}_{a,1}=Q\left(\frac{\left\|\mathbf{v}^{L_{a}}\right\|}{2\sigma}\right)\text{.}\end{split} (61)

Appendix C Proof of Lemma 2

The central problem is to solve out the integral of hh in the RHS of (26). Recall that hh follows Rayleigh distribution, whose PDF is given as

g1(h)=2hΩexp(h2Ω).g_{{}_{1}}\left(h\right)=\frac{2h}{\Omega}\mathrm{exp}\left(\frac{-h^{2}}{\Omega}\right)\text{.} (62)

Insert (62) into the RHS of (26), the integral is transformed as

hexp(h2u28σ2sin2θ)g1(h)dh=0+exp(h2u28σ2sin2θ)2hΩexp(h2Ω)dh\xlongequalz=u28σ2sin2θ1Ω0+exp[(z+1Ω)h2]dh2=1zΩ+1=8σ2sin2θΩu2+8σ2sin2θ,\begin{split}&\int_{h}{\mathrm{exp}}{\left(-\frac{{h^{2}}{u^{2}}}{{8}{\sigma^{2}}{\sin^{2}}\theta}\right)}g_{{}_{1}}(h)\mathrm{d}h\\ &=\int_{0}^{+\infty}{\mathrm{exp}}{\left(-\frac{{h^{2}}{u^{2}}}{{8}{\sigma^{2}}{\sin^{2}}\theta}\right)}\cdot\frac{2h}{\Omega}\mathrm{exp}\left(\frac{-h^{2}}{\Omega}\right)\mathrm{d}h\\ &\xlongequal{z=\frac{u^{2}}{8{{\sigma}^{2}}{{\sin}^{2}}\theta}}\frac{1}{\Omega}\int_{0}^{+\infty}\mathrm{exp}\left[-\left(z+\frac{1}{\Omega}\right){h^{2}}\right]\mathrm{d}{h^{2}}\\ &=\frac{1}{z\Omega+1}=\frac{8{\sigma^{2}}{\sin^{2}}\theta}{{\Omega}u^{2}+8{\sigma^{2}}{\sin^{2}}\theta}\text{,}\end{split} (63)

with which the RHS of (26) becomes

upU(u)hexp(h2u28σ2sin2θ)g1(h)dh=upU(u)8σ2sin2θΩu2+8σ2sin2θ.\begin{split}&\sum_{u}p_{{}_{U}}(u)\int_{h}{\mathrm{exp}}{\left(-\frac{{h^{2}}{u^{2}}}{{8}{\sigma^{2}}{\sin^{2}}\theta}\right)}g_{1}(h)\mathrm{d}h\\ &=\sum_{u}p_{{}_{U}}(u)\cdot\frac{8{\sigma^{2}}{\sin^{2}}\theta}{{\Omega}u^{2}+8{\sigma^{2}}{\sin^{2}}\theta}\text{.}\end{split} (64)

Recall that U=Ua,1=f(𝐱a,1(𝐌))f(𝐱a,1(𝐌))U=U_{a,1}=\operatorname{f}(\mathbf{x}_{a,1}(\mathbf{M}))-\operatorname{f}(\mathbf{x}_{a,1}(\mathbf{M^{\prime}})). Appendix A draws the conclusion that both f(𝐱a,1(𝐌))\operatorname{f}(\mathbf{x}_{a,1}(\mathbf{M})) and f(𝐱a,1(𝐌))\operatorname{f}(\mathbf{x}_{a,1}(\mathbf{M^{\prime}})) follow the uniform distribution over [0,2c1][0,2^{c}-1], i.d.d. As a result, (64) extends to

upU(u)8σ2sin2θΩu2+8σ2sin2θ\xlongequal[Y=f(𝐱a,1(𝐌))]X=f(𝐱a,1(𝐌))iΨjΨpX(i)pY(j)8σ2sin2θΩ(ij)2+8σ2sin2θ=iΨjΨ22c8σ2sin2θΩ(ij)2+8σ2sin2θ.\begin{split}&\sum_{u}p_{{}_{U}}(u)\cdot\frac{8{\sigma^{2}}{\sin^{2}}\theta}{{\Omega}u^{2}+8{\sigma^{2}}{\sin^{2}}\theta}\\ &\xlongequal[Y=\operatorname{f}(\mathbf{x}_{a,1}(\mathbf{M^{\prime}}))]{X=\operatorname{f}(\mathbf{x}_{a,1}(\mathbf{M}))}\sum_{i\in\Psi}\sum_{j\in\Psi}p_{{}_{X}}(i)p_{{}_{Y}}(j)\frac{8{\sigma}^{2}{\sin}^{2}{\theta}}{{\Omega}(i-j)^{2}+8{\sigma}^{2}{\sin}^{2}{\theta}}\\ &=\sum_{i\in\Psi}\sum_{j\in\Psi}2^{-2c}\frac{8{\sigma}^{2}{\sin}^{2}{\theta}}{{\Omega}(i-j)^{2}+8{\sigma}^{2}{\sin}^{2}{\theta}}\text{.}\end{split} (65)

