• 沒有找到結果。

DMT Optimal Codes Constructions for Multiple-Access MIMO Channel

N/A
N/A
Protected

Academic year: 2021

Share "DMT Optimal Codes Constructions for Multiple-Access MIMO Channel"

Copied!
24
0
0

加載中.... (立即查看全文)

全文

(1)

DMT Optimal Codes Constructions for

Multiple-Access MIMO Channel

Hsiao-feng (Francis) Lu, Member, IEEE, Camilla Johanna Hollanti, Member, IEEE,

Roope Iikanpoika Vehkalahti, Member, IEEE, and Jyrki Lahtonen, Member, IEEE

Abstract—Explicit code constructions for multiple-input mul-tiple-output (MIMO) multiple-access channels (MAC) with users are presented in this paper. The first construction is dedi-cated to the case of symmetric MIMO-MAC where all the users have the same number of transmit antennas and transmit at the same level of per-user multiplexing gain . Furthermore, we assume that the users transmit in an independent fashion and do not cooperate. The construction is systematic for any values of , and . It is proved that this newly proposed construction achieves the optimal MIMO-MAC diversity-multiplexing gain tradeoff (DMT) provided by Tse et al. at high-SNR regime.

In the second part of the paper we take a further step to inves-tigate the MAC-DMT of a general MIMO-MAC where the users are allowed to have different numbers of transmit antennas and can transmit at different levels of multiplexing gain. The exact op-timal MAC-DMT of such channel is explicitly characterized in this paper. Interestingly, in the general MAC-DMT, some users might not be able to achieve their single-user DMT performance as in the symmetric case, even when the multiplexing gains of the other users are close to 0. Detailed explanations of such unexpected result are provided in this paper. Finally, by generalizing the code construc-tion for the symmetric MIMO-MAC, explicit code construcconstruc-tions are provided for the general MIMO-MAC and are proved to be optimal in terms of the general MAC-DMT.

Index Terms—Cyclic division algebras (CDAs), diversity-mul-tiplexing gain tradeoff (DMT), multiple access channel (MAC), multiple-input multiple-output (MIMO) channel, space-time block codes (STBCs).

I. INTRODUCTION

D

URING the last decade extensive research has been car-ried out in the design of point-to-point space-time (ST) codes [1], [2] for multiple-input multiple-output (MIMO) com-munication systems. ST codes based on cyclic division algebras

Manuscript received April 24, 2009; revised June 09, 2010; accepted December 01, 2010. Date of current version May 25, 2011. This work was supported in part by the Taiwan National Science Council under Grant NSC 98-2221-E-009-045-MY3, through a grant for H.-F. Lu; in part by the Finnish Academy of Science and Letters and the Academy of Finland (Grant 210280) through grants for C. J. Hollanti; and in part by the Emil Aaltonen Foundation, Finland, through a grant for R. I. Vehkalahti.

The material in this paper was presented in part at the IEEE International Symposium on Information theory, Seoul, Korea, June 2009.

H.-F. Lu is with the Department of Electronical Engineering, National Chiao Tung University, Hsinchu 30010, Taiwan, R.O.C. (e-mail: francis@mail.nctu. edu.tw).

C. J. Hollanti is with the Department of Mathematics, FI-20014 University of Turku, Turku, Finland (e-mail: cajoho@utu.fi).

R. I. Vehkalahti is with the Department of Mathematics, FI-20014 University of Turku, Turku, Finland (e-mail: roiive@utu.fi).

J. Lahtonen is with the Department of Mathematics, FI-20014 University of Turku, Turku, Finland (e-mail: lahtonen@utu.fi).

Communicated by B. S. Rajan, Associate Editor for Coding Theory. Digital Object Identifier 10.1109/TIT.2011.2136170

(CDAs) [3]–[7] that can also be regarded as a kind of algebraic lattice codes and/or as a kind of linear dispersion ST codes [8] have been shown to perform extremely well. The error perfor-mance of these codes have been shown to be very close to the outage bound not only for practical numbers of antennas but also at moderate SNR values.

For high-SNR regime, the same point-to-point CDA-based ST codes have been shown [4] to be optimal in terms of the diversity-multiplexing tradeoff (DMT) proposed by Zheng and Tse [9]. Specifically, let and be respectively the numbers of transmit and receive antennas at transmitter and receiver ends.

Let , , denote the multiplexing gain such

that the actual transmission rate equals

SNR (1)

Assuming a MIMO Rayleigh block fading channel, it was shown [4] that at multiplexing gain , the CDA-based ST codes achieve the optimal codeword error probability

SNR (2)

at high-SNR regime, where by we mean the exponential equality defined in [9]. That is, we write SNR SNR if

SNR SNR

The notations of and are similarly defined. The exponent is commonly known as the DMT [9] and is given by a piecewise linear function connecting the points

for . Furthermore,

rep-resents the largest diversity gain that can be achieved by any point-to-point ST codes under Rayleigh block fading channel whenever the channel remains static for at least a block of channel uses [4] and varies independently from one block to another.

For other types of fading statistics, the CDA-based ST codes are also known [4] to be capable of achieving the optimal error performance in such channels that include Rician, Weibull and Nakagami as special cases. ST codes that are optimal in all fading statistics are coined approximately universal codes [4], [10].

If coding across independent fading blocks is allowed, the multiblock CDA code [6] has been shown to be approximately universal as well. In particular, it achieves codeword error probability

SNR (3)

(2)

at multiplexing gain , where is the number of independent fading blocks occupied by the code. The exponent

is known as the multi-block DMT [6], [9] when coding is applied over independent fading blocks. Therefore, the multiblock CDA-based ST code is optimal in terms of the multiblock DMT at high-SNR regime. More important, (3) indicates that the code has error probability decreasing to zero as approaches

in-finity whenever . Hence, the multiblock ST code

could potentially achieve the MIMO ergodic channel capacity at high-SNR regime and simultaneously be optimal in terms of the multiblock DMT at every discrete value .

Motivated by the promising outcome in the point-to-point scenario, the aim of this paper is to investigate the code con-struction for the multiple-access channel (MAC) scenario. We will concentrate on the uplink transmission from multiple mo-bile users to a common base station (or access point). Both the mobile users and the base station may be equipped with mul-tiple antennas.

Consider a MIMO-MAC with mobile users. For simplicity, we first focus on the case of symmetric MIMO-MAC [11], where each user is equipped with transmit antennas and communi-cates independently to the base station that has receive an-tennas. Furthermore, we assume that all the users transmit at the same level of multiplexing gain. With a slight abuse of notation, hereafter we will denote by the per-user multiplexing gain in

the symmetric MIMO-MAC. Let , be respectively

the ST codes used by the th user, . Each

code , , consists of matrices and

satisfies the following power constraint:

SNR (4)

where by we mean the Frobenius norm of matrix .

Furthermore, we require SNR for all such that

every user transmits at the same multiplexing gain . Let be

the channel matrix of the th user. We assume is

fixed for a block of channel uses. is known completely to the receiver at base station but unknown to all the users. Entries of are modeled as i.i.d. complex Gaussian random variables to model the MIMO Rayleigh block fading channel. Let be the signal matrix transmitted by the th user; then the signal matrix received at base station is given by

(5)

where is the noise matrix with i.i.d.

en-tries. When each user’s information is encoded independently, Tse et al. [11] proved that the tradeoff between the diversity gain and multiplexing gain in a symmetric MIMO-MAC is gov-erned by the following theorem.

