Generalized Union Bound for Space-Time Codes Cong Ling Department of Electronic Engineering, King’s College London (email: [email protected]) Abstract⎯Gallager’s second bounding technique, also known as the generalized union bound, is employed to derive a new upper bound on the error probability of space-time codes with maximum-likelihood (ML) decoding on quasi-static Rayleigh fading channels. The new bound is distinguished by two characteristics: unlike the classical union bound, the new bound is rapidly convergent and is only a few decibels away from simulation results; compared to Gallager’s first bound, it has better computational efficiency and numerical stability. Hence, the new bound is a useful tool for performance analysis and computer search of good space-time codes. Moreover, the correlation between fading coefficients is easily accommodated by the new bound. The application of the new bound to convolutional coding on block fading channels is also demonstrated, and an improved version is derived for the bit error probability of maximum a posteriori probability (MAP) decoding.

1

Index Terms⎯Gallager bounds, optimization, space-time coding, union bound, weight spectrum.

I. INTRODUCTION A typical scenario in the study of space-time coding assumes a channel changing so slowly that the fading coefficients can be modeled as static during an entire frame. The standard union bound is often found to be divergent on such quasi-static fading channels. This is because the pairwise error probability decreases at a polynomial rate limited by the diversity order of the channel, which will be overwhelmed by the typically exponential growth of the weight spectrum. Expurgating some codewords from the weight spectrum can alleviate the divergence problem to some degree but is not

This work was supported in part by the National Science Foundation of China under Grant 60402026.

1

adequate. Consequently, the “limit-before-average” (LBA) technique [1], originally developed for convolutional codes on block fading channels, has been applied to space-time codes [2], [3], [4]. Despite its tightness, this technique is analytically not very attractive in that it involves a multi-dimensional integral. The multi-integral is usually evaluated by using the Monte Carlo method, which can be lengthy especially for low error probabilities. In [5], [6], computationally efficient upper bounds for space-time codes were derived on the basis of Gallager’s first bounding technique. In particular, the ellipsoidal bound of [6] is rather tight. The idea is to reduce the computation of the lengthy multi-integral to an optimization problem, for which efficient solutions are known to be existent. This philosophy is akin to that of the classical Chernoff bound, in which a one-dimensional integral is replaced by the optimization of a single parameter [7]. Not only does this reduce the computation complexity, it also gives more insights into the problem. Following the line of thought initiated in [6], this paper continues to develop a new upper bound for space-time codes by exploiting Gallager’s second bounding technique, sometimes referred to as the generalized union bound [8]. The new bound has the salient feature that it only involves the computation of matrix determinants in the optimization process. Note that the ellipsoidal bound requires the computation of residues, which tends to be numerically unstable for high multiplicities of poles [6]. Namely, the new bound is numerically more stable and can be evaluated faster, although it is not as tight as the ellipsoidal bound. Moreover, the bound readily takes into account the case of spatially correlated fading. Therefore, the new bound offers an alternative to existing analytic tools for space-time codes. Such tools are useful in the design or search of good codes. The remainder of this paper is organized as follows. Section II sets up the system model of space-time coding in quasi-static fading channels and briefly reviews existing upper bounds. In Section III, the new bound is derived for the frame error rate (FER) and for the bit error rate (BER). Section IV demonstrates its application to convolutional coding over block fading channels and further tightens the bound on the BER of systematic codes. Numerical results are given in Section V to assess the tightness and convergence properties of the bound. Finally, conclusions are drawn in

2

Section VI.

II. SYSTEM MODEL AND PREVIOUS WORK Consider a space-time coded system with nT transmit antennas and nR receive antennas as depicted in Fig. 1. At each time slot t, nT parallel output symbols of the trellis encoder ct1 , ct2 ,

, ctnT

are simultaneously transmitted from nT antennas. Following the convention [9], it is assumed that the elements of the signal constellations are scaled by a factor

Es

so that the average energy of

the constellation is unity. Suppose that the frame length is L time slots. The space-time code matrix of size nT × L is defined as ⎡ c11 ⎢ 2 c c=⎢ 1 ⎢ ⎢ nT ⎣⎢c1

c12 c22 c2nT

c1L ⎤ ⎥ cL2 ⎥ . ⎥ ⎥ cLnT ⎦⎥

The received signal at the jth receive antenna is a noisy superposition of nT transmitted signals corrupted by fading, given by nT

yt = Es ∑ α ti , j cti + ηt j j

(1)

i =1

where the complex additive white Gaussian noise (AWGN) ηt j has zero mean and variance N 0 / 2 per dimension, and α ti , j is the fading gain from transmit antenna i to receive antenna j, modeled as complex Gaussian with zero mean and variance 0.5 per dimension. Accordingly γ s

Es / N 0 is the

symbol SNR per transmit antenna (and the symbol SNR per receive antenna is given by nT γ s .) A maximum-likelihood (ML) decoder (Viterbi algorithm) is used at the receiver. Here we focus on quasi-static fading, i.e., the gain α ti , j = α i , j remains constant during a frame and varies independently from one frame to another. However the model accommodates spatial correlation between fading coefficients. In matrix form, the signal model can be expressed as 3

y = Es αc + η

(2)

where y is the nR × L received signal matrix obtained by stacking ytj , η is the nR × L noise matrix obtained by stacking ηt j , for j = 1, 2, …, nR and t = 1, 2, …, L, and α is the fading matrix ⎡ α 1,1 ⎢ 1,2 α α=⎢ ⎢ ⎢ 1, nR ⎣α

α 2,1 α 2,2

α n ,1 ⎤ ⎥ α n ,2 ⎥ T

T

⎥ ⎥ α nT ,nR ⎦

α 2,n

R

.

Given the fading realization α, the pairwise error probability that the decoder decides another codeword cˆ while c being actually sent is given by

⎛ γ ⎞ Pe (c, cˆ | α ) = Q ⎜⎜ s d 2 (c, cˆ | α ) ⎟⎟ ⎝ 2 ⎠

(3)

where L

nR

nT

2

d (c, cˆ | α ) = ∑∑ ∑ α (c − cˆ ) =tr ( αCα H ) 2

i, j

i t

t =1 j =1 i =1

i t

(4)

represents the squared Euclidean distance for given α. In (4), the codeword-difference correlation matrix C is defined as C

B (c, cˆ) B H (c, cˆ ),

c − cˆ .

B (c, cˆ)

For uncorrelated fading, d 2 (c, cˆ | α ) can be rewritten as [9] nR

nT

d 2 (c, cˆ | α ) = ∑∑ λi β i , j

2

(5)

j =1 i =1

where β i , j ’s are independent and identically distributed (iid) complex Gaussian random variables with zero mean and variance 0.5 per dimension, and λi for i = 1, 2, …, nT are eigenvalues of C. The first-event error probability PE is the probability of having a decoding error event2 starting from a given time epoch (say t = 0) [10], [11]. One can bound the frame error rate Pf as [10], [12] Pf ≤ LPE . 2

An error event occurs when the decoder leaves the correct path and then returns.