Appendix D Proof Of Monotony

Rayleigh(θ;σ,La)=(iΨjΨ22c8σ2sin2θΩ(ij)2+8σ2sin2θ)La.\small\mathcal{F}_{\text{Rayleigh}}(\theta\text{;}\sigma\text{,}L_{a})={\left(\sum_{i\in\Psi}\sum_{j\in\Psi}2^{-2c}\frac{8{\sigma}^{2}{\sin}^{2}{\theta}}{{\Omega}(i-j)^{2}+8{\sigma}^{2}{\sin}^{2}{\theta}}\right)}^{L_{a}}\text{.} (66)

The monotony of Rayleigh(θ;σ,La)\mathcal{F}_{\text{Rayleigh}}(\theta\text{;}\sigma\text{,}L_{a}) equals to the monotony of

FRayleigh(θ)8σ2sin2θΩ(ij)2+8σ2sin2θ,F_{\text{Rayleigh}}(\theta)\triangleq\frac{8{\sigma}^{2}{\sin}^{2}{\theta}}{{\Omega}(i-j)^{2}+8{\sigma}^{2}{\sin}^{2}{\theta}}\text{,} (67)

where θ[0,π2]\theta\in[0,\frac{\pi}{2}].

Denoting 8σ2=b8{\sigma}^{2}=b and Ω(ij)2=k{\Omega}(i-j)^{2}=k, the derivative of F(θ)F(\theta) is given by

dFRayleigh(θ)dθ=2kbsinθcosθ(k+sin2θ)20,\frac{\mathrm{d}F_{\text{Rayleigh}}(\theta)}{\mathrm{d}\theta}=2kb\frac{\mathrm{sin}\theta\mathrm{cos}\theta}{{(k+\mathrm{sin}^{2}\theta)}^{2}}\geq 0\text{,} (68)

which indicates the monotonically increasing property of function FRayleigh(θ)F_{\text{Rayleigh}}(\theta).

As a consequence, Rayleigh(θ;σ,La)\mathcal{F}_{\text{Rayleigh}}(\theta\text{;}\sigma\text{,}L_{a}) is a monotonically increasing function.

Nakagami(θ;σ,La)=(iΨjΨ22c(8mσ2sin2θΩ(ij)2+8mσ2sin2θ)m)La.\begin{split}&\mathcal{F}_{\text{Nakagami}}(\theta\text{;}\sigma\text{,}L_{a})\\ &={\left(\sum_{i\in\Psi}\sum_{j\in\Psi}2^{-2c}{\left(\frac{8m{\sigma}^{2}{\sin}^{2}{\theta}}{{\Omega}(i-j)^{2}+8m{\sigma}^{2}{\sin}^{2}{\theta}}\right)}^{m}\right)}^{L_{a}}\text{.}\end{split} (69)

Because m0.5m\geq 0.5, the monotony of Nakagami(θ;σ,La)\mathcal{F}_{\text{Nakagami}}(\theta\text{;}\sigma\text{,}L_{a}) equals to the monotony of

FNakagami(θ)8mσ2sin2θΩ(ij)2+8mσ2sin2θ,F_{\text{Nakagami}}(\theta)\triangleq\frac{8m{\sigma}^{2}{\sin}^{2}{\theta}}{{\Omega}(i-j)^{2}+8m{\sigma}^{2}{\sin}^{2}{\theta}}\textbf{,} (70)

where θ[0,π2]\theta\in[0,\frac{\pi}{2}].

Similar to the proof for Rayleigh(θ;σ,La)\mathcal{F}_{\text{Rayleigh}}(\theta\text{;}\sigma\text{,}L_{a}), we denote that 8mσ2=b8m{\sigma}^{2}=b and Ω(ij)2=k{\Omega}(i-j)^{2}=k. Since the result is the same as (68), Nakagami(θ;σ,La)\mathcal{F}_{\text{Nakagami}}(\theta\text{;}\sigma\text{,}L_{a}) is also a monotonically increasing function.