Theorem 1 (Symmetric MAC-DMT [11]): In a symmetric

MIMO-MAC with users, each having transmit antennas

Fig. 1. The MAC DMT forK = 3 users with n = 2 and n = 2.

and transmitting independently at multiplexing gain , the max-imal possible diversity gain is given by

(6)

where is the point-to-point DMT for transmit

antennas, receive antennas and multiplexing gain defined as before (or see [9], [11]). Equation (6) is termed optimal sym-metric MAC-DMT. The multiplexing gain for nonnegative di-versity gain is bounded between

(7)

Compared with the point-to-point scenario, the decrease of maximal multiplexing gain by a factor of (see in of (7)) is due to the sharing of receive antennas among

users and the fact that . Equation (6)

also shows that when the level of multiplexing gain is low

such that , each user is able to retain

his single-user performance, i.e., , as

if there were no other users in the channel. On the other hand, when the level of multiplexing gain is high and , the MIMO-MAC system would operate in the antenna pooling region [11], and single-user performance can no longer be maintained. As a consequence, a much lower diversity gain

dominates the error performance in this region.

In Fig. 1 we demonstrate the above facts of the symmetric

MAC-DMT for the case of users, and . It

(3)

and antenna pooling regions is at and

the cut-off point of is at .

The construction of MAC-DMT optimal codes calls for a coding scheme that independently encodes, but simultaneously transmits, each mobile user’s information over the MIMO channel such that at receiver end, the decoding of all users’ signals achieves the best possible error performance dictated by the MAC-DMT. Thus, a coding scheme is called MAC-DMT

optimal if it achieves the following error performance under

joint decoding

SNR

A. Prior Work

Several works have been reported in this area. Nam et al. [12] presented the first MAC-DMT optimal scheme using a class of structured multiple-access random lattice ST codes. For the constructions of deterministic codes, below we briefly review some relevant earlier papers. Almost all deal exclusively with the two-user symmetric MIMO-MAC case, i.e., .

1) [13] extended the pairwise-error-probability-based design criteria of point-to-point ST codes to the MAC case for

users and , . An explicit (4 4)

two-user MIMO code,1i.e., a (2 4) code for each user, based

on independent Alamouti blocks [2] is also introduced in [13]. Yet, we remark that such code does not achieve the optimal symmetric MAC-DMT (6).

2) In [14] Badr and Belfiore proposed an explicit algebraic

code for and . The idea can be extended

to bigger values of . The determinant of the code matrix is nonzero thanks to a ”twisting element.” However, the determinant is vanishing. The decay of determinants of this two-user MIMO-MAC code was carefully studied in [15]. It was shown that the code is MAC-DMT optimal, when . Whether this code achieves the optimal MAC-DMT also when remains an open question. In [15] it was shown, however, that the code fails to satisfy the criteria for achieving optimal MAC-DMT set forth in [16], when . This alone does not mean that their code could not be optimal, as the criteria in [16] is sufficient, but not necessary (see [17] for justification of this claim). 3) Some explicit, algebraic code constructions for and

were introduced by Hong and Viterbo in [18]. A design criterion based on an approximation of truncated union bound was proposed. With such criterion they con-structed a code that outperforms in error performance the aforementioned (4 4) two-user code [13].

4) Badr and Belfiore [19] proposed another (4 4) two-user MIMO-MAC code which is obtained by adding a twist ma-trix to the (2 2) Golden ST code [20], [21] such that the

1In this paper, by an(m2n) code we mean a code consisting of (m2n) code

matrices, wherem is the number of transmit antennas required for transmission, andn is the number of channel uses. The number m can be either n or Kn , depending on the discussion. Whenm = n , the code is for each user’s use. Whenm = Kn , we mean the vertical concatenation of all users’ codes as an overall code. Notationn 2 n without parenthesis is used for the channel dimensions.

overall code matrix is nonsingular whenever all the subma-trices associated with each user are nonzero. However, be-cause of this additional matrix, the overall code matrix, though nonsingular, could be ill-conditioned at high-SNR regime, thereby resulting in a vanishing determinant, simi-larly as did their earlier one-antenna code [14] already dis-cussed above.

5) [22] addressed the problem of whether there exists a two-user MIMO-MAC code satisfying the nonvanishing deter-minant (NVD) property. This problem concerns whether the twisted Golden MIMO-MAC code [19] can be further improved to avoid the disadvantage of having a vanishing determinant. The answer is negative. [22] shows that if all the overall code matrices are nonsingular whenever the submatrices from each user are nonzero, then some of them must have determinant arbitrarily close to zero, i.e., have vanishing determinants.

6) By removing the matrix, [22] reported another code con-struction and proved its MAC-DMT optimality for and for any values of and . Computer simulations showed that this code outperforms the (4 4) code of [19] at all SNR values. Another important contribution reported in [22] was that, for the two-user MAC case, one does not need the whole code matrix to be nonsingular, and hence introducing the additional rotation matrix is not neces-sary from the MAC-DMT point of view.

7) In [16], Coronel et al. studied the optimal DMT perfor-mance of a selective fading MIMO-MAC and provided a sufficient criterion for designing MAC-DMT optimal codes for any and . Noting that the Rayleigh block fading channel is a flat fading channel, a simplification of their criterion requires the product concatenation of codes from any subsets of users to satisfy the NVD property such that the error probabilities associated with these sub-codes do not exceed the corresponding outage probability. However, as already pointed out in [22], such codes do not exist for the case of . A further investigation of their criterion can be found in [17].

B. Principal Results

A complete solution to the problem of constructing

MIMO-MAC codes for users that are MAC-DMT

op-timal in Rayleigh MIMO-MAC is presented in this paper. The paper consists of two parts. The first part provides the constructions of MAC-DMT optimal codes for the symmetric MIMO-MAC. The second part is on the code construction for the general MIMO-MAC where the users are allowed to have different numbers of transmit antennas and can transmit at dif-ferent levels of multiplexing gain.

C. Outline

A general result on the nonexistence of NVD MIMO-MAC codes is presented in Section II. This result suggests that the design criterion proposed by Coronel et al. [16] might be too strict to yield any MAC-DMT optimal codes. A relaxed design criterion is then provided in this section.

(4)

In Section III, we present a new code construction for the

symmetric MIMO-MAC for any , and . Several nice

properties of the proposed code are presented in this section. We prove that this newly proposed construction is MAC-DMT op-timal and meets the relaxed design criterion given in Section II. For ease of reading, the proof of MAC-DMT optimality is rele-gated to Section V.

In Section IV we investigate the MAC-DMT in a general MIMO-MAC where the users are allowed to have different num-bers of transmit antennas and transmit at different levels of mul-tiplexing gain. The exact general MAC-DMT in such channel will be given in Section IV-B, and it will be seen that unlike the symmetric case, some users in the general MIMO-MAC are no longer able to achieve their single-user performance even if the multiplexing gains of other users are extremely close to zero. The reasons for such unexpected result will be carefully explained therein. Finally, in Section IV-D the newly proposed code construction for symmetric channels will be extended to cater to the general MIMO-MAC. The MAC-DMT optimality of the generalized construction will be presented in Section VI.

II. RELAXED DESIGN CRITERION OFMAC-DMT OPTIMALCODES

In this section, we first present a rigorous, yet negative, re-sult on the nonexistence of a MIMO-MAC lattice code that has the NVD property. This result suggests that the design criteria proposed by Coronel et al. [16] might be too strict to yield any MAC-DMT optimal codes. Following this, a relaxed design cri-terion will be presented and will be met by all subsequent con-structions of MIMO-MAC codes in this paper.

