Rate-Splitting Multiple Access for Overloaded Multi-group Multicast: A First Experimental Study

Xinze Lyu, , Sundar Aditya,  and Bruno Clerckx This work was supported in part by UKRI Impact Acceleration Account (IAA) grant EP/X52556X/1.X. Lyu, S. Aditya and B. Clerckx are with the Dept. of Electrical and Electronic Engg., Imperial College London, London SW7 2AZ, U.K. (e-mail: {x.lyu21, s.aditya, b.clerckx}@imperial.ac.uk).
Abstract

Multi-group multicast (MGM) is an increasingly important form of multi-user wireless communications with several potential applications, such as video streaming, federated learning, safety-critical vehicular communications, etc. Rate-Splitting Multiple Access (RSMA) is a powerful interference management technique that can, in principle, achieve higher data rates and greater fairness for all types of multi-user wireless communications, including MGM. This paper presents the first-ever experimental evaluation of RSMA-based MGM, as well as the first-ever three-way comparison of RSMA-based, Space Divison Multiple Access (SDMA)-based and Non-Orthogonal Multiple Access (NOMA)-based MGM. Using a measurement setup involving a two-antenna transmitter and two groups of two single-antenna users per group, we consider the problem of realizing throughput (max-min) fairness across groups for each of three multiple access schemes, over nine experimental cases in a line-of-sight environment capturing varying levels of pathloss difference and channel correlation across the groups. Over these cases, we observe that RSMA-based MGM achieves fairness at a higher throughput for each group than SDMA- and NOMA-based MGM. These findings validate RSMA-based MGM’s promised gains from the theoretical literature.

Index Terms:
Rate-Splitting Multiple Access (RSMA), Multi-group multicast (MGM), RSMA prototy**, RSMA measurements.

I Introduction

Multi-group multicast (MGM) is a special case of multi-user wireless communications, wherein a multi-antenna transmitter (TX) jointly communicates with several groups of users simultaneously. The g𝑔gitalic_g-th group (integer g𝑔gitalic_g) has Ngsubscript𝑁𝑔N_{g}italic_N start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT users, where Ng1subscript𝑁𝑔1N_{g}\geq 1italic_N start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT ≥ 1 in general, and each user within this group desires the same message from the TX. The special case of a single group corresponds to multicast/broadcast communications, whereas Ng=1subscript𝑁𝑔1N_{g}=1italic_N start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT = 1 (for each g𝑔gitalic_g) corresponds to unicast communications.

MGM is a key physical (PHY) layer enabler of several group applications, such as video streaming, push-to-talk group radio communications, location-based services, safety-critical vehicular communications, etc [1]. The importance of these applications in future wireless networks is driving ongoing standardization efforts on how best to support multicast/broadcast service (MBS) in future wireless standards, most notably in 3GPP (Third Generation Partnership Project) [2].

Analogous to unicast, MGM can be realized through linear precoding (beamforming) at the TX, but with two key differences that need to be factored into the precoder design:

  • MGM involves the additional constraint that each user within a group must be able to decode its desired message. In general, a wide range of spatial geometries may be possible in terms of group composition, resulting in a great variety of channels experienced by users in a group. However, a special case of considerable practical interest is where users within a group are closely situated and experience similar channels, spanning a spectrum from similar channel strengths at one end, all the way to high spatial correlation/alignment at the other. This could arise in applications like gaming, video conferencing, vehicular swarms, etc. Hence, we focus on this special case in this paper.

  • Additionally, in many MGM applications such as satellite communications [3] and Internet-of-Things (IoT) [4], the number of users is also likely to exceed the number of antennas at the TX. Hence, we further focus on such overloaded scenarios for which interference management is even more challenging, since the interference power at all users cannot be suppressed to arbitrarily low levels.