Rician(θ;σ,La)=(iΨjΨ22c8(K+1)σ2sin2θΩ(ij)2+8(K+1)σ2sin2θexp(8K(K+1)σ2sin2θΩ(ij)2+8(K+1)σ2sin2θK))La,\small\begin{split}&\mathcal{F}_{\text{Rician}}(\theta\text{;}\sigma\text{,}L_{a})\\ &=\Bigg{(}\sum_{i\in\Psi}\sum_{j\in\Psi}2^{-2c}\frac{8(K+1){\sigma}^{2}{\sin}^{2}{\theta}}{{\Omega}(i-j)^{2}+8(K+1){\sigma}^{2}{\sin}^{2}{\theta}}\\ &\qquad\cdot{\mathrm{exp}\left(\frac{8K(K+1){\sigma}^{2}{\sin}^{2}{\theta}}{{\Omega}(i-j)^{2}+8(K+1){\sigma}^{2}{\sin}^{2}{\theta}}-K\right)\Bigg{)}}^{L_{a}}\text{,}\end{split} (71)

Because exponential function is monotonically increasing, the monotony of Rician(θ;σ,La)\mathcal{F}_{\text{Rician}}(\theta\text{;}\sigma\text{,}L_{a}) equals to the monotony of

FRician(θ)8(K+1)σ2sin2θΩ(ij)2+8(K+1)σ2sin2θ.F_{\text{Rician}}(\theta)\triangleq\frac{8(K+1){\sigma}^{2}{\sin}^{2}{\theta}}{{\Omega}(i-j)^{2}+8(K+1){\sigma}^{2}{\sin}^{2}{\theta}}\text{.} (72)

where θ[0,π2]\theta\in[0,\frac{\pi}{2}].

Similar to the proof for Rayleigh(θ;σ,La)\mathcal{F}_{\text{Rayleigh}}(\theta\text{;}\sigma\text{,}L_{a}), we denote 8(K+1)σ2=b8(K+1){\sigma}^{2}=b and Ω(ij)2=k{\Omega}(i-j)^{2}=k. As the result is the same as (68),Rician(θ;σ,La)(\ref{derivation}),\mathcal{F}_{\text{Rician}}(\theta\text{;}\sigma\text{,}L_{a}) is also a monotonically increasing function.

Appendix E Proof of Theorem 2

For Spinal codes over the Nakagami-m fading channel, the only difference is to replace Rayleigh distribution with Nakagami-m distribution. The PDF Nakagami-m distribution is given by

g2(h)=2mmΓ(m)Ωmh2m1exp(mh2Ω),g_{{}_{2}}(h)=\frac{2m^{m}}{\Gamma(m)\Omega^{m}}h^{2m-1}\text{exp}\left(\frac{-mh^{2}}{\Omega}\right)\text{,} (73)

where hh is the channel fading parameter, m0.5m\geq 0.5 and the Gamma function is defined as

Γ(m)=0+ettm1dt.\Gamma(m)={\int_{0}^{+\infty}}\mathrm{e}^{-t}t^{m-1}\mathrm{d}t\text{.} (74)

Substitute (73) into the integral of (26) and denote z=u28σ2sin2θz=\frac{u^{2}}{8{{\sigma}^{2}}{{\sin}^{2}}\theta}, after which the integral of hh is transformed as

hexp(h2u28σ2sin2θ)g2(h)dh=0+ezh22mmΓ(m)Ωmh2m1emh2Ωdh\xlongequal[h=tz+mΩ]t=(z+mΩ)h2mmΓ(m)(zΩ+m)m0+ettm1dt=mm(zΩ+m)m.\begin{split}&\int_{h}{\mathrm{exp}}{\left(-\frac{{h^{2}}{u^{2}}}{{8}{\sigma^{2}}{\sin^{2}}\theta}\right)}g_{{}_{2}}(h)\mathrm{d}h\\ &=\int_{0}^{+\infty}\mathrm{e}^{-{z}{h^{2}}}\cdot\frac{2m^{m}}{\Gamma(m)\Omega^{m}}h^{2m-1}\mathrm{e}^{-\frac{mh^{2}}{\Omega}}\mathrm{d}h\\ &\xlongequal[h=\sqrt{\frac{t}{z+\frac{m}{\Omega}}}]{t=\left(z+\frac{m}{\Omega}\right)h^{2}}\frac{m^{m}}{\Gamma(m){(z\Omega+m)}^{m}}{\int_{0}^{+\infty}}\mathrm{e}^{-t}t^{m-1}\mathrm{d}t\\ &=\frac{m^{m}}{{(z\Omega+m)}^{m}}\text{.}\end{split} (75)