4

By extending [11, Eq. (6.9)] to space-time codes, one obtains the union bound on the first-event error probability PE (α ) ≤

1

c∈

cˆ ≠ c

conditioned on fading gain matrix α, where with cardinality

⎛ γs 2 ⎞ d (c, cˆ | α ) ⎟⎟ , ⎝ 2 ⎠

∑∑ Q ⎜⎜

(6)

is the set of codewords starting at time epoch zero,

, and cˆ − c runs over first error events diverging from c at this epoch. Eq. (6)

can be rearranged into ⎛ γ ⎞ PE (α ) ≤ ∑ AC Q ⎜⎜ s tr ( αCα H ) ⎟⎟ C ⎝ 2 ⎠

(7)

where AC denotes the average multiplicity of first error events having codeword-difference correlation matrix C. The set of pairs {C, AC} is referred to as the weight spectrum of a space-time code in this paper. The weight spectrum is computable in a number of ways, e.g., [11], [13], [14]. The standard union bound then proceeds to average (7) over the fading coefficients, which, due to (5), leads to an expression in terms of the eigenvalue spectrum {λ, Aλ} of the code. That is, the standard union bound is completely determined by the eigenvalue spectrum of a code. However, this classical approach often leads to a divergent upper bound on quasi-static fading channels. A tight bound is obtained by observing that the error rate never exceeds unity; so the union bound condition on α can be limited before averaging over fading coefficients [1]. Mathematically, the LBA bound reads

Pf ≤ ∫ min [1, LPE (α ) ] f α (α )dα

(8)

where f α (α ) stands for the probability density function (pdf) of fading coefficients. Due to the nonlinearity arising from the operation min(1, x), the above bound does not admit a closed form, but has to be numerically evaluated. This difficulty was circumvented in [5], [6] by making use of the following inequality, which originated in Gallager [15] and Fano [16] for Gaussian channels, and is called Gallager’s first bounding technique in literature,

5

Prob(error) ≤ Prob(error, α ∉ R ) + Prob(α ∈ R ) where R is the Gallager region. In the context of space-time codes, R represents the state of deep fading. Then one apply the standard union bound outside of R , yielding the upper bound Pf ≤ L ∫ PE (α ) f α (α )dα + ∫ f α (α )dα . R

R

Note that the LBA technique is tightest in this framework (but without closed form).

{

}

An ellipsoid R E = α : tr ( αFα H ) ≤ 1

(F

0 positive definite) is chosen as the Gallager

region in [6]. This leads to the ellipsoidal bound on the frame error rate: ⎧ π /2 γC 1 ⎪ Pf ≤ min ⎨− L ∑ AC ∫ I + s 2 F 0 4sin θ π 0 C ⎪⎩

−1 ⎡ γ C ⎛ ⎞ −1 s s Res ⎢ s e I + s ⎜ I + F ∑ 4sin 2 θ ⎠⎟ ⎢ ⎝ k =1 ⎣

− nR K 1

K2

+ 1 + ∑ Res ⎡ s −1e s I + sF ⎣ k =1

− nR

{

To speed up optimization, a sphere R S = α : tr ( αα H ) ≤ r 2

− nR

⎤ , s1k < 0 ⎥dθ ⎥ ⎦ (9)

⎫ , sk2 < 0 ⎤ ⎬ . ⎦⎭

}

can be used, which simplifies to the

spherical bound 2 π /2 K ⎧⎪ ⎡ e sr 1 Pf ≤ min ⎨− L ∑ Aλ ∫ ∑ Res ⎢ r >0 π 0 k =1 λ ⎢⎣ s ⎩⎪

λiγ s ⎞ ⎛ ∏ ⎜ s +1+ ⎟ 4sin 2 θ ⎠ i =1 ⎝ nT

− nR

nR nT 2 n ⎫ ⎤ 2 r ⎪ , sk < 0 ⎥dθ + 1 − e − r ∑ ⎬. n=0 n ! ⎪ ⎥⎦ ⎭

(10)

III. GENERALIZED UNION BOUND In this section, Gallager’s second bounding technique is employed to derive a new upper bound for space-time codes, which allows for easier evaluation than the LBA bound and Gallager’s first bound. Gallager originally presented this technique in the context of error exponents for random coding [8], which differed from his first bounding technique proposed for concrete codes and particularly for low-density parity-check (LDPC) codes [15]. It was until 1998 that Duman and Salehi extended the second bound to concrete codes and to turbo codes in particular [17]. For recent studies on Gallager bounds see [18], [19], [20]. A scrutiny into Gallager’s second bounding technique reveals that the essential idea is the

6

following upper bound on the probability of a union ρ

P ( ∪ Ai ) ≤ ⎡⎣ ∑ P( Ai ) ⎤⎦ , 0 ≤ ρ ≤ 1

(11)

which is also referred to as the generalized union bound [18]. Obviously it will degenerate into the standard union bound if ρ = 1; on the other hand it will be reduced to the trivial upper bound P ( ∪ Ai ) ≤ 1 if ρ = 0.

By this inequality one has Pf = ∫ P ( Error | α ) f α (α )dα ≤ ∫ [ LPE (α )] f α (α )dα . ρ

(12)

This can be derived alternatively from the LBA bound by observing that

min [1, LPE (α ) ] ≤ [ LPE (α ) ] , 0 ≤ ρ ≤ 1 . ρ

Yet, the expression as reads in (12) is intractable for further analysis, for the power of ρ is raised on a sum of first-event error probabilities. To get around this obstacle, it is necessary to loosen the bound (12) by applying the Jensen inequality. Letting ψ (α ) be an arbitrary pdf of α, the above bound can be rewritten as Pf ≤ ∫ψ (α )ψ −1 (α ) [ LPE (α )] f α (α )dα = ∫ψ (α ) {ψ −1/ ρ (α ) [ LPE (α ) ] f α1/ ρ (α )} dα. ρ

ρ

(13)

Invoking the Jensen inequality E[ x ρ ] ≤ ( E[ x]) ρ for 0 ≤ ρ ≤ 1, x ≥ 0 [21] yields the desirable form: Pf ≤ min

0≤ ρ ≤1

{

}

ρ

1−1/ ρ 1/ ρ ∫ψ (α) [ LPE (α)] fα (α)dα .

(14)

This weakened bound (due to the use of Jensen’s inequality) retains the nice property that it is neither greater than unity nor than the standard union bound. Accordingly an improved upper bound is expected by tuning ρ, and the bound must be convergent (since it is always less than unity and is monotonically increasing as more terms of the weight spectrum are added.) It remains to optimize the “tilting” pdf ψ (α ) and parameter ρ for the best upper bound. The principal consideration is the easy computation of the resulting bound, which leads to the following remarks:

7

z

Had the parameter ρ been permitted to change with α, i.e, ρ = ρ ( α ) , the bound would be tighter. Unfortunately the Jensen inequality appears to be inapplicable. For this reason ρ is chosen as a constant with respect to α in this paper.

z

The optimum tilting pdf can be derived by using the variational method [22, Appendix C], which however makes (14) regress to (12). It is somewhat tricky to select ψ (α ) in order to result in a closed-form yet tight bound. For Rayleigh fading channels, it is natural to consider a complex Gaussian tilting pdf.