The above factors have given rise to the following precoder design approaches in the literature, motivated by well-known analogues for unicast:

  • a)

    SDMA-based MGM: Most of the existing literature [5, 6, 7, 8, 9, 10, 11, 12, 13, 14] on MGM focuses on this approach, which is an extension of Space Division Multiple Access (SDMA) for unicast. Thus, each group is treated as a super-user, resulting in one precoder per group, and the precoders are designed to suppress the power of the inter-group interference so that it can be treated as noise. While there are several well-motivated objectives for precoder design in MGM, arguably the most popular in the literature is (max-min) rate fairness across the different groups111Other objectives for MGM precoder design include (i) minimizing transmit power subject to minimum rate constraints at each group [9, 10, 11], (ii) maximizing sum rate under a per antenna power constraint [12], and (ii) extending zero-forcing and MMSE precoding for MGM [14].. For SDMA-based MGM, this problem has been investigated analytically in many settings, namely [5, 6] (with a sum power constraint), [7] (with a per antenna power constraint), [8] (with low-complexity precoder design for large-scale transmit antennas) and [9] (for line-of-sight groups), [10] (for satellite communications), [13] (for cooperative integrated terrestrial-satellite communications). However, to the best of our knowledge, an experimental evaluation of SDMA-based MGM does not exist in the literature. The existing experimental studies have focused on implementing MGM at higher layers [15, 16], instead of the PHY layer.

  • b)

    NOMA-based MGM: This approach is an extension of power domain Non-Orthogonal Multiple Access (NOMA), and is guided by the principle that decoding the interference rather than treating it as noise is more effective for overloaded scenarios, particularly for unicast [17]. Like above, this approach also yields one precoder per group, but for the two-group case, one of the groups employs successive interference cancellation (SIC) to decode and subtract the interference, just like the stronger user in unicast NOMA. Precoder optimization222Note: NOMA-based MGM has also been investigated for single-antenna systems [18] and multi-antenna systems without precoder optimization [19, 20]. for NOMA-based MGM was investigated in [21, 22] (objective: minimizing transmit power subject to minimum rate constraints at each group). However, even for the overloaded scenario, NOMA-based MGM exhibited significant gains over SDMA-based MGM only when there is considerable disparity in channel strength across the groups [22]. Moreover, for groups of closely-spaced users with large angular separation between groups, it is plausible that SDMA-based MGM may still be effective in suppressing inter-group interference by exploiting the spatial domain, even in overloaded scenarios. The potential underperformance of NOMA-based MGM relative to SDMA-based MGM in such scenarios, despite a higher receiver complexity, is due to the former’s inability to exploit the multi-antenna multiplexing gain [23, 24]. For a clearer picture of the circumstances under which NOMA-based MGM is superior to SDMA-based MGM (in terms of realizing max-min rate fairness across groups, say), an experimental study of the two schemes over a variety of group spatial geometries is needed, something that is missing in the literature.

  • c)

    RSMA-based MGM: Inspired by the superiority of Rate-Splitting Multiple Access (RSMA) over both SDMA and NOMA for unicast [25, 26], this approach bridges a) and b) above to address their limitations. Specifically, each group partially decodes the inter-group interference and partially treats it as noise [27]. All groups have the same signal processing complexity for retrieving their desired messages (one stage of SIC, regardless of the number of groups). Unlike a) and b), this approach yields one more precoder than the number of groups (details in Section II). Precoder optimization for RSMA-based MGM was investigated in [24, 28, 29] (objective: max-min rate fairness across groups), where it was demonstrated analytically that RSMA-based MGM achieved a higher minimum rate than the other two approaches. This was validated using link-level simulations in [30] (under perfect CSIT) and [31, 29, 32] (for satellite communications). However, for a clearer picture of the extent of such gains, an experimental comparison of the three MGM approaches over a variety of group spatial geometries is needed.

Indeed, it is surprising that an experimental evaluation of MGM for any of the above multiple access schemes is missing from the literature. In this paper, we address this gap by conducting the first-ever experimental comparison of the fairness performance of RSMA-, SDMA- and NOMA-based MGM. Our contributions are as follows:

  • For two groups with two users per group, we formulate the practically relevant MCS-limited max-min throughput fairness problem for RSMA-based MGM. In the process, we explain how RSMA-based MGM has an advantage over SDMA- and NOMA-based MGM in terms of realizing max-min fairness through a flexible allocation of the common stream (Section II-B).

  • Using our RSMA prototype (Section III), we empirically solve the above-defined MCS-limited max-min fairness problem through measurements. In our experiments (Section IV), we realize nine cases, with the intention of capturing a wide variety in terms of pathloss difference and spatial correlation between the two groups (subject to closely situated users in each group). Over these nine cases, we observe that RSMA-based MGM achieves fairness at a higher minimum throughput than SDMA- and NOMA-based MGM. This is consistent with theoretical predictions [24, 28, 30].