Consider a symmetric MIMO-MAC with users, each having transmit antennas and communicating independently to the base station at the same level of multiplexing gain . Let , be respectively the space-time codes

used by the th user, , all satisfying the

power constraint (4). If independent Gaussian random code-books were used, i.e., the entries of code matrices are i.i.d. SNR random variables for all , Tse et al. [11] showed that the event of users in error has probability upper bounded by

SNR (8)

where is the event of users in outage. Note that the overall

error event . The union bound on

gives

(9)

Since the right-hand-side of (9) has a negative SNR-exponent equal to defined in (6), (9) proved the achievability of MAC-DMT claimed by Theorem 1 based on the argument of Gaussian random codebooks.

We next turn our attention to the deterministic ST codes. From the point-to-point perspective, it is known [4] that ST

codes satisfying the NVD property have the same error prob-ability as the outage events. Thus, for any MIMO-MAC code

, set

where

SNR

To see how is chosen, we offer the following insight. For each

, the code has size SNR so that it is of

multiplexing gain . An explicit construction of was given in [4] where the code is seen as a real algebraic ST lattice code with

dimension . Hence there are SNR PAM

signals selected from each dimension and SNR for all . Thus, the constant is chosen such that the code

satisfies the power constraint (4).

From [4], it is easy to prove the following theorem which in turn gives a sufficient criterion for designing MAC-DMT op-timal codes. We remark that this theorem is an alternative state-ment of the result given in [16] under certain restrictions, and we refer the interested readers to [17] for the connections.

Theorem 2 ([16]): Let be given as above. For

any , let be

the product concatenation of , defined by

.. .

If for all pairs of distinct code matrices ,

, the difference matrix ..

. (10)

satisfies , where by we mean the

Her-mitian transpose of matrix , then the codes are

jointly MAC-DMT optimal.

Proof: Note that the imposed condition implies that the

code satisfies the NVD property for any . Along sim-ilar lines as in [4], it can be shown that the error event associated with code , i.e., the error event of users in in error, has probability upper bounded by

SNR

where is the event of users in in outage. Now taking union bound over all possible as in (9) completes the proof.

Remark 1: The condition of for all is called the full NVD criterion and is actually equivalent to the criterion given by Coronel et al. in [16] with certain re-strictions, see [17] for details. It should be noted that this full NVD condition is only sufficient, not necessary. However, the following result suggests that this condition might be too strong

(5)

and precludes the existence of codes meeting the criterion. We

call the stronger condition the exactly

full NVD criterion.

Theorem 3: For any and for any , there do not exist any linear MIMO-MAC codes2that satisfy the exactly full

NVD criterion.

Proof: For ease of reading, the proof is relegated to

Appendix A.

Roughly speaking, the proof of Theorem 3 shows that while it is possible to construct DMT-optimal codes for each user, as the existing CDA-based ST codes [4] would do, it

is impossible for the product code to have an

exactly full NVD. Any such product code would have difference

matrices such that is extremely close

to zero at high-SNR regime. In terms of conventional rank and coding gain design criteria of ST codes, this means that even if the code achieves full diversity gain, it necessarily loses sig-nificantly in coding gain. Therefore, it becomes meaningless to say that the code achieves full rank and full diversity. We may conclude that the exactly full NVD condition is in practice too strict to yield MAC-DMT optimal codes.

Another implication from the proof of Theorem 3 is that the exactly full NVD condition can be met only if the users coop-erate in their transmission. Without cooperation, the exactly full NVD condition can never be met and the determinant must be vanishing.

On the other hand, we may relax the exactly full NVD con-dition without adversely affecting the DMT performance. To do so, we will partition the error events in a different manner. Given

the set of users , let , , denote the error

event when the users in are in error and the corresponding error matrix (cf. (10)) has rank exactly . Clearly event defined in the proof of Theorem 2 is a disjoint union of

. Now the codes are jointly

MAC-DMT optimal if the following holds.

Theorem 4 (Relaxed Design Criterion): Let be defined as above. Then they are jointly MAC-DMT optimal if the error events have probabilities upper bounded by

SNR (11)

for all and for all

. Furthermore, as for design of MAC-DMT optimal codes we require at least that

SNR (12)

for all and for all .

While (12) might be the most relaxed condition for designing MAC-DMT optimal codes, in this paper we will focus on con-dition (11).

2Here by linear codes we mean codes having linear dispersion forms [8] or

having a lattice structure. Almost all existing ST codes are linear, for example, the Alamouti codes [2], the CDA-based ST codes [3]–[7], [13], [18]–[21], [23], etc.

The rationale behind the above theorem is the observation that

the error probabilities SNR with are

not dominant in the overall DMT performance. Hence we could relax the conditions such that the event

has probability larger than the corresponding outage probability, but no larger than the dominant error probability. That is, we could allow

SNR (13)

but would still require SNR

Relaxation (13) would not affect the overall DMT performance. Compared with the exactly full NVD condition required by The-orems 2, Theorem 4 relaxes greatly the code design criterion in the following ways.

1) We do not require the difference matrix to be non-singular and to satisfy the NVD property when all the

com-ponent matrices are nonzero, which has been

shown to be impossible by Theorem 3.

2) Should the difference matrix happen to be singular, (11) requires the resulting error performance must be no

worse than SNR for some , , in

order to maintain the MAC-DMT optimality.

3) In Theorem 2, events with were required

to have probability absolutely zero. This is too strict and would preclude the existence of MAC-DMT optimal codes.

III. MAC-DMT OPTIMALCODECONSTRUCTION FOR

SYMMETRICMIMO-MAC CHANNELS

For the symmetric MIMO-MAC coded system with users, each having transmit antennas and transmitting at multi-plexing gain , in this section we will propose a systematic code construction that is MAC-DMT optimal for any combi-nations of , , , and . The construction does not assume any cooperation among the users. Furthermore, compared with the MAC-DMT optimal two-user code proposed in [22] where a sign change is required in the code matrices, here in the proposed method each user encodes his own information using an identical encoder. This greatly simplifies the hardware implementation of these encoders.

A. Proposed Construction

Given the number of users , let be the smallest odd integer such that , i.e.,

if even,

if odd. (14)

The construction calls for the following number fields. Let be a number field that is a cyclic Galois extension of

(6)

Fig. 2. Field extensions required by the proposed code constructions.

with degree , where . Let be

an-other cyclic Galois extension of with degree . Let and

be the generators of Galois groups and

with degrees and , respectively. The fields and are

chosen3such that . Let be

the compositum of and . See Fig. 2 for the relation among the required number fields. The readers are referred to [4], [22], [24] for the constructions of such number fields.

Let be a cyclic division algebra with

(15) where

(16) (17) for . The element is an indeterminate satisfying

, and is some suitable nonnorm element.

4 By we mean the complex conjugate of and is the

algebraic closure of in [26]–[28]. Notice that and is unimodular. It has been shown [5] that with such unimodular

, is always a cyclic division algebra.

Remark 2: While in the above we have set to be of form such that is unimodular, it might be possible that in some CDAs, the nonnorm element is actually an th root of unity for some integer and is already unimodular. See [29] for such example construction. Should it be the case, we could set , and the discussion below can be easily modified to show that the MAC-DMT optimality of the proposed constructions remains to hold. Therefore, for simplicity, here we will focus only on the case of .

Remark 3: We note that by construction the Galois groups of

the numbers fields are

3A more general condition on and is that the automorphisms and 

commute.

4A sufficient criterion for finding a suitable nonnorm element is given in

[25, Theorem 1]. Also, we refer the interested readers to [4, Theorems 10 and 11] for two explicit constructions of .

where in the last line denotes the direct product of the groups generated by and , respectively. It should also be noted that the automorphisms and commute, i.e.,

due to the direct product of two groups.