A. Determinant Bound For the sake of convenience, it is assumed that the pdf of fading coefficients admits the form f α (α ) =

1

π

nR nT

R

e− tr( αR

nR

−1 H

α )

,

where R, positive definite, is the spatial correlation matrix on the transmitter side. Note that this is often a valid model of the down-link in mobile communications, where, because of the lack of rich scattering, much wider spacing between the antennas is needed at the base station to achieve independent fading than at the mobile [23]. To obtain a closed-form upper bound, one may choose a zero-mean complex Gaussian tilting pdf

ψ (α ) =

1

π

nR nT

F

nR

e − tr( αF

−1 H

α )

,

F

0.

(15)

Substituting this pdf into (14) and using the alternative form of the Q function [24] Q( x) =

1

π

π /2

∫e



x2 2sin 2 θ

dθ ,

for x > 0,

0

one obtains ρ

⎛ ⎡ γ C R −1 (1− ρ ) F −1 ⎤ H ⎞ ⎧ ⎫ − tr ⎜ α ⎢ s 2 + − ⎥ α ⎟⎟ ⎜ ⎢ 4sin θ ρ ρ / 2 π ⎥⎦ ⎝ ⎣ ⎠ ⎪⎪ ⎪⎪ 1 e α Pf ≤ ⎨ ∫ L ∑ AC ∫ dθ d ⎬ . n (1−1/ ρ ) n /ρ π 0 π nT nR F R R R C ⎪ ⎪ ⎪⎭ ⎩⎪

(16)

Exchanging the order of integration in (16) and applying the integral formula of the complex

8

Gaussian pdf [7, Eq. (B.10)] yield ⎧⎪ A Pf ≤ ⎨ L∑ C π ⎩⎪ C

π /2



R



nR

F

ρ

⎛1 ⎞ nR ⎜ −1⎟ ⎝ρ ⎠

0

γ sC

4sin 2 θ

+

R −1

ρ



(1 − ρ ) F −1

− nR

ρ

ρ

⎫⎪ dθ ⎬ , ⎭⎪

(17)

where the integral with respect to α converges in the region

γ sC

4sin 2 θ

+

R −1

ρ



(1 − ρ ) F −1

ρ

0

for any θ ∈ [0, π / 2] and any C in the weight spectrum. The region of convergence remains the same if sin 2 θ is removed, since sin 2 θ ≤ 1 and since C is positive definite. The bound (17) remains to be minimized for the tightest upper bound. This will be referred to as the determinant bound henceforth, for the calculation of matrix determinants is involved in (17). In summary, the determinant bound is given by

⎧⎪ A Pf ≤ min ⎨ L ∑ C ⎩⎪ C π

π /2



nR

−1

R F

ρ

0

γ s CF R −1 F (1 − ρ ) I + − ρ ρ 4 sin 2 θ

− nR

⎫⎪ dθ ⎬ ⎭⎪

ρ

(18)

subject to the following constraints 0 ≤ ρ ≤ 1, F

γ sC 4

+

R −1

ρ



0, (1 − ρ ) F −1

0, ∀ C ∈ {C , AC }.

ρ

(19)

The constrained nonlinear optimization constitutes the main computation cost of the new bound (note that the optimization under the constraint of F

0 is also needed in the ellipsoidal bound

(9).) There exist a number of efficient algorithms of constrained optimization in standard texts [25]. In this paper, a simpler way is used. The first and second constraints in (19) can be eliminated by writing ρ = exp(−ς 2 ) and by applying the Cholesky decomposition F = TT H [26], respectively. The third is handled by imposing a high penalty (e.g., 1010) if it is violated. Then ς and T are free parameters, which enables the use of unconstrained optimization procedures. A couple of remarks concerning the determinant bound follow. z

Since the calculation of determinants is easier than that of residues, the new bound has lower

9

computational complexity and better numerical stability than the ellipsoidal bound. z

The new bound assumes the full knowledge of the weight spectrum {C, AC}. It is incomplete to determine the bound with the knowledge of the eigenvalue spectrum.

B. Eigenvalue Bound If the tilting pdf ψ (α ) is characterized by a single parameter, the bound can be further simplified at some expense of tightness. Letting F = rR, one has ⎧⎪ A Pf ≤ ⎨ L∑ C ⎩⎪ C π

π /2

∫r

nR nT

ρ

0

rγ s CR rI (1 − ρ ) I + − ρ 4sin 2 θ ρ

− nR

ρ

⎫⎪ dθ ⎬ , ⎭⎪

(20)

under the constraint that the matrix within the determinant operator is positive definite. A salient feature of (20) is that it is determined by the eigenvalues of CR. If there is uncorrelated fading at the transmitter, i.e., F = rI, then (20) will be solely determined by the eigenvalues λi ( i = 1,

, nT ) of

C. Also, the positive definite condition can be simplified to

r>

1− ρ , 1 + λi ργ s / 4sin 2 θ

for 0 ≤ ρ ≤ 1 .

Again, sin 2 θ can be removed without affecting the constraint. Let λmin denote the globally minimum eigenvalue in the whole eigenvalue spectrum. Then this simple version of the new bound admits the form − nR π / 2 nR nT nT ⎧⎪ ⎫⎪ ⎡ rγ s λi ⎛ 1 − r ⎞⎤ Aλ ρ + − θ 1 Pf ≤ min ⎨ L∑ r d ⎬ ⎢ ⎥ ∏ ⎜ 2 π ∫0 ρ ⎟⎠ ⎦ i =1 ⎣ 4sin θ ⎝ ⎩⎪ λ ⎭⎪

ρ

(21)

where

r>

1− ρ 1 + λmin ργ s / 4

,

for 0 ≤ ρ ≤ 1 .

(22)

This is referred to as the eigenvalue bound, since it is determined by the eigenvalue spectrum. In the case of correlated fading, the bound is still valid if λi is understood as the eigenvalue of CR. Some remarks on the eigenvalue bound follow. z

Since the eigenvalue bound is a special case of the determinant bound, it is necessarily looser. 10

z

The eigenvalue bound can be evaluated faster than the determinant bound, as there are just two parameters ρ and r to be optimized.

z

The eigenvalue bound is a counterpart of the spherical bound (10) but easier to compute, as the spherical bound involves the calculation of residues.

C. Bound on BER To derive Gallager’s second bound on the BER of space-time codes, one also starts from the LBA bound ⎡1 ⎛ γ ⎞⎤ i Pb ≤ ∫ min ⎢ , ∑ AC ,i Q ⎜⎜ s tr ( αCα H ) ⎟⎟ ⎥ f α (α )dα , ⎢⎣ 2 C ,i k ⎝ 2 ⎠ ⎥⎦

(23)

where AC ,i denotes the average number of first-event errors with correlation matrix C and information weight i, and k is the number of information bits per trellis level. The bound (23) is an extension of [1, (11)] to space-time codes. Aided by the inequality [21, p. 314] min{a, b} ≤ a ρ b1− ρ , for 0 ≤ ρ ≤ 1, a, b ≥ 0 ,

(24)

one obtains 1− ρ

⎛1⎞ Pb ≤ ∫ ⎜ ⎟ ⎝2⎠

ρ

⎡ i ⎛ γ ⎞⎤ H ⎢ ∑ AC ,i Q ⎜⎜ s tr ( αCα ) ⎟⎟ ⎥ f α (α )dα . ⎢⎣ C ,i k ⎝ 2 ⎠ ⎥⎦

(25)