I-A Notation

Column vectors are represented by lowercase bold letters (e.g., 𝐡𝐡\mathbf{h}bold_h). |||\cdot|| ⋅ | denotes the magnitude of scalars and cardinality of sets. 𝔼[]𝔼delimited-[]{\mathbb{E}}[\cdot]blackboard_E [ ⋅ ], ()H,,(\cdot)^{H},~{}\|\cdot\|,~{}\cup( ⋅ ) start_POSTSUPERSCRIPT italic_H end_POSTSUPERSCRIPT , ∥ ⋅ ∥ , ∪ and \emptyset denote the expectation operator, the Hermitian operator, the Euclidean norm, set union and the empty set, respectively. 𝒞𝒩(0,σ2)𝒞𝒩0superscript𝜎2\mathcal{CN}(0,\sigma^{2})caligraphic_C caligraphic_N ( 0 , italic_σ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) denotes the circularly symmetric complex Gaussian distribution with zero mean and variance σ2superscript𝜎2\sigma^{2}italic_σ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT.

II System Model

Without loss of generality, we consider MGM involving two groups, with each group comprising two single-antenna users (i.e., N1=N2=2subscript𝑁1subscript𝑁22N_{1}=N_{2}=2italic_N start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT = italic_N start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT = 2). Each user belongs to exactly one group, and we assume that the composition of the groups is known as apriori to the TX. For brevity, let (u,g)𝑢𝑔(u,g)( italic_u , italic_g ) denote user u𝑢uitalic_u in group g𝑔gitalic_g. All the users within a group desire the same message from the TX, which is communicated over two stages – in the first stage, the TX acquires channel state information (CSIT), which is used in the second stage to design the RSMA-based MGM precoders and transmit signal. The two stages are described in more detail next.

II-A Stage 1: CSIT Acquisition

We consider data transmission using OFDM signals over Ncsubscript𝑁𝑐N_{c}italic_N start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT subcarriers. Let Ntsubscript𝑁𝑡N_{t}italic_N start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT denote the number of antenna elements at the TX, and let 𝐡u,g[k]Ntsubscript𝐡𝑢𝑔delimited-[]𝑘superscriptsubscript𝑁𝑡{\mathbf{h}}_{u,g}[k]\in{\mathbb{C}}^{N_{t}}bold_h start_POSTSUBSCRIPT italic_u , italic_g end_POSTSUBSCRIPT [ italic_k ] ∈ blackboard_C start_POSTSUPERSCRIPT italic_N start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT end_POSTSUPERSCRIPT denote the slowly varying, flat fading channel experienced by (u,g)𝑢𝑔(u,g)( italic_u , italic_g ) over the k𝑘kitalic_k-th subcarrier (k=0,,Nc1𝑘0subscript𝑁𝑐1k=0,\dots,N_{c}-1italic_k = 0 , … , italic_N start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT - 1). Through orthogonal pilots transmitted by the TX, (u,g)𝑢𝑔(u,g)( italic_u , italic_g ) obtains an estimate of 𝐡u,g[k]subscript𝐡𝑢𝑔delimited-[]𝑘{\mathbf{h}}_{u,g}[k]bold_h start_POSTSUBSCRIPT italic_u , italic_g end_POSTSUBSCRIPT [ italic_k ], denoted by 𝐡^u,g[k]subscript^𝐡𝑢𝑔delimited-[]𝑘\hat{{\mathbf{h}}}_{u,g}[k]over^ start_ARG bold_h end_ARG start_POSTSUBSCRIPT italic_u , italic_g end_POSTSUBSCRIPT [ italic_k ]. To reduce CSI feedback overhead, (u,g)𝑢𝑔(u,g)( italic_u , italic_g ) evaluates the wideband CSI by averaging 𝐡u,g[k]subscript𝐡𝑢𝑔delimited-[]𝑘{\mathbf{h}}_{u,g}[k]bold_h start_POSTSUBSCRIPT italic_u , italic_g end_POSTSUBSCRIPT [ italic_k ] over the subcarriers, i.e.,