Given multiplexing gain , let SNR be the base alphabet defined as

SNR SNR SNR

then the corresponding information set is

SNR SNR

(18)

where is an integral basis of . It should

be noted that for SNR

and that SNR . Let

be the left-regular map that maps elements in into square matrices with entries in . Specifically, given with

is given by

..

. ... . .. ... (19)

Note that the field is the center of the division algebra ,

meaning that for any and .

Equiva-lently we have

showing that the matrix-product commutes.

Proposition 5 ([4], [26]): Let and be defined as above. Then

for all , where .

Having defined the above, the encoding of each user’s data stream proceeds as follows. Given the multiplexing gain , the th user first partitions his binary data steam into blocks of SNR bits. Then using the integral basis and set SNR defined above, each block of binary bits is mapped in a one-one fashion to a symbol

(7)

SNR . The encoding is performed indepen-dently at each user’s end.

Given SNR , the th user actually sends out the

following signal matrix through his transmit

antenna array in channel uses

(20)

where and where is a normalizing constant such

that

SNR SNR

Hence we have

SNR (21)

Remark 4: The above construction of the MIMO-MAC codes

is reminiscent of the multiblock ST code presented in [6]. Some key differences are highlighted below.

1) In the proposed construction we require the length of the

code to be where must be an odd integer.

2) The number fields and are required such that the automorphisms and commute. This was not needed in [6].

3) The element of the CDA must be unimodular, and we

have set .

We use the following example to illustrate the proposed con-struction.

Example 1: We consider the case of and . By construction is the smallest odd integer such that

. Then it can be shown that with and

the number fields and meet the required

conditions of , and .

Furthermore, we have . The generators

and for the Galois groups and are

given respectively by

The set is an integral basis for .

As the prime ideal of remains inert in and

, following from [4] this gives an appropriate nonnorm

ele-ment . Hence we have . With ,

is a CDA of index 2 which is also a central simple -algebra [26]. Next let

for with SNR . The Galois conjugates of

are for example given by

where and

. With the above, the signal matrix of the first

user is given by , where

SNR and

By vertically concatenating the signal matrices from all users, the overall MIMO-MAC code of the users is

..

. . .. ...

SNR

(22)

For ease of code performance analysis that comes later we set , i.e.,

..

. . .. ...

SNR

(23)

Remark 5: Below we briefly compare the proposed

construc-tion of with another MAC-DMT optimal code constructed for users in [22]. The latter MIMO-MAC code takes the fol-lowing form

SNR

(24)

The construction of requires a number field

with and such that

, and

. Here by “1” of we mean the trivial au-tomorphism. The field and the element are defined as be-fore. The element is taken from the cyclic division algebra for some indeterminate . SNR is the base-in-formation set defined similarly as SNR in (18). Thus, com-pared with the present proposed construction, we see that requires an additional sign change at the second block matrix of the second user’s code. This sign change is essential to en-sure an NVD-like property. It also endows with another nice property that the transmission of code matrices in takes only channel uses, less than that required by . However, this additional sign change might complicate system design as the system must constantly check which user requires a sign change and which user does not. Such disadvantage does not exist in the proposed construction of . Everything works perfectly after patching an extra block of transmission when is even. An-other drawback of is the difficulty of generalization to the

cases of .

Let be the channel matrix of the th user. We

(8)

(5), given the overall transmitted code matrix , the re-ceived signal matrix at receiver end is

(25) is the noise matrix whose entries are i.i.d. random variables, and is the th block received signal matrix given by

and

B. Properties of the Proposed Construction

To simplify the analysis of code performance, below we de-fine the extended versions of and .

..

. . .. ...

SNR (26) (27)

Given the overall signal matrix , let be any

signal matrix such that the upper submatrix of

equals . Then we can rewrite (25) as

(28) where

if odd, if even.

By we mean the all-zero matrix of proper size. Noting (25) and (28) are equivalent, henceforth we will work only with the extended codes and , rather than and . We next show several nice properties possessed by and .

Property 1: For any , we have

(29)

Proof: We first claim

(30) To see this, notice that

..

. . .. ...

..

. . .. ...

where the last equality follows from the fact that is even, hence the claim (30) is proved. Next, we show

(31) To this end, define

(32) where is the indeterminate defined as in (15). Since from (17)

for all , it is clear that ,

where . Now we have

.. . . .. ... .. . . .. ... . .. .. . . .. ... . ..

where we have used the fact that since

by construction. Thus, as is fixed by both and , we

see that .

Finally, from the definition of (19), the matrix

. ..

has entries in for all and

since SNR

is the ring of algebraic integers in number field . It then follows that

Summarizing the above results, we conclude that

(9)

Property 2: Let .. . ... . .. ... (33) and .. . . .. ...

with , SNR , where by we mean the

transpose of vector . Let be the maximal number of rows in that are linearly independent as a left -module; then

(34) where the rank is measured in the complex number field .

Proof: To find out the rank of matrix , we use the el-ementary row operations from Gaussian elimination method. Note that the same row operations can be performed on whose entries are in . Extra care must be taken because multiplica-tion in is noncommutative. Further, we note that elementary row operations on are equivalent to the block elementary row operations on . By this we mean that, say is a

elementary matrix with entries in ; then it is clear

where is the natural extension of to the central

simple matrix algebra over [26], i.e.,

(35)

From hypothesis, assume is the maximal

subset of the rows of that are linearly independent over . Then it follows that there are leading ones in the row-reduced matrix of . Equivalently, the same block elementary opera-tions would reduce matrix into a matrix whose main diagonal consists of identity matrices, each of size , after permuting the columns if necessary. This completes the proof.

Property 2 shows that the overall code matrix might not always have full rank , and the rank of is always a multiple of . This is not too much of a surprise as it is straight-forward to see that in (26) if some ’s are identical, then the overall code matrix cannot be nonsingular.

Compared with the constructions proposed in [18], [19], the matrix of the present construction could be singular even when the component matrices are all distinct and nonzero as shown by Property 2. Nevertheless, we will prove in Section V that in order to achieve the optimal MAC-DMT performance at high-SNR regime, it is unnecessary to construct codes such that is nonsingular whenever all the component matrices are distinct and nonzero.

Before rigorously proving the above statement, a heuristic way to see this is the following. Since the users communicate

independently to the base station, for any overall MIMO-MAC code it is impossible for all the code matrices to be non-singular as some component matrices of the th user could be zero. Also, from the pairwise error probability point of view,

for any , can be singular at least when the

information symbols transmitted by some users are the same. The rank of overall code matrices is at best a multiple of . Therefore, intuitively speaking, perhaps it would not hurt to make things a bit worse in the sense that the difference matrix can be singular in other cases. By this we mean that if there are distinct information symbols in the difference

ma-trix , the maximal possible rank of is . We

claim that it would not hurt in the DMT sense if the

construc-tion can provide only rank for some with .

The reason for this actually follows from Theorem 4 that the error events of users in error but getting only rank distance do not dominate the error performance in the final DMT performance. Therefore, we strongly speculate that such difference matrices do not have to achieve the same rank as the Gaussian random code does. The rank can be less, as long as the resulting error performance is not worse than those

of and .

Although we do not need the whole code to satisfy the full NVD property as in the point-to-point scenario, an alternative NVD-like property is preferred and is given as below.

Property 3: Let be defined as in (33) and assume that is a subset of rows of that are linearly in-dependent as a left -module. Define

..

. (36)

i.e., is the submatrix of consisting of the corresponding linearly independent rows, where is the natural exten-sion of . Then

(37) where by we mean the hermitian transpose of matrix .

Proof: First, it follows from Property 2 that

since has full row rank and by assumption.

To show , we shall first verify that

is fixed under automorphisms and . For , it can be seen from the proof of Property 1 that

and

..