It is worth mentioning that the factor (1/ 2)1− ρ represents an improvement upon the original Gallager bound. Interestingly, the inequality (24) is also used to derive the Chernoff bound in information-theoretic literature [21, p. 315]; and Gallager’s bound essentially applies (24) in the special case b = 1. Proceeding as for the determinant bound on the frame error probability, one arrives at 1− ρ

⎛1⎞ Pb ≤ min ⎜ ⎟ ⎝2⎠

⎧⎪ iAC ,i 1 π / 2 −1 n ⎨∑ ∫ RF ⎪⎩ C ,i k π 0

R



γ s CF 4 sin θ 2

under the constraints (19). 11

−1

+

R F

ρ



(1 − ρ ) I

ρ

− nR

⎫⎪ dθ ⎬ ⎪⎭

ρ

(26)

The eigenvalue-spectrum version of the bound can be derived in a similar manner, which reads − nR 1− ρ π / 2 nR nT nT ⎫⎪ ⎡ rγ s λi ⎛ 1 − r ⎞⎤ 1⎞ ⎧ ⎪ iAλ,i 1 ⎛ ρ + − Pb ≤ min ⎜ ⎟ ⎨∑ r d θ 1 ⎬ ⎢ ⎥ ∏ ⎜ ⎟ 2 ρ ⎠⎦ ⎝ 2 ⎠ ⎪ λ k π ∫0 i =1 ⎣ 4sin θ ⎝ ⎪⎭ ⎩

ρ

(27)

under the constraints (22), where Aλ,i is similarly defined.

D. Discussions Aside from tackling quasi-static fading channels, the method of this paper differs from [17], [18], [19] in that no partition of codewords is made. Recall that in the context of turbo codes, the code is partitioned into a set of subcodes according to the distance spectrum, and Gallager’s second bound is applied to each subcode individually. To extend the idea to space-time codes, one may partition the code into a set of subcodes according to the weight spectrum {C, AC}. In Appendix A, however, it is shown that this approach leads to a divergent bound for space-time codes. A zero-mean Gaussian tilting pdf has been chosen in this section so that the closed-form expression (17) could be easily derived. A natural question to ask is whether the bound could be tightened by using a nonzero-mean Gaussian tilting pdf, for one would have more free parameters to optimize. Unfortunately, it is shown in Appendix B that this is impossible, which means that the bounds of this paper are best possible for Rayleigh fading and complex Gaussian tilting pdf’s. The ellipsoidal and determinant bounds have the common starting point⎯the LBA technique. Both can be considered means to overcome the multi-integral of the LBA technique at some cost of tightness. Since the LBA technique in essence uses the union bound in a wise way, all of these bounds only require the knowledge of the weight spectrum. All these bounds are convergent, which implies that the contribution of the error events with C or λ must decreases exponentially for exponentially increasing AC or Aλ. Therefore, the first few terms of the weight or eigenvalue spectrum usually dominate the performance. In practice, it is often enough to truncated the spectrum at a certain threshold. This will be confirmed by numerical results in Section V. In Table I, we compare the computational complexity of various bounds for space-time codes. 12

The transfer-function technique [11] is applicable to the LBA bound but not to the tight bounds (nor to the traditional union bound), for the matrix determinant or eigenvalue is not an additive measure. The LBA bound has 2nTnR real-valued variables (i.e., α) to integrate, whose complexity grows exponentially; at each step of the integration, the inversion of an NP-by-NP matrix, of complexity

O( N P3 ) , is required to evaluate the transfer function, where NP is number of states of the product trellis [11]. Hence, this method is only practical for codes with small trellises. On the other hand, we keep the first NS terms of the spectrum when computing other bounds. The Hermitian matrix F 0 in the ellipsoidal and determinant bounds has nT2 real free parameters. While the bounds derived in this paper have one more parameter ρ to optimize than [6], the major advantage is the avoidance of residue computations. We consider two numerical methods of residue computation: one is the perturbation method with complexity O(nT2 nR2 ) [27], the other is saddle-point integration [7] using, say K points, where it is observed that K has to be very large here. Unfortunately, both methods tend

to be unstable for poles of high multiplicities during the optimization process, no matter how large K is. Finally, the factors nT3 and nT for the new bounds come from the computations of determinant in (18) and multiplication in (21), respectively. TABLE I COMPARISON OF COMPUTATIONAL COMPLEXITY LBA-Transfer Ellipsoidal Spherical Determinant Function Bound Bound Bound Bound Number of Real 2nTnR nT2 nT2 +1 1 Parameters (to integrate) O(nT2 nR2 N S ) O(nT2 nR2 N S ) Computations O(nR nT2 N P3 ) O(nT3 N S ) per Step or O( KN S ) or O( KN S )

Eigenvalue Bound 2 O(nT N S )

IV. APPLICATION TO CONVOLUTIONAL CODING ON BLOCK FADING CHANNELS In this section, the new bound is applied to performance analysis of convolutional codes on block fading channels, a problem that motivated Malkamäki and Leib to develop the LBA bound [1]. The system model is illustrated in Fig. 2, where the output of a rate k/n (binary) encoder is interleaved over n subchannels (i.e., the n output bits at each stage of encoding are assigned to the subchannels 13

respectively.) For simplicity it is assumed that a single receive antenna is used in each subchannel. With slight modification, this problem can be put into the framework of performance bounds for space-time codes. Suppose that the convolutional code has the component-distance spectrum {d, i, Ad,i }, where Ad,i is the number of first error events with (Hamming) distance vector d = (d1, …, dn) and i information bit errors. The elements of d represent the distances on the subchannels. The bounds derived in Section III are valid verbatim for convolutional codes with the substitution: ⎡ d1 ⎢0 d n ]) 4 ⎢ ⎢ ⎢ ⎣0

C ← 4diag([ d1

0 d2

0

0⎤ 0 ⎥⎥ . ⎥ ⎥ dn ⎦

By inspection it is obvious that a diagonal matrix will be sufficient for F, namely, F = diag([f1, … fn]), fj > 0. The main contribution of this section is the derivation of an improved bound on the BER of systematic convolutional codes. As observed by Hu and Miller [28], there is room of improvement if the number ½ in the LBA technique is replaced by a better bound. In particular, this is easy to do for systematic codes. The method is to consider a trivial decoder that makes symbol-by-symbol hard decisions on the systematic bits, which are transmitted over the first subchannel without loss of generality. The BER of this trivial decoder, given by Pb (α 1 ) = Q

(

)

12 1 2γ s | α 1 |2 ≤ e −γ s |α | , 2

(28)

must be higher than that of the minimum-BER decoder. It is widely known that the maximum a posteriori probability (MAP) decoding yields the minimum BER [11]. Therefore, what derived in the following is valid for MAP decoding. While the ML decoding only minimizes the sequence error probability, its BER performance is very close to that of MAP decoding. Replacing ½ by (28), and employing inequality (24) again, one obtains

14

n ⎡1 ⎛ 12 i Pb ≤ ∫ min ⎢ e−γ s |α | , ∑ Ad ,i Q ⎜ 2γ s ∑ d j α j ⎜ j =1 d ,i k ⎢⎣ 2 ⎝ 1− ρ