As mentioned above, z=u28σ2sin2θz=\frac{u^{2}}{8{{\sigma}^{2}}{{\sin}^{2}}\theta}, substitute zz into (LABEL:eq48):

mm(zΩ+m)m=(8mσ2sin2θΩu2+8mσ2sin2θ)m.\begin{split}\frac{m^{m}}{{(z\Omega+m)}^{m}}={\left(\frac{8m{\sigma}^{2}{\sin}^{2}{\theta}}{{\Omega}u^{2}+8m{\sigma}^{2}{\sin}^{2}{\theta}}\right)}^{m}\text{.}\end{split} (76)

Then the process is similar to Theorem 1, by denoting that

Nakagami(θ;σ,La)=(iΨjΨ22c(8mσ2sin2θΩ(ij)2+8mσ2sin2θ)m)La,\begin{split}&\mathcal{F}_{\text{Nakagami}}(\theta\text{;}\sigma\text{,}L_{a})\\ &={\left(\sum_{i\in\Psi}\sum_{j\in\Psi}2^{-2c}{\left(\frac{8m{\sigma}^{2}{\sin}^{2}{\theta}}{{\Omega}(i-j)^{2}+8m{\sigma}^{2}{\sin}^{2}{\theta}}\right)}^{m}\right)}^{L_{a}}\text{,}\end{split} (77)

which is monotonically increasing proved in Appendix D, and that

Nakagami(La,σ)=r=1NbrNakagami(θr;σ,La),\mathscr{F}_{\text{Nakagami}}\left(L_{a}\text{,}\sigma\right)=\sum_{r=1}^{N}b_{r}\mathcal{F}_{\text{Nakagami}}(\theta_{r}\text{;}\sigma\text{,}L_{a})\text{,} (78)

we can get the explicit bound.

Appendix F Proof of Theorem 3

For Spinal codes over the Rician fading channel, the only difference is to replace Rayleigh distribution with Rician distribution. The PDF of Rician distribution is given by

g3(h)=2(K+1)hΩexp(K+(K+1)h2Ω)I0(2K(K+1)Ωh),\small g_{{}_{3}}(h)=\frac{2(K+1)h}{{\Omega}\,{\mathrm{exp}\left(K+\frac{(K+1){h^{2}}}{\Omega}\right)}}I_{0}\left(2\sqrt{\frac{K(K+1)}{\Omega}}h\right)\text{,} (79)

where hh is the channel fading parameter, and I0()I_{0}(\cdot) is the modified Bessel function of the first kind with order zero, given by

I0(x)=k=0+(x2/4)kk!Γ(k+1).I_{0}(x)=\sum_{k=0}^{+\infty}\frac{{(x^{2}/4)}^{k}}{k!\Gamma(k+1)}\text{.} (80)

Substitute (79) into (26) and denote z=u28σ2sin2θz=\frac{u^{2}}{8{{\sigma}^{2}}{{\sin}^{2}}\theta}, then the integral of hh is transformed as

hexp(h2u28σ2sin2θ)g3(h)dh=0+ezh2g3(h)dh=2(K+1)ΩeK0+he(K+1Ω+z)h2I0(2K(K+1)Ωh)dh=2eKm=0+Km(K+1)m+1m!Γ(m+1)Ωm+10+h2m+1e(K+1Ω+z)h2dh.\small\begin{split}&\int_{h}{\mathrm{exp}}{\left(-\frac{{h^{2}}{u^{2}}}{{8}{\sigma^{2}}{\sin^{2}}\theta}\right)}g_{{}_{3}}(h)\mathrm{d}h=\int_{0}^{+\infty}\mathrm{e}^{-{z}{h^{2}}}g_{{}_{3}}(h)\mathrm{d}h\\ &=\frac{2(K+1)}{{\Omega}\mathrm{e}^{K}}{\int_{0}^{+\infty}}{h\mathrm{e}^{-\left(\frac{K+1}{\Omega}+z\right)h^{2}}I_{0}\left(2\sqrt{\frac{K(K+1)}{\Omega}}h\right)\mathrm{d}h}\\ &=\frac{2}{\mathrm{e}^{K}}\sum_{m=0}^{+\infty}\frac{K^{m}{(K+1)}^{m+1}}{m!\Gamma(m+1)\Omega^{m+1}}{\int_{0}^{+\infty}}h^{2m+1}\mathrm{e}^{-\left(\frac{K+1}{\Omega}+z\right)h^{2}}\mathrm{d}h\text{.}\end{split} (81)

Let t=(z+K+1Ω)h2t=\left(z+\frac{K+1}{\Omega}\right)h^{2}, h=tz+K+1Ωh=\sqrt{\frac{t}{z+\frac{K+1}{\Omega}}}, we have