(10)

..

. . .. ...

for some column permutation matrix of size ,

where and ,

. Now it follows that

as , and we have proved is fixed by

.

For , again from the proof of Property 1 we see that

and .. . . .. ... . .. . .. where .. . ... . .. ... ... (38)

From (38) it is clear that since by

con-struction. Therefore, we see that

Taking into account that it follows

that . So far, we have proved that

is fixed by both and . This in turn implies that . Finally, the proof is complete after noting that has entries in .

In Property 2 we have shown that the overall code matrix might not have full rank, and when that happens, its rank always equals for some . The number indicates the number of users whose transmitted signal vectors, when regarded as rows of matrix in (33), are linearly independent over . Further, Property 3 shows that even when is singular and fails to have

NVD, i.e., fails to satisfy , the submatrix

formed by the transmitted signal matrices of those users still

satisfies the NVD property. Such result can be further extended to yield the following property on the nonzero eigenvalues of

.

Property 4: Let and be defined as above with

. Let be the

nonzero eigenvalues of . Then

(39)

Proof: Here we take an information theoretic ap-proach to prove the first inequality. To this end, let be a complex Gaussian random vector of length with zero mean and covariance matrix

Without loss of generality we can assume that linearly inde-pendent users are the first users and corresponds to the th

user, . Hence the covariance matrix of the

subvector equals

We have the following inequality for the differential entropies

of and

(40) Notice that the covariance matrix of can be decomposed as

for some unitary matrix . is a diagonal

matrix whose nonzero entries are the ’s. Thus setting we have

Now combining the above results proves the first inequality in (39). The second inequality in (39) follows directly from

Prop-erty 3 and from .

Remark 6: The above property shows that despite can be singular, the product of the nonzero eigenvalues of is al-ways bounded from below by 1. This can be regarded as a re-laxation of the conventional NVD property. In the design of ST codes, satisfying the NVD criterion is a sufficient condition to achieve the optimal point-to-point DMT performance. To guar-antee NVD in the point-to-point MIMO, we require all the users to cooperate fully as already seen in Theorem 3. However, it is not allowed in MIMO-MAC where users transmit indepen-dently their own information to the common receiver. Thus, in MIMO-MAC we do not demand full NVD, and only partial NVD is required as shown in (39).

(11)

C. MAC-DMT Optimality of the Proposed Construction

Armed with the properties discussed in the previous section, below we are able to show the proposed code is MAC-DMT optimal.

Theorem 6: Given multiplexing gain , the proposed code

defined as in (22) achieves the following diversity gain (41)

over Rayleigh block fading channel with channel coherence

time channel uses. Thus, is MAC-DMT optimal.

Proof: The proof is relegated to Section V for ease of

reading.

IV. MAC-DMT OPTIMAL CODES FORGENERAL

MIMO-MAC SYSTEMS

In [11], Tse et al. focused on analyzing the DMT in a sym-metric MIMO-MAC system. By symsym-metric we mean that every mobile user in the system has the same number of transmit an-tennas and transmits at the same level of multiplexing gain. However, the symmetric MIMO-MAC might not be practical enough. In the near future, the mobile communication is likely to be at a transition stage, migrating from conventional SISO (single-input single-output) to MIMO. In fact, such transition already takes place in wireless local area networks where some old laptops have single transmit antenna while the latest ones could have more than two transmit antennas. In the mixture of SISO and MIMO communication environment, one would ex-pect the mobile users having different numbers of transmit an-tennas. Furthermore, in practice it is often possible that mobile users transmit at different rates because of the different plans they purchase from the service provider. The different rate im-plies a different level of multiplexing gain in the DMT sense. It is then of fundamental importance that we must have a gen-eral code construction that works for any MIMO-MAC systems where the mobile users are allowed to have different numbers of transmit antennas and can transmit at different levels of multi-plexing gains. In the previous sections we have provided a sys-tematic construction for the symmetric MIMO-MAC and have proved that it achieves the optimal MAC-DMT. Below we will extend these results to the general channel.

A. Decoding in General MIMO-MAC

There can be at least two decoding methods in the gen-eral MIMO-MAC, depending on how much computational complexity one can afford. The first decoder is the joint ML decoder, by which we mean the following. Assuming there are users, each transmitting using a codebook that consists

of ST code matrices, for . Let

be the signal matrix transmitted by the th user, and let

be the received signal matrix; then the joint ML decoder seeks

the optimal joint ML estimate by

(42)

where . This joint ML decoder

was used in [11] for analyzing the MAC-DMT performance in symmetric MIMO-MAC.

However, the above joint ML decoder might not be optimal in terms of the error performance of each user. For the th user, the truly optimal decoder, though having extremely high com-putational complexity, is the individual ML decoder that seeks optimal ML estimate by

(43)

where and

. The difference be-tween the individual and joint ML decoders is analogous to that between the BCJR and Viterbi decoders [30] for the decoding of convolutional codes. It is easy to see that the individual ML decoder always outperforms the joint ML decoder.

In the next two subsections we will examine the MAC-DMT performances of these two decoders. Obviously we expect there might exist certain performance loss in the joint ML decoder, compared to the individual ML decoder.

B. MAC-DMT for General MIMO-MAC With Joint Decoding

Consider a general MIMO-MAC system with mobile users. Let denote the number of transmit antennas of the

th user, , and let be the corresponding

multiplexing gain. Assuming receive antennas at the base station, the first major result of this section is the following.

Theorem 7 (General Joint MAC-DMT): Let , , and be defined as above. If joint decoding is performed at receiver end, the optimal MAC-DMT of such system is given by

(44) for i.i.d. Rayleigh block fading channel that is fixed for at least

The minimization in (44) is taken over all possible nonempty

subsets , and

(45) is the total number of transmit antennas of users in . The notion of is the conventional point-to-point DMT.

(12)

Fig. 3. Joint MAC-DMTd (r ; r ) of general MIMO-MAC with two users.

Fig. 4. Joint MAC-DMTd (0; r ) of general MIMO-MAC with two users.

Prior to proving Theorem 7, we shall give an example illus-trating this theorem and in particular, show some unexpected effects resulting from joint decoding.

Example 2: For simplicity, here we consider a general

MIMO-MAC system with two users. The first user has transmit antenna and transmits at multiplexing gain ; the second user has transmit antennas and transmits at multiplexing gain . Assume there are receive an-tennas at receiver end. Using (44) the resulting MAC-DMT is shown in Fig. 3. First, it is interesting to note that unlike the symmetric MIMO-MAC where all users have same number of transmit antennas and transmit at same level of multiplexing gain, here the second user cannot achieve his single-user DMT performance even when . This effect is shown in Fig. 4. While this is quite unexpected, such phenomenon can be easily explained. Recall that the DMT is an asymptotic result. Strictly speaking, the multiplexing gain is defined as

SNR SNR

and is the actual transmission rate. Therefore, when we say it does not necessarily mean . It simply means that the rate of the first user grows much slower than SNR. For example, an ST code that is fixed and does not vary with

SNR has multiplexing gain 0 since the rate is a constant. But the rate is bounded away from 0.

Having learned the above, in our example given the multi-plexing gain for some positive very close to 0, the DMT performance of joint decoder would be dominated by erroneous decoding of the first user’s signals when is small. It is also easy to confirm this observation from pairwise error probability

(PEP) analysis. Assume , but , i.e., the

codes are fixed and do not vary with SNR. Since the two users do not cooperate, for any distinct pairs of overall code matrices, the maximal possible rank is the minimum of and . Hence the resulting maximal possible diversity gain equals

which equals 2 in this example. Therefore, the PEP analysis confirms that the single-user DMT performance cannot be achieved for small values of as shown in Fig. 4.