𝐡^u,g:=1Nck=0Nc1𝐡^u,g[k].assignsubscript^𝐡𝑢𝑔1subscript𝑁𝑐subscriptsuperscriptsubscript𝑁𝑐1𝑘0subscript^𝐡𝑢𝑔delimited-[]𝑘\hat{{\mathbf{h}}}_{u,g}:=\frac{1}{N_{c}}\sum^{N_{c}-1}_{k=0}\hat{{\mathbf{h}}% }_{u,g}[k].over^ start_ARG bold_h end_ARG start_POSTSUBSCRIPT italic_u , italic_g end_POSTSUBSCRIPT := divide start_ARG 1 end_ARG start_ARG italic_N start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT end_ARG ∑ start_POSTSUPERSCRIPT italic_N start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT - 1 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_k = 0 end_POSTSUBSCRIPT over^ start_ARG bold_h end_ARG start_POSTSUBSCRIPT italic_u , italic_g end_POSTSUBSCRIPT [ italic_k ] . (1)

II-B Stage 2: RSMA-based MGM Signal Design

Refer to caption
Figure 1: An illustration of RSMA-based MGM.

Fig. 1 depicts RSMA-based MGM in operation. Let W1subscript𝑊1W_{1}italic_W start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT and W2subscript𝑊2W_{2}italic_W start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT denote the messages meant for groups 1 and 2, respectively. At the TX, each Wgsubscript𝑊𝑔W_{g}italic_W start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT (g=1,2𝑔12g=1,2italic_g = 1 , 2) is split (by the message splitter block) into common and private portions denoted by Wc,gsubscript𝑊𝑐𝑔W_{c,g}italic_W start_POSTSUBSCRIPT italic_c , italic_g end_POSTSUBSCRIPT and Wp,gsubscript𝑊𝑝𝑔W_{p,g}italic_W start_POSTSUBSCRIPT italic_p , italic_g end_POSTSUBSCRIPT, respectively (i.e., Wg=Wc,gWp,gsubscript𝑊𝑔subscript𝑊𝑐𝑔subscript𝑊𝑝𝑔W_{g}=W_{c,g}\cup W_{p,g}italic_W start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT = italic_W start_POSTSUBSCRIPT italic_c , italic_g end_POSTSUBSCRIPT ∪ italic_W start_POSTSUBSCRIPT italic_p , italic_g end_POSTSUBSCRIPT). The two common portions – Wc,1subscript𝑊𝑐1W_{c,1}italic_W start_POSTSUBSCRIPT italic_c , 1 end_POSTSUBSCRIPT and Wc,2subscript𝑊𝑐2W_{c,2}italic_W start_POSTSUBSCRIPT italic_c , 2 end_POSTSUBSCRIPT – are then combined (by the message combiner block) into a common message, Wcsubscript𝑊𝑐W_{c}italic_W start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT. The three message components – Wcsubscript𝑊𝑐W_{c}italic_W start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT, Wp,1subscript𝑊𝑝1W_{p,1}italic_W start_POSTSUBSCRIPT italic_p , 1 end_POSTSUBSCRIPT and Wp,2subscript𝑊𝑝2W_{p,2}italic_W start_POSTSUBSCRIPT italic_p , 2 end_POSTSUBSCRIPT – are then individually encoded and modulated to form data streams sc[k]subscript𝑠𝑐delimited-[]𝑘s_{c}[k]italic_s start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT [ italic_k ], s1[k]subscript𝑠1delimited-[]𝑘s_{1}[k]italic_s start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT [ italic_k ] and s2[k]subscript𝑠2delimited-[]𝑘s_{2}[k]italic_s start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT [ italic_k ], respectively, over the subcarriers (k=0,,Nc1𝑘0subscript𝑁𝑐1k=0,\cdots,N_{c}-1italic_k = 0 , ⋯ , italic_N start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT - 1). sc[]subscript𝑠𝑐delimited-[]s_{c}[\cdot]italic_s start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT [ ⋅ ] is referred to as the common stream, whereas s1[](s2[])subscript𝑠1delimited-[]subscript𝑠2delimited-[]s_{1}[\cdot]~{}(s_{2}[\cdot])italic_s start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT [ ⋅ ] ( italic_s start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT [ ⋅ ] ) is referred to as the private stream of group 1 (group 2).