⎛1 ⎞ ≤ ∫ ⎜ e−γ s |α | ⎟ ⎝2 ⎠ 12

n ⎡ i ⎛ j γ A Q 2 ⎢∑ ⎜⎜ s∑dj α d ,i j =1 ⎢⎣ d ,i k ⎝

2

⎞⎤ ⎟⎟ ⎥ fα (α )dα ⎠ ⎥⎦

(29)

ρ

2

⎞⎤ ⎟⎟ ⎥ fα (α )dα . ⎠ ⎥⎦

After some manipulations, one arrives at the determinant bound ⎧ iA 1 π / 2 n 1⎞ ⎪ ⎛ Pb ≤ min ⎜ ⎟ ⎨∑ d ,i f j1/ ρ ∏ ∫ ⎝ 2 ⎠ ⎪ d ,i k π 0 j =1 ⎩ 1− ρ

−1 ⎡γ sd j f j f j ⎛ 1 ⎞⎤ ⎪⎫ ⎢ 2 + + (γ sδ j − 1) ⎜ − 1⎟ ⎥ dθ ⎬ ρ ⎝ ρ ⎠⎦ ⎣ sin θ ⎪⎭

ρ

(30)

where δ j = 1 if j = 1 and = 0 otherwise. The constraints of the minimization are 0 ≤ ρ ≤ 1, f j > 0,

γ sd j f j +

fj

⎛1 ⎞ + (γ sδ j − 1) ⎜ − 1⎟ > 0, ∀ d j ∈ {d , i, Ad ,i }. ρ ⎝ρ ⎠

In this case, the determinant of matrix reduces to the product of diagonal elements. While regular interleaving is employed in Fig. 2, the analysis can be easily extended to random interleaving. Only the component-distance spectrum has to be recalculated (see [29]). However, full diversity is not always achieved under random interleaving [30]. This is because the probability of having codewords with component distance di = 0 for some i will be nonzero.

IV. NUMERICAL RESULTS In this section, numerical results on the performance bounds for a variety of codes are presented. As mentioned earlier, the knowledge of the weight spectrum is needed to calculate the bounds. The weight enumeration of space-time code generally consists of two steps. The product-state diagram is first reduced to a diagram with minimum product states [11], [14]; then the information about the weight spectrum can be extracted from the reduced diagram in a number of ways. The trellis-search method of [13] can be used instead, which is however only practical for the search of the first few terms of the spectrum. In this section, we use a method similar to those given by Viterbi et al. [31] and by McEliece [32] for convolutional codes. This method only requires the recursion of a set of difference equations, which is more manageable and computationally less burdensome. 15

First, the new bound is examined on uncorrelated fading channels. Fig. 3 shows the results of the (7, 5) 4-state BPSK code [13], where the determinant bound is compared with the ellipsoidal bound and simulation results. To examine the convergence behavior, the weight spectrum is truncated at a threshold H, i.e., a matrix C will be excluded if any entry of C has absolute value greater than H. Other choices of the threshold are possible (e.g., the determinant in [13]); H is natural for the method used in this paper. The determinant bound is then computed for the truncated weight spectrum with different values of H (the standard union bound is not depicted since it is too loose.) The new bound converges fast, and no growth is observed if H goes beyond 40. It is seen that the new bound is 2 dB looser than the ellipsoidal bound, and is about 3 dB looser than simulation points. To further study the convergence behavior, the determinant and traditional union bounds are compared in Fig. 4 for the 4 and 16-state QPSK codes of Tarokh, Seshadri, and Calderbank (TSC) [9]. The traditional union bound obviously diverges, and gives conflicting results. In contrast, the determinant bound exhibits rapid convergence, correctly predicting the superiority of the 16-state code. Indeed, rapid convergence is a distinct feature of the new bounds. As a result, the number of terms in the weight (or eigenvalue) spectrum to be included does not have to be very large. In practice, a safe way is to try different values of H, and truncate at a threshold beyond which no (or little) further growth is observed. Fig. 5 compares the determinant bound with the eigenvalue bound for the TSC 16-state QPSK code, where the ellipsoidal bound and the spherical bound are also shown. It is seen that the eigenvalue bound is roughly 2 dB looser than the determinant bound, which is in turn looser than the ellipsoidal bound by another 2 dB. The reduction from F to a single parameter r appears to incur more expense for Gallager’s second bounding technique here, for the spherical bound is only 0.5 dB looser than the ellipsoidal bound. Next, the effect of correlated fading on the tightness of the bound is illustrated in Fig. 6 for the TSC 4-state QPSK code with two transmit antennas. The correlation matrix is given by

16

⎡1 R=⎢ ⎣μ

μ⎤ 1 ⎥⎦

where μ denotes the correlation coefficient. The LBA bound for this code was given in [3, Fig.5] for correlated fading. In Fig. 6, the determinant bound is plotted for different values of μ = 0, 0.65, and 0.85. It can be seen that the bound keeps the gap of about 2 dB from the respective simulation points, regardless of the correlation coefficient. Fig. 7 demonstrates the new bound on BER for the (1, 3) 2-state BPSK code, where the LBA bound is included as a benchmark. It is seen that the determinant bound is about 4 dB looser than simulation points, and it retains the gap to the LBA bound within 3 dB. It is a general trend that the bounds on BER are less tight than that those on FER [1]. This is because ½ is a poor estimate of the BER, whereas 1 is a relatively good estimate of the FER in deep fading because of the multiplication by L. To show the improved determinant bound for convolutional codes on block fading channels, a recursive systematic convolutional (RSC) code with generator polynomial (7, 5) is used in Fig. 8. MAP decoding is implemented by the BCJR algorithm [11]. Only the BER of MAP decoding is depicted, as very little difference exists between the performance of MAP and ML decoders. It is seen that the bound (30) results in nearly 2 dB improvement, and it is about 4.5 dB from simulated BER. The simulated performance of the RSC code is slightly better than that of the nonsystematic nonrecursive (7, 5) code (cf. [1, Fig. 2]).

V. CONCLUSIONS In this paper, the generalized union bound for space-time codes has been developed. The new bound only requires the calculation of matrix determinants or eigenvalues (followed by optimization), thereby being easier to compute and numerically more stable than the existing bound based on Gallager’s first bounding technique. While this comes at the cost of some tightness, the new bound exhibits fast convergence and is able to predict the relative performance of codes. Therefore, it is a useful tool in performance analysis and code search. The method is applicable to 17

other codes on fading channels, which is demonstrated by the analysis of convolutional codes in block fading. The improved bound on the BER of systematic convolutional codes may prompt a similar effort towards systematic space-time codes. However, the QR decomposition to recover the systematic bit requires the condition nR ≥ nT, which appears to be restrictive. For this reason, this has not been pursued in this paper. The nonlinear optimization dictates the computing time of the new bound. A simple unconstrained optimization procedure was employed in this paper. Better optimization algorithms will speed up the computation of the bound and give more robust results. On the other hand, it is interesting to see if there exists a better “tilting” pdf ψ (α ) than the complex Gaussian pdf.