Before concluding this example we remark that the loss in DMT for the second user can in fact be recovered if an individual ML decoder is used. We will come back to this in Section IV-C. The proof of Theorem 7 follows along similar lines of that of symmetric MAC-DMT provided by Tse et al. in [11]. Specifi-cally, let

SNR

denote the actual transmission rate of the th user. Given the subset of users, let denote the following outage event

(46) where

• is the overall channel matrix, is

the channel matrix of size of the th user, • is the total number of transmit antennas defined by

• contains the transmitted signal vectors of users in and is defined as

• is the received signal vector given by

where is the complex Gaussian random noise vector, and • consists of transmitted signals of users not in . Let denote the overall outage event. It is clear that

(13)

Following similar arguments as in [11] it is straightfor-ward to see that the error probability of joint decoding

is lower bounded by

SNR

(47) To establish the converse, we take the random codebook ap-proach similar to that used by Tse et al. in [11]. Let be the code-book of the th mobile user, consisting of code matrices that are randomly generated by some complex Gaussian random generator. Further, satisfies the desired multiplexing gain,

SNR

Let denote the event that the signal matrices of users in are erroneously decoded by the joint decoder. Then arguing similarly as in [11], it can be shown that

SNR whenever

Thus, using union bound we have

SNR provided that

This proves Theorem 7.

C. MAC-DMT for General MIMO-MAC With Individual ML Decoding

In the previous section we investigated the MAC-DMT for a general MIMO-MAC with joint decoding at the receiver end. We also observed in Example 2 that certain DMT performance loss could result from the use of joint decoder. However, such loss can be safely avoided by the use of individual ML decoder. Recall that for the th user, the truly optimal decoder, though having extremely high computational complexity, is the indi-vidual ML decoder that seeks optimal ML estimate by

(48)

where and

. Clearly (48) outperforms (42) in error performance, but at a cost of much higher computational complexity.

Without loss of generality, below we focus on the error per-formance of the individual ML decoding for the th user. To dis-tinguish the DMT performances of decoders (42) and (48), we shall call the DMT of the latter the individual MAC-DMT and

will denote it by .

To characterize the DMT performance of the individual ML decoder, we only need to consider the outage events (cf. (46)) in which the th user is a member of . Event with is not counted as an outage for the th user for obvious reasons. Thus, along similar lines as in the proof of Theorem 7 we can show the following.

Theorem 8 (General Individual MAC-DMT): Let , , and be defined as before. If individual ML decoding is per-formed at receiver end for the th user, the optimal individual MAC-DMT is given by

(49) where the minimization is taken over all

under the condition and is defined in (45).

Proof: For brevity we only outline the proof. Let denote the outage event of the th user; then following from the above discussion it can be seen that

since if , the th user is not in outage. Now let denote the error probability of the in-dividual decoder for the th user; then it can be shown that

SNR

where the first inequality follows from [9, Lemma 5]. To show the converse, let denote the error event that the signal ma-trices of the users in are erroneously decoded under joint de-coding. Then simply note that the error probability of an indi-vidual ML decoder is upper bounded by that of a joint ML de-coder, i.e.,

where the right-hand-side gives the probability of a joint ML decoder when the signal of the th user is erroneously decoded.

(14)

Fig. 5. Comparison between the joint MAC-DMT and the individual MAC-DMT of the second user whenr = 4r = r.

Now using the union bound argument and along similar lines as in the proof of Theorem 7 it can be shown that

SNR

SNR This completes the proof.

With the above result, we now come back to Example 2 to investigate the individual MAC-DMT of the second user.

Example 3 (Continued From Example 2): In Example 2 we

have considered the specific case of , , ,

and . Assuming the second user transmits at mul-tiplexing gain , from Theorem 8 the individual MAC-DMT of the second user is

Hence we see that the single-user performance of the second user is recovered by the use of an individual ML decoder. To illustrate further the difference in MAC-DMT between (42) and (48), in Fig. 5 we compare the MAC-DMT performances of joint

and individual decoders at . It can be clearly

seen that the individual ML decoder outperforms significantly the joint ML decoder at low-multiplexing-gain regime.

Another comparison between the DMT performances of both decoders at is given in Fig. 6. It shows that the

joint ML decoder (given by ) is not optimal for

the second user. The truly optimal individual ML decoder for

the second user has DMT performance .

Further-more, the individual ML decoder for the second user achieves

the single-user DMT performance as long as .

On the other hand, for the first user who has lesser number of

Fig. 6. Comparison between the joint MAC-DMT and the individual MAC-DMT whenr = r = r.

transmit antennas, the DMT performances of the joint and in-dividual decoders are the same and are actually equal to his single-user performance .

Next we could apply Theorem 8 to the case of symmetric MIMO-MAC to see how the error probabilities of joint and individual ML decoders compare. The comparison is given in the following corollary. It shows that in the symmetric MIMO-MAC there is no difference in terms of MAC-DMT performance between the joint and individual ML decoders.

Corollary 9: For symmetric MIMO-MAC with users, each having transmit antennas and transmitting at multiplexing gain , let denote the error probability of the joint ML decoder and denote the error probability of the th individual ML decoder. Then

and in terms of DMT we have

for all .

Proof: It suffices to show only the equality in DMT. First,

from Theorem 8 we have

and the proof is complete after noting that the right-hand-side of the above is the same as the MAC-DMT given in Theorem 1.

Before concluding the subsection we have the following re-marks. First, while the individual ML decoder could achieve a much higher DMT performance as seen in Examples 2 and 3, the computational complexity required by (48) is often extremely high. Thus, the individual ML decoder has widely been consid-ered as being impractical in multiuser detections. The reason for including this receiver is only to clarify the unexpected DMT performance loss of the joint ML decoder in Example 2.

As the individual ML decoder is rarely used, below we will not consider this receiver anymore. We will regard the joint MAC-DMT given in Theorem 7 as the optimal MAC-DMT in

(15)

practice, although it is now clear that it is not the best one can actually achieve.

D. General MAC-DMT Optimal Codes

So far we have provided the optimal MAC-DMT (44) for the general MIMO-MAC system with users where the th user has transmit antennas and transmits at multiplexing gain . To have a deterministic code for the general MIMO-MAC, we can extend the code construction given in Section III for sym-metric MIMO-MAC to the present case.

Let , , , and be defined as before. For brevity we only present the construction when is odd. Codes for even can be constructed by simply patching an extra coded block to each user’s code matrices, similar to that described in Section III. Henceforth we will drop the subscript “ ” in , , , ,

, etc. for simplicity.

Given and , we first define

(50)

as the maximal number of transmit antennas among all users. In general, the number can be either preknown to all the users, or explicitly specified among any groups of users. Next, let be the number field that is cyclic Galois over

with degree , and let be another cyclic

Galois extension of with degree . Let be the generator of the Galois group , and similarly let be the generator of . The fields and are required to satisfy

or are required such that and commute. Finally, we set to be the compositum of fields and . Similar to Section III, with some suitable unimodular , we have

(51) as an appropriate central simple division -algebra with for , where is an indeterminate satisfying .

Given the multiplexing gain of the th user, we set the cor-responding base alphabet and information set as follows:

SNR SNR SNR

(52) and

SNR SNR

(53)

where is an integral basis of . Unlike

the construction for symmetrical MIMO-MAC, here the infor-mation set can be different among users as each user has dif-ferent level of multiplexing gain.

Let denote the left-regular map of elements in into

ma-trices of size whose entries are in (similar to

of (19)); then the ST code of the th user is given by

SNR

(54) where

SNR (55)

such that the power constraint (4) is satisfied.