0+h2m+1e(K+1Ω+z)h2dh=0+ettm2(z+K+1Ω)m+1dt=Γ(m+1)2(z+K+1Ω)m+1.\begin{split}&{\int_{0}^{+\infty}}h^{2m+1}\mathrm{e}^{-\left(\frac{K+1}{\Omega}+z\right)h^{2}}\mathrm{d}h\\ &=\int_{0}^{+\infty}\mathrm{e}^{-t}\cdot\frac{t^{m}}{2{\left(z+\frac{K+1}{\Omega}\right)}^{m+1}}\mathrm{d}t\\ &=\frac{\Gamma(m+1)}{2{\left(z+\frac{K+1}{\Omega}\right)}^{m+1}}\text{.}\end{split} (82)

With (LABEL:eq59) in hand, (LABEL:eq58) can be rewritten as

hexp(h2u28σ2sin2θ)g3(h)dh=1KeKK(K+1)Ωz+K+1m=0+1m!(K(K+1)Ωz+K+1)m=1KeKK(K+1)Ωz+K+1exp(K(K+1)Ωz+K+1).\begin{split}&\int_{h}{\mathrm{exp}}{\left(-\frac{{h^{2}}{u^{2}}}{{8}{\sigma^{2}}{\sin^{2}}\theta}\right)}g_{{}_{3}}(h)\mathrm{d}h\\ &=\frac{1}{K\mathrm{e}^{K}}\cdot\frac{K(K+1)}{{\Omega}z+K+1}\cdot\sum_{m=0}^{+\infty}\frac{1}{m!}{\left(\frac{K(K+1)}{{\Omega}z+K+1}\right)}^{m}\\ &=\frac{1}{K\mathrm{e}^{K}}\cdot\frac{K(K+1)}{{\Omega}z+K+1}\cdot\mathrm{exp}\left(\frac{K(K+1)}{{\Omega}z+K+1}\right)\text{.}\end{split} (83)

Now that z=u28σ2sin2θz=\frac{u^{2}}{8{{\sigma}^{2}}{{\sin}^{2}}\theta}, (83) comes into

1KeKK(K+1)Ωz+K+1exp(K(K+1)Ωz+K+1)=8eK(K+1)σ2sin2θΩu2+8(K+1)σ2sin2θexp(8K(K+1)σ2sin2θΩu2+8(K+1)σ2sin2θ).\small\begin{split}&\frac{1}{K\mathrm{e}^{K}}\cdot\frac{K(K+1)}{{\Omega}z+K+1}\cdot\mathrm{exp}\left(\frac{K(K+1)}{{\Omega}z+K+1}\right)\\ &=\frac{8\mathrm{e}^{-K}(K+1){\sigma^{2}}{\sin^{2}}\theta}{{\Omega}u^{2}+8(K+1){\sigma^{2}}{\sin^{2}}\theta}\mathrm{exp}\left(\frac{8K(K+1){\sigma^{2}}{\sin^{2}}\theta}{{\Omega}u^{2}+8(K+1){\sigma^{2}}{\sin^{2}}\theta}\right)\text{.}\end{split} (84)

Then the process is similar to Theorem 1. By denoting that

Rician(θ;σ,La)=(iΨjΨ22c8(K+1)σ2sin2θΩ(ij)2+8(K+1)σ2sin2θexp(8K(K+1)σ2sin2θΩ(ij)2+8(K+1)σ2sin2θK))La,\small\begin{split}&\mathcal{F}_{\text{Rician}}(\theta\text{;}\sigma\text{,}L_{a})\\ &=\Bigg{(}\sum_{i\in\Psi}\sum_{j\in\Psi}2^{-2c}\frac{8(K+1){\sigma}^{2}{\sin}^{2}{\theta}}{{\Omega}(i-j)^{2}+8(K+1){\sigma}^{2}{\sin}^{2}{\theta}}\\ &\qquad\cdot{\mathrm{exp}\left(\frac{8K(K+1){\sigma}^{2}{\sin}^{2}{\theta}}{{\Omega}(i-j)^{2}+8(K+1){\sigma}^{2}{\sin}^{2}{\theta}}-K\right)\Bigg{)}}^{L_{a}}\text{,}\end{split} (85)

which is monotonically increasing proved in Appendix D, and that

Rician(La,σ)=r=1NbrRician(θr;σ,La),\mathscr{F}_{\text{Rician}}\left(L_{a}\text{,}\sigma\right)=\sum_{r=1}^{N}b_{r}\mathcal{F}_{\text{Rician}}(\theta_{r}\text{;}\sigma\text{,}L_{a})\text{,} (86)

we can get the explicit bound.