APPENDIX A. Partition Let CC denote the subcode consisting of the reference codeword and other codewords with codeword-difference correlation matrix C. The frame error rate for the subcode is bounded as (by applying the Chernoff bound Q(x) ≤ e− x

2

/2

/ 2) ρ

Pf ,CC ≤ ∫ ⎡⎣ LPE ,CC (α ) ⎤⎦ fα (α )dα α ρ

⎛ γ C ⎞ ⎡1 − tr ⎜α s α H ⎟ ⎤ −1 H 1 ⎝ 4 ⎠ ≤ ∫ ⎢ LAC e e − tr(α R α ) dα ⎥ nn nR α 2 ⎢⎣ ⎥⎦ π R T R

ρ

⎛1 ⎞ ργ s CR = ⎜ LAC ⎟ +I 4 ⎝2 ⎠

− nR

.

Thus, the upper bound for the overall code is given by − nR ρ ⎧⎪⎛ 1 ⎫⎪ ⎞ ργ s CR + I ⎬, Pf ≤ ∑ min ⎨⎜ LAC ⎟ 0≤ ρ ≤1 4 ⎠ C ⎩⎪⎝ 2 ⎭⎪

(31)

where ρ is optimized for each subcode individually. This approach works if a small number of terms of the weight spectrum are accounted for. However, the bound will be divergent as more terms are added. This is because

( LAC )

ρ

remains exponential in C no matter how ρ changes, if AC increases

18

exponentially. B. Nonzero-Mean Tilting pdf It is shown here that the determinant bound cannot be further improved even if a nonzero-mean Gaussian tilting pdf is used in Jensen’s inequality. Let the tilting pdf with nonzero mean α be given by

ψ (α ) =

1

π

nT nR

F

nR

e

(

),

− tr ( α − α ) F −1 ( α − α ) H

0.

F

(32)

Substituting this pdf into (14), one has ⎧ ⎪⎪ A Pf ≤ ⎨ ∫ L∑ C C π ⎪ ⎩⎪

π /2

∫ dθ

e

⎛ ⎛ γ C R −1 ⎞ H (1− ρ ) F −1 ( α − α )H − tr ⎜ α ⎜ s 2 + ⎟α −( α −α ) ⎜ ⎜ 4sin θ ρ ⎟ ρ ⎠ ⎝ ⎝

πn n F T R

0

nR (1−1/ ρ )

R

⎞ ⎟ ⎟ ⎠

nR / ρ

ρ

⎫ ⎪⎪ dα ⎬ . ⎪ ⎭⎪

(33)

It can be derived after some manipulations that the integral is formally the same as (16), with an extra exponential term

e

−1 ⎛ ⎡ ⎤ R −1 (1− ρ ) F −1 ⎞ (1− ρ ) F −1 ⎥ H (1− ρ ) F −1 (1− ρ ) F −1 ⎛ γ s C − tr ⎜ α ⎢ − − + − α ⎜⎜ ⎟⎟ 2 ⎜⎜ ⎢ ⎥ ρ ρ ρ ρ ⎝ 4sin θ ρ ⎠ ⎦ ⎝ ⎣

⎞ ⎟ ⎟⎟ ⎠

=e

−1 ⎛ ⎡ −1 ⎜ ⎢⎛ γ s C R −1 ⎞ ρ F ⎤⎥ H − tr ⎜ α ⎜ + α ⎟⎟ − 2 ⎜ ⎜ ⎢⎣⎝ 4sin θ ρ ⎠ 1− ρ ⎥⎦ ⎝

⎞ ⎟ ⎟ ⎟ ⎠

where the matrix inversion lemma [26] is used. With these the upper bound is given by ⎧ ⎪ ⎪ A Pf ≤ ⎨ L∑ C ⎪ C π ⎪ ⎩

π /2

∫ 0

dθ R −1F

nR / ρ

e

−1 ⎛ ⎡ −1 ⎜ ⎢⎛ γ s C R −1 ⎞ ρ F ⎤⎥ H α − tr ⎜ α ⎜⎜ + 2 ⎟⎟ − ⎜ ⎢⎣⎝ 4sin θ ρ ⎠ 1− ρ ⎥⎦ ⎝

⎞ ⎟ ⎟ ⎟ ⎠

γ s CF R −1F (1 − ρ ) I + − ρ ρ 4sin 2 θ

nR

⎫ ⎪ ⎪ ⎬ ⎪ ⎪ ⎭

ρ

(34)

under the same constraints as (19). The exponential term in (34) represents an extra degree of freedom in optimization. Unfortunately, the matrix definite, because

19

⎛ γ C R−1 ⎞⎟ ⎜ s + ⎜ ⎟ 2 ⎝ 4sin θ ρ ⎠

−1



ρF 1− ρ

will always be negative

γ sC

4sin 2 θ

+

R −1

ρ



(1 − ρ ) F −1

ρ

−1 ⎛ γ sC ⎛ γ sC R −1 ⎞ (1 − ρ ) F −1 ⎤ R −1 ⎞ ⎡ + =⎜ + ⎥ ⎟ ⎟ ⎢I − ⎜ 2 2 ρ ⎠ ρ ⎥⎦ ⎝ 4 sin θ ρ ⎠ ⎢⎣ ⎝ 4sin θ −1 ⎛ γ sC R −1 ⎞ ⎡⎛ γ s C R −1 ⎞ ρ F ⎤ ⎛ (1 − ρ ) F −1 ⎞ + + − =⎜ ⎢ ⎥⎜− ⎟ ⎜ ⎟ ⎟. 2 ρ ⎠ ⎢⎣⎝ 4sin 2 θ ρ ⎠ 1 − ρ ⎥⎦ ⎝ ρ ⎝ 4sin θ ⎠

Note that − (1 − ρ ) F / ρ is negative definite for 0 < ρ < 1 . Since −1

positive definite and since

γ sC 4 sin 2 θ

+

R −1

ρ



(1 − ρ ) F −1

ρ

γ sC 4 sin 2 θ

+

R −1

ρ

is apparently

is constrained to be positive definite,

−1

−1 R ⎞ ρF ⎛ γ sC ⎜ 4 sin 2 θ + ρ ⎟ − 1 − ρ must be negative definite [26]. ⎝ ⎠

Hence, one has

e

−1 ⎛ ⎡ −1 ⎜ ⎛ γ C R −1 ⎞ ρ F ⎤⎥ H − tr ⎜ α ⎢⎜ s 2 + − α ⎟ ⎜ ⎟ ⎜ ⎣⎢⎝ 4sin θ ρ ⎠ 1− ρ ⎦⎥ ⎝

⎞ ⎟ ⎟ ⎟ ⎠

≥1

for any α . It is therefore concluded that a nonzero-mean tilting Gaussian pdf does not lead to any improvement of the bound.

ACKNOWLEDGMENT The author would like to thank Xiaofu Wu and T. M. Duman for helpful discussions.