The streams are then linearly precoded, giving rise to the transmit signal 𝐱[k]𝐱delimited-[]𝑘{\mathbf{x}}[k]bold_x [ italic_k ] over the k𝑘kitalic_k-th subcarrier, which can be expressed as follows:

𝐱[k]=𝐩csc[k]+𝐩1s1[k]+𝐩2s2[k].𝐱delimited-[]𝑘subscript𝐩𝑐subscript𝑠𝑐delimited-[]𝑘subscript𝐩1subscript𝑠1delimited-[]𝑘subscript𝐩2subscript𝑠2delimited-[]𝑘\mathbf{x}[k]=\mathbf{p}_{c}s_{c}[k]+\mathbf{p}_{1}s_{1}[k]+{\mathbf{p}}_{2}s_% {2}[k].bold_x [ italic_k ] = bold_p start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT italic_s start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT [ italic_k ] + bold_p start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT italic_s start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT [ italic_k ] + bold_p start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT italic_s start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT [ italic_k ] . (2)

where 𝐩csubscript𝐩𝑐{\mathbf{p}}_{c}bold_p start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT is referred to as the common stream precoder and 𝐩gsubscript𝐩𝑔{\mathbf{p}}_{g}bold_p start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT the private stream precoder of group g𝑔gitalic_g. Let 𝐏:=[𝐩c,𝐩1,𝐩2]assign𝐏subscript𝐩𝑐subscript𝐩1subscript𝐩2{\mathbf{P}}:=[{\mathbf{p}}_{c},~{}{\mathbf{p}}_{1},~{}{\mathbf{p}}_{2}]bold_P := [ bold_p start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT , bold_p start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT , bold_p start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ] denote the collection of precoders.

Remark 1.

𝐏𝐏{\mathbf{P}}bold_P is designed as a function of the (imperfect) CSIT acquired in Stage 1 [see (8)-(9)]. However, to avoid messy notation, this functional dependence is not shown explicitly.

Remark 2.

Assuming unit symbol power (i.e., 𝔼[|sc[k]|2]=𝔼[|s1[k]|2]=𝔼[|s2[k]|2]=1𝔼delimited-[]superscriptsubscript𝑠𝑐delimited-[]𝑘2𝔼delimited-[]superscriptsubscript𝑠1delimited-[]𝑘2𝔼delimited-[]superscriptsubscript𝑠2delimited-[]𝑘21{\mathbb{E}}[|s_{c}[k]|^{2}]={\mathbb{E}}[|s_{1}[k]|^{2}]={\mathbb{E}}[|s_{2}[% k]|^{2}]=1blackboard_E [ | italic_s start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT [ italic_k ] | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ] = blackboard_E [ | italic_s start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT [ italic_k ] | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ] = blackboard_E [ | italic_s start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT [ italic_k ] | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ] = 1), 𝐩c2superscriptnormsubscript𝐩𝑐2\|{\mathbf{p}}_{c}\|^{2}∥ bold_p start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT ∥ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT and 𝐩g2superscriptnormsubscript𝐩𝑔2\|{\mathbf{p}}_{g}\|^{2}∥ bold_p start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT ∥ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT represent the power allocated to the common stream and group g𝑔gitalic_g’s private stream, respectively. In (2), the same precoders are applied across all subcarriers, which translates to uniform power allocation over the subcarriers. This suboptimal choice is driven by our use of imperfect CSIT in (1) to design 𝐏𝐏{\mathbf{P}}bold_P. The optimal power allocation – dictated by waterfilling – requires knowledge of 𝐡u,g[k]subscript𝐡𝑢𝑔delimited-[]𝑘{\mathbf{h}}_{u,g}[k]bold_h start_POSTSUBSCRIPT italic_u , italic_g end_POSTSUBSCRIPT [ italic_k ] at the TX for all k𝑘kitalic_k.

The received signal at (u,g)𝑢𝑔(u,g)( italic_u , italic_g ), denoted by yu,g[k]subscript𝑦𝑢𝑔delimited-[]𝑘y_{u,g}[k]italic_y start_POSTSUBSCRIPT italic_u , italic_g end_POSTSUBSCRIPT [ italic_k ], is given by:

yu,g[k]subscript𝑦𝑢𝑔delimited-[]𝑘\displaystyle y_{u,g}[k]italic_y start_POSTSUBSCRIPT italic_u , italic_g end_POSTSUBSCRIPT [ italic_k ] =𝐡u,gH[k]𝐱[k]+nu,g[k]absentsuperscriptsubscript𝐡𝑢𝑔𝐻delimited-[]𝑘𝐱delimited-[]𝑘subscript𝑛𝑢𝑔delimited-[]𝑘\displaystyle=\mathbf{