Given , , the overall code is obtained

by vertically concatenating the code matrices from each user,

..

. (56)

The overall code matrix is a square matrix of size . Below we will present some nice properties of which are essential to proving its MAC-DMT optimality.

The first property extends Property 2 of the symmetric MAC code in Section III-B.

Property 5: For any SNR , define ..

. ... . .. ... (57)

and let

..

. . .. ...

be the corresponding overall code matrix where .

Then SNR. Further, let be the maximal number of

rows of that are linearly independent as a left -module. Then (58) where the rank is measured in the complex number field .

Proof: The first claim can be easily verified from the

set-tings of and SNR , and is thus omitted for brevity. For the second, to determine the rank of , it suffices to consider the rank of the unscaled code matrix

..

. . .. ... (59)

Notice that is a code matrix of the code defined in (27) for

the symmetric MIMO-MAC when we set and

(16)

The next property generalizes Property 3 in Section III-B where we were interested in the Gram determinant of the un-scaled code matrix. Here, for the purpose of analyzing the general MAC-DMT performance of the proposed code, we will seek directly the Gram determinant of the overall matrix .

Property 6: Let be defined as in (57) and assume that is a subset of rows of that are linearly in-dependent as a left -module. Let

..

. . .. ... (60)

be the submatrix of consisting of the corresponding rows. Then

SNR (61)

Proof: Arguing similarly to the proof of Property 5, set

..

. . .. ... (62)

Then we have

. .. and

by Property 3 since is a submatrix of the code matrix (cf. (59)) of the code (cf. (27)) for the symmetric MIMO-MAC

when setting and . The result now

follows from

and from the definition of in (55).

The two properties above are exactly what we need to prove the MAC-DMT optimality of the proposed general MIMO-MAC code in (56). Hence, with these properties we can prove the following theorem.

Theorem 10: Given and , with odd, the proposed code defined in (56) achieves the general joint MAC-DMT

(63)

over a Rayleigh block fading channel that remains static for at

least channel uses. is MAC-DMT optimal.

Proof: The proof is similar to that of Theorem 6 and is

relegated to Section VI for ease of reading.

The proof to Theorem 10 can in fact be further extended to show that the proposed code (56) achieves the optimal indi-vidual MAC-DMT(49), provided that an indiindi-vidual ML decoder for each user is used at the receiver end. This result along with the proof will be presented in Corollary 12 of Section VI.

V. PROOF OFTHEOREM6

Here we only prove the case of odd. The case of even

and can be proved using similar arguments, and

will therefore be briefly handled in a remark following the proof.

A. Proof Overview

In this subsection we provide an overview of the proof for the case of odd, along with a few insights to the proof. Given the overall channel matrix

(64) we will provide an upper bound on the codeword error proba-bility of the joint decoder at receiver end. Let

be a subset of users, and let denote the event that 1) the signal of the th user is erroneously decoded if and only

if , and further that

2) the rank distance between overall transmitted code matrix and the erroneously decoded overall signal matrix is only

for some .

Specifically, let SNR denote the information symbol transmitted by the th user, and let be the corresponding de-coding output at receiver; then the event can be formu-lated as follows:

(65)

where from the proposed construction (cf. (27)) we have ..

. . .. ...

and

..

. . .. ...

Note that the difference matrix has exactly nonzero rows, and by Property 2 we see the rank distance

. Hence it makes sense

for the second requirement of error event that

(17)

Thus, it can be seen from the union bound argument that the codeword error probability is upper bounded by

(66)

The event is a further partition of the event considered by Tse et al. in [11]. We discuss this in more detail in the fol-lowing remark.

Remark 7: With regard to the Gaussian random codebook

considered by Tse et al. [11], it is straightforward to see

is empty with probability one if , since the compo-nent matrices associated with each user are complex Gaussian

random matrices of size for some . In other

words, if for all and otherwise, then

the error matrix would have rank with probability one. Therefore, one can rewrite (66) as

(67)

and recover the same union bound used in [11].

Unlike [11] where the authors analyzed each summand of (67) by a union bound argument with a Gaussian random codebook, here we will focus on the error probability of a deterministic codebook (cf. (27)), and

at-tempt to upper bound the probability by using a

joint ML decoder. To this end, in Section V-B we will examine the minimum Euclidean distance among the noise-free received code matrices contained in . It should be noted that here by minimum Euclidean distance, we mean the minimum Euclidean distance among only the pairs of code matrices in , not the whole code . Thus, the minimum Euclidean distance will be a function of , , and .

Once we obtain the minimum Euclidean distance, we will analyze the error performance of a bounded distance decoder, which will be used as an upper bound on that of the ML de-coder. The bounded distance decoder results in an error only when the noise matrix has norm larger than half of the min-imum Euclidean distance. More precisely, let be the overall channel matrix defined in (64) and be known to the decoder; let be the overall MIMO-MAC code, where is the codebook of the th user. The minimum Euclidean dis-tance among all code matrices in is defined as

which is dependent upon . Given the received signal matrix , the bounded distance decoder outputs

if , and declares a decoding failure

oth-erwise. Thus, only the received signal matrices that are within distance from the original transmitted overall code ma-trix can be correctly decoded in the bounded distance decoder. Other received signal matrices would result in either a decoding

error (i.e., decoding into an erroneous code matrix) or a de-coding failure (i.e., cannot find a code matrix within distance ). Though this decoder is suboptimal compared to the ML decoder, its error performance can be mathematically ana-lyzed.

The error performance analysis following this outline will be given in Section V-C. Finally, in Section V-F we briefly discuss the proof for the case of even .

B. Lower Bounds on the Minimum Distance Among Noise-Free Received Signal Matrices

For any with

..

. . .. ...

and

..

. . .. ...

let , , and let , where

is the channel matrix associated with the th user. Given the channel matrix , below we provide a lower bound on the squared Euclidean distance between and , i.e.,

(68)

We distinguish the following two cases which correspond to the

error events and with , respectively.

1) For event we have for

, otherwise, and

. In this case, let and be defined as in (36) and let

be the equivalent channel matrix; then we have

Let be the set of ordered nonzero

eigenvalues of where , and

let be the ordered nonzero

eigenvalues of . Then we have

(69) Note that

(70) where the first inequality follows from Property 3, and the second exponential equality is because is fixed and is independent of SNR.

(18)

By repeatedly using the arithmetic mean-geometric mean

inequality and (70) as in [4], [6] for , we

have

(71)

(72)

SNR SNR

SNR (73)

where (72) follows from (70) and where in (73) we have set

SNR

Hence the SNR exponent of is lower bounded

by

(74)

2) The second case corresponds to event which

means for ,

otherwise, and . In other

words, the nonzero rows

are not linearly independent over . From Property 2 we can assume without loss of generality that

are linearly independent for some .

Let and let and be defined as in (36)

with respect to the set . Set

and . Property 2 in turn implies that

. .. ..

. ... ...

(75)

for some square matrices , where ..

. . .. ... (76)

Similar to the previous case, let

be the equivalent channel matrix; then the difference of the noise-free received signal matrices can be rewritten as

(77) where

(78) is an alternative channel equivalent matrix and

(79)

for .

Let be the set of ordered nonzero

eigenvalues of , where , and

let be the ordered nonzero

eigen-values of . Notice that

from Property 3. Arguing similarly as in the first case shows that SNR (80) for , where SNR (81) (82) and (83)

(19)

Remark 8: We remark that (83) shows the last term is , instead of being as in (74). For readers who may wonder why these two terms are different given both events concern the case of users in error, the major reason is due to the distance bounding techniques, i.e., the repeated arithmetic mean-geo-metric mean inequalities, we have used in the above.