REFERENCES [1] E. Malkamäki and H. Leib, “Evaluating the performance of convolutional codes over block fading channels,” IEEE Trans. Inform. Theory, vol. 45, pp. 1643–1646, July 1999. [2] G. Caire and G. Colavolpe, “On low-complexity space-time coding for quasi-static fading channels,” IEEE Trans. Inform. Theory, vol. 49, pp. 1400–1416, June 2003. [3] A. Stefanov and T. M. Duman, “Performance bounds for space-time trellis codes,” IEEE Trans. Inform. Theory, vol. 49, pp. 2134–2140, Sept. 2003. [4] A. P. des Rosiers and P. H. Siegel, “On performance bounds for space-time codes on fading channels,” IEEE Trans. Commun., vol. 52, pp. 1688-1697, Oct. 2004. [5] H. Bouzekri and S. L. Miller, “Distance spectra and performance bounds of space-time trellis codes over quasi-static fading channels,” IEEE Trans. Inform. Theory, vol. 50, pp. 1820–1831, Aug. 2004. [6] C. Ling, K. H. Li and A. C. Kot, “Gallager bounds for space-time codes,” IEEE Communication Theory Workshop

20

2004, Capri, Italy, May 2004. Available online at http://www.ntu.edu.sg/home5/pg01854370/publications. [7] C. W. Helstrom, Elements of Signal Detection and Estimation. Englewood Cliffs, NJ: Prentice-Hall, 1995. [8] R. G. Gallager, Information Theory and Reliable Communication. New York, NY: John Wiley & Sons, 1968. [9] V. Tarokh, N. Seshadri, and A. R. Calderbank, “Space-time codes for high data rate wireless communication: Performance criterion and code construction,” IEEE Trans. Inform. Theory, vol. 44, pp. 744–765, Mar. 1998. [10] S. S. Pietrobon, “On the probability of error of convolutional codes,” IEEE Trans. Inform. Theory, vol. 42, pp. 1562–1568, Sept. 1996. [11] C. Schlegel, Trellis Coding. New York: IEEE Press, 1997. [12] G. Caire and E. Viberbo, “Upper bound on the frame error probability of terminated trellis codes”, IEEE Commun. Lett., vol. 2, pp. 2–4, Jan. 1998. [13] D. Aktas and M. P. Fitz, “Distance spectrum analysis of space-time trellis-coded modulations in quasi-static Rayleigh fading channels,” IEEE Trans. Inform. Theory, vol. 49, pp. 1335-1344, Dec. 2003. [14] J. Shi and R. Wesel, “Efficient computation of trellis code generating functions,” IEEE Trans. Commun., vol. 52, pp. 219-227, Feb. 2004. [15] R. G. Gallager, Low-Density Parity-Check Codes. Cambridge, MA: MIT Press, 1963 [16] R. M. Fano, Transmission of Information. Cambridge, MA: MIT Press, 1961. [17] T. M. Duman and M. Salehi, “New performance bounds for turbo codes,” IEEE Trans. Commun., vol. 46, pp. 717-723, June 1998. [18] D. Divsalar and E. Biglieri, “Upper bounds on error probabilities of coded systems beyond the cutoff rate,” IEEE Trans. Commun., vol. 51, pp. 2011-2018, Dec. 2003. [19] S. Shamai (Shitz) and I. Sason, “Variations on the Gallager bounds, connections, and applications,” IEEE Trans. Inform. Theory, vol. 48, pp. 3029-3051, Dec. 2002. [20] I. Sason, S. Shamai (Shitz) and D. Divsalar, “Tight exponential upper bounds on the ML decoding error probability of block codes over fully interleaved fading channels,” IEEE Trans. Commun., vol. 51, pp. 1296-1305, Aug. 2003. [21] T. M. Cover and J. A. Thomas, Elements of Information Theory. New York: John Wiley & Sons, 1991. [22] S. Benedetto and E. Biglieri, Principle of Digital Communication. New York: Kluwer Academic/Plenum Publishers, 1999. [23] W. C. Y. Lee, Mobile Communications Engineering, 2nd Ed. McGraw-Hill, 1998. [24] M. K. Simon and M.-S. Alouini, Digital Communication over Fading Channels. New York: John Wiley & Sons, 2000. [25] B. S. Gottfried and J. Weisman, Introduction to Optimization Theory. Englewood Cliffs, NJ: Prentice-Hall, 1973. [26] R. A. Horn and C. R. Johnson, Matrix analysis. Cambridge, UK: Cambridge University Press, 1993. [27] S. Siwamogsatham, M. P. Fitz, and J. Grimm, “A new view of performance analysis of transmit diversity schemes in correlated Rayleigh fading,” IEEE Trans. Inform. Theory, vol. 48, pp. 950–956, Apr. 2002. [28] J. Hu and S. Miller, “An improved upper bound on the performance of convolutional codes over quasi-static fading channels,” in Proc. Globecom, San Francisco, CA, 2003, pp. 1593–1597. [29] X. Wu, H. Xiang, C. Ling et al., “Performance analysis of turbo-like codes over noncoherent block AWGN and fading channels,” IEEE Trans. Wireless Commun., accepted for publication. [30] J. J. Boutros, E. C. Strinati, and A. G. Fabregas, “Turbo code design for block-fading channels,” In Proc. 42nd Ann. Allerton Conf. Commun., Control and Computing, Allterton, IL, Sept.-Oct. 2004. [31] R. J. McEliece, “How to compute weight enumerators for convolutional codes,” in Communications and Coding, M. Darnel and B. Honary, Eds. New York: Wiley, 1998, ch. 6, pp. 121-141. [32] A. J. Viterbi, A. M. Viterbi, J. Nicolas and N. T. Sindushayana, “Perspectives on interleaved concatenated codes with iterative soft output decoding,” in Proc. Int. Symp. Turbo Codes, Brest, France, Sept. 1997, pp. 47-54.

21

ηt1

α t1,1 Data Trellis Encoder

α t1,n

R

α tn

T

ML Decoder

,1

α tn

T

, nR

ηtn

R

Fig. 1. Model of space-time coding with nT transmit and nR receive antennas.

ηt1 α t1 Data

Convolutional Encoder

ML/MAP Decoder

α tn ηtn Fig. 2. Model of convolutional coding over a block fading channel with n subchannels.

22

0

10

-1

Frame Error Rate

10

-2

10

Determinant Bound, H = 40 Determinant Bound, H = 20 Determinant Bound, H = 12 Ellipsoidal Bound, H = 40 Simulation -3

10

4

6

8 10 12 14 16 Symbol SNR per Receive Antenna (dB)

18

20

Fig. 3. Performance bounds for the (7, 5) 4-state BPSK code for nT = 2, nR = 1 and L = 130.

1

10

Union Bound, 16-State Union Bound, 4-State Determinant Bound, 16-State Determinant Bound, 4-State 0

Frame Error Rate

10

-1

10

-2

10

2

4

6

8 10 Threshold H

12

14

16

Fig. 4. Convergence behavior for the TSC 4 and 16-state QPSK codes for nT γ s = 20 dB, nT = 2, nR = 1 and L = 130.

23

0

10

-1

Frame Error Rate

10

-2

10

Eigenvalue Bound Determinant Bound Spherical Bound Ellipsoidal Bound Simulation -3

10

4

6

8

10 12 14 16 18 20 Symbol SNR per Receive Antenna (dB)

22

24

Fig. 5. Performance bounds for the TSC 16-state QPSK code for nT = 2, nR = 1 and L = 130. 0

10

μ = 0.85, 0.65, 0

-1

Frame Error Rate

10

-2

10

Determinant Bound Simulation -3

10

10

12

14 16 18 20 22 24 Symbol SNR per Receive Antenna (dB)

26

28

Fig. 6. Performance bounds for the TSC 4-state QPSK code for nT = 2, nR = 1 and L = 130 on correlated fading channels.