In general, when the equivalent channel matrix of (78), and similarly when the channel matrix with , has

rank , the rank of the product matrix would be

since is of full rank . Thus our lower bound on

the norm would only capture the smaller

eigen-values of , which are all nonzero. Furthermore, one reason for introducing the equivalent channel matrix , rather than working with is that, algebraically speaking, the norm could be zero as is singular and all the rows of could lie in the left-null space of . However, since is random, this occurs with probability zero. In other words, if we apply the series of arithmetic mean-geometric mean inequalities to the matrix product , we could end up with the trivial algebraic inequality

even the right-hand-side has probability 0. Whether the above could happen depends on the relations among , , , and . While there is nothing wrong with the algebraic inequality itself, this bound can actually be further tightened by introducing the equivalent channel so that we can focus on error events that have probability larger than zero.

Remark 9: Another heuristic way to see why the last term of

equals follows from the base-alphabet SNR de-fined in Section III-A. Recall that in the construction of

CDA-based ST code for point-to-point channel [4], [6], to achieve the DMT optimality therein we would set the base-al-phabet as

SNR SNR SNR

such that the resulting exponent equals

then along the same lines as in [4], [6] one can prove such code is approximately universal and achieves diversity gain . However, it is because we set the base-alphabet as SNR , which has size

SNR SNR

meaning an -fold increase in the multiplexing gain, we expect the error probability associated with event has diversity

gain .

C. Upper Bounds on Codeword Error Probability

Having obtained the squared minimum Euclidean distances among the signal matrices associated with error

event , and among the signal matrices

associated with error event , below we proceed to an-alyze the error performance of the proposed construction. The analysis resembles the sphere bounding technique used in [4], [6] which is essentially a bounded-distance decoding technique. That is, the bounded-distance decoder declares an error only when the noise has norm larger than half of the minimum Eu-clidean distance. Clearly, the error performance of a bounded-distance decoder serves as an upper bound on that of a joint ML decoder.

First, since the lower bounds on the Euclidean distance hold for all , we define

Then, using the bounded distance decoder discussed in Section V-A, the probability of error event given channel matrix can be upper bounded by

(84) where the inequality follows from the property of a bounded

distance decoder. Hence we see that if

. On the other hand, we may replace the above

upper bound of with the trivial upper bound

when . Thus, it implies

(20)

Since the above bounds do not depend on the specific choices of , from (66) the union bound on the codeword error

prob-ability gives

(85)

Remark 10: One can regard the probability

as a further upper bound on the union bound

in (66), and the second type of probability

(86)

as an upper bound on . Furthermore,

the event of users in error has probability upper bounded by

(87) It should be noted that in (86) we have over-estimated the number of choices of -linearly independent rows out of nonzero rows in the difference matrix

that can happen in the event . Even with this over-estimate, noting

for all within the range of interest, we can rewrite (85) as

(88) Below we investigate the diversity orders of each term in (88).

D. Diversity Gain of the First Case

For each , , we have

(89)

SNR SNR

SNR (90)

where , , and

where for . Equation (89) follows

from [31]–[33], and (90) is given in [9] since is a matrix of

size having entries that are i.i.d. complex

Gaussian random variables. The quantity

repre-sents the point-to-point DMT of an MIMO Rayleigh fading channel at multiplexing gain .

E. Diversity Gain of the Second Case

Similarly, for the second set of maximizations in (88) we have

for each that

SNR SNR

(91) where is defined in (78) and is of size . Noting that the entries of are correlated complex Gaussian random variables, we invoke the following result which was shown inde-pendently in [34, Corollary 1] and [35, Theorem 3] to simplify the analysis.

Theorem 11 ( [34], [35]): The diversity order of outage

prob-ability for Rayleigh fading channels with arbitrary full rank cor-relations is unchanged from the case of i.i.d. Rayleigh fading. Moreover, if the channel matrix can be decoupled as

where has independent and regular entries, then the optimal DMT for channel is the same as that for .

Armed with Theorem 11, the analysis of the diversity gain of the second case is now easy. A direct application of the above theorem gives

SNR SNR

(21)

Summarizing results of (88), (90) and (92) gives SNR

and

This completes the proof.

Remark 11: The above proof shows that

SNR

for error events , , and

SNR

for . This is exactly what is shown in Theorem 2. Furthermore, in the event of users in error, the proposed code has error probability

SNR SNR

F. Proof Outline for Even

The proof of Theorem 6 can be adapted to cater to the case when the number of users is even. Here we discuss only briefly what the changes are. Firstly, with

in mind, i.e., , the result (68) of the squared Euclidean distance between and remains to hold. Similarly, the fur-ther lower bounds on in (74) and (83) stay without changes except that one should keep the following in mind.

1) The parameter of the first case, where

, and out of ’s are distinct, has value from 1

up to . This is because , and we

can always assume without affecting the

value of . Thus the diversity gain resulting from the first case is

Compared with the case of odd , (90) has up to . 2) The parameters and in the second case can be argued

similarly as the above, and we have

. Hence the diversity gain of this case is

Overall, it shows the MAC-DMT optimality of the proposed construction remains to hold.

VI. PROOF OFTHEOREM10

The proof of Theorem 10 is similar to that of Theorem 6. Therefore, we will skip the most of the details and highlight only the key differences.

First, for any subset

of users, again let denote the event that the decoder has made an error in decoding users signals, but only the rows formed by the difference signal matrices of some users, say user , are linearly independent over . In other words, let be a pair of distinct information symbols of

the th user for . Set the submatrices and as in

(62) with and . Then the event

corresponds to the case when .

Let denote the channel matrix of the th user

that is known completely to receiver. Since the code matrices

of the th users are of size , here we

will assume without loss of generality that only the first rows of are used for transmission via the transmit antennas of the th user, and the remaining rows are discarded during either encoding or transmission. On the other hand, we could extend the channel matrix to an equivalent channel matrix of size by adding on the right an all-zero matrix with appropriate size. That is, we set

and the received signal matrix can be written as

where is the noise matrix of size .

For the event , let

be the overall equivalent channel matrix, and let

be the ordered nonzero eigenvalues of with

Similarly, let be the ordered

nonzero eigenvalues of . Then the

min-imum squared Euclidean distance between and is

bounded by

SNR

(93)

數據

Fig. 1. The MAC DMT for K = 3 users with n = 2 and n = 2.
Fig. 2. Field extensions required by the proposed code constructions.
Fig. 3. Joint MAC-DMT d (r ; r ) of general MIMO-MAC with two users.
Fig. 5. Comparison between the joint MAC-DMT and the individual MAC-DMT of the second user when r = 4r = r.

參考文獻

相關文件

An additional senior teacher post, to be offset by a post in the rank of APSM, is provided to each primary special school/special school with primary section that operates six or

Wang, Solving pseudomonotone variational inequalities and pseudocon- vex optimization problems using the projection neural network, IEEE Transactions on Neural Networks 17

For the proposed algorithm, we establish a global convergence estimate in terms of the objective value, and moreover present a dual application to the standard SCLP, which leads to

Define instead the imaginary.. potential, magnetic field, lattice…) Dirac-BdG Hamiltonian:. with small, and matrix

If the best number of degrees of freedom for pure error can be specified, we might use some standard optimality criterion to obtain an optimal design for the given model, and

Corollary 13.3. For, if C is simple and lies in D, the function f is analytic at each point interior to and on C; so we apply the Cauchy-Goursat theorem directly. On the other hand,

Corollary 13.3. For, if C is simple and lies in D, the function f is analytic at each point interior to and on C; so we apply the Cauchy-Goursat theorem directly. On the other hand,

[r]