24

0

10

Determinant Bound LBA Bound Simulation -1

Bit Error Rate

10

-2

10

-3

10

-4

10

0

2

4

6 8 10 12 14 16 Symbol SNR per Receive Antenna (dB)

18

20

Fig. 7. Performance bounds on BER of the (1, 3) 2-state BPSK code for nT = 2 and nR = 1.

0

10

Determinant Bound Improved Determinant Bound LBA Bound Simulation (MAP Decoding) -1

Bit Error Rate

10

-2

10

-3

10

0

2

4

6

8

10 12 Eb/N0 (dB)

14

16

18

20

Fig. 8. Performance bounds on BER of the (7, 5) RSC code on block fading channels with n = 2. 25

Generalized Union Bound for Space-Time Codes - Semantic Scholar

Department of Electronic Engineering, King's College London (email: cling@ieee.org) ... Index Terms⎯Gallager bounds, optimization, space-time coding, union ...

305KB Sizes 0 Downloads 450 Views

Recommend Documents

Generalized Union Bound for Space-Time Codes - Semantic Scholar
Cambridge, MA: MIT Press, 1963. [16] R. M. Fano ... Cambridge, UK: Cambridge University Press, 1993. ... Wireless Commun., accepted for publication. [30] J. J. ...

Approximate Test Risk Bound Minimization ... - Semantic Scholar
GALE program of the Defense Advanced Research Projects Agency, Contract. No. ...... recognition,” Data Mining and Knowledge Discovery, vol. 2, no. 2, pp.

Multimodal MRI reveals secondarily generalized ... - Semantic Scholar
The protocol included a 3D T1-weighted fast field-echo sequence [TR 11 ms, ... and less prone to increased neurotransmitter traffic than the signal reception.

Complexity of stochastic branch and bound for ... - Semantic Scholar
such methods is that in most problems of interest, the optimal solution involves ..... an analytical bound based on sampling a small set of beliefs and [12], which ...

A Generalized Mechanism for Perception of Pitch ... - Semantic Scholar
Jan 14, 2009 - (Oz, O1, O2, Iz). Frontal electrodes were se- lected based on regions of interest defined for the EAN in previous studies (Koelsch et al.,. 2007).

A Generalized Mechanism for Perception of Pitch ... - Semantic Scholar
Jan 14, 2009 - Subjects had no prior exposure to the musical system used in the present study. All research was approved by the Committee for the Pro- tection of Human Subjects at UC Berkeley. Procedure. Participants were seated in a sound-attenuated

Complexity of stochastic branch and bound for ... - Semantic Scholar
regret bounds [4] in a distribution-free framework. The aim of ..... The straightforward proof can be found in App. A. The only controllable variable is the branch-.

Linear Network Codes: A Unified Framework for ... - Semantic Scholar
This work was supported in part by NSF grant CCR-0220039, a grant from the Lee Center for. Advanced Networking, Hewlett-Packard 008542-008, and University of ..... While we call the resulting code a joint source-channel code for historical ...

Binary Codes Embedding for Fast Image Tagging ... - Semantic Scholar
tagging is that the existing/training labels associated with image exam- ..... codes for tags 'car' and 'automobile' be as close as possible since these two tags.

Linear Network Codes: A Unified Framework for ... - Semantic Scholar
Page 1 ..... For any n × ⌊nR⌋ matrix bn, we can build a linear channel code with .... For any n × n matrix cn, we can build a joint source-channel code for the.

Generalized Silver Codes
gineering, Indian Institute of Science, Bangalore 560012, India (e-mail: [email protected]; [email protected]). Communicated by E. Viterbo, Associate ...... 2006, he was with Robert Bosch India limited, Bangalore. Currently, he is a wor

Nonlinear adventures at the zero lower bound - Semantic Scholar
Jun 11, 2015 - consumption, inflation, and the one auxiliary variable. The Smolyak .... t has a recursive structure in two auxiliary variables x1;t and x2;t that satisfy εx1;t ¼ рεА1Юx2;t and have laws of ...... We start at the unconditional me

A Bound on the Label Complexity of Agnostic ... - Semantic Scholar
to a large pool of unlabeled examples, and is allowed to request the label of any .... Examples. The canonical example of the potential improvements in label complexity of active over passive learning is the thresholds concept space. Specifically ...

A Generalized Data Detection Scheme Using ... - Semantic Scholar
Oct 18, 2009 - We evaluated the performance of the proposed method by retrieving a real data series from a perpendicular magnetic recording channel, and obtained a bit-error rate of approximately 10 3. For projective geometry–low-density parity-che

Generalized and Doubly Generalized LDPC Codes ...
The developed analytical tool is then exploited to design capacity ... error floor than capacity approaching LDPC and GLDPC codes, at the cost of increased.

Generalized Additive Codes, Self-Dual Codes and ...
x = (x0,x1,··· ,xn−1) ∈ Zn. 4 is an arbitrary vector, and suppose ϕ(xj)=(xj,0,xj,1) for all j = 0,1,··· ,n − 1, then. ϕ(x0,x1,··· ,xn−1)=(x0,0,x1,0,··· ,xn−1,0,x0,1,x1,1,··· ,xn−1,1). (2.3). In this paper, we shall consid

Improved quantum hypergraph-product LDPC codes - Semantic Scholar
Leonid Pryadko (University of California, Riverside). Improved quantum ... Example: Suppose we take LDPC code [n,k,d] with full rank matrix ; then parameters of ...

Improved quantum hypergraph-product LDPC codes - Semantic Scholar
Leonid Pryadko (University of California, Riverside). Improved quantum ... Example: Suppose we take LDPC code [n,k,d] with full rank matrix ; then parameters of ...

On Generalized Weights for Codes over Zk
Jun 22, 2011 - introduced the GHWR of a linear code C over a finite chain ring and studied some properties of the GHWR. For any g, 1 ≤ g ≤ rank(C), we define the higher weight spectrum as. A g i = |{D : D is a Zk-submodule of C with rank(D) = g a

Anesthesia for ECT - Semantic Scholar
Nov 8, 2001 - Successful electroconvulsive therapy (ECT) requires close collaboration between the psychiatrist and the anaes- thetist. During the past decades, anaesthetic techniques have evolved to improve the comfort and safety of administration of

Considerations for Airway Management for ... - Semantic Scholar
Characteristics. 1. Cervical and upper thoracic fusion, typically of three or more levels. 2 ..... The clinical practice of airway management in patients with cervical.

Czech-Sign Speech Corpus for Semantic based ... - Semantic Scholar
Marsahll, I., Safar, E., “Sign Language Generation using HPSG”, In Proceedings of the 9th International Conference on Theoretical and Methodological Issues in.

Discriminative Models for Semi-Supervised ... - Semantic Scholar
and structured learning tasks in NLP that are traditionally ... supervised learners for other NLP tasks. ... text classification using support vector machines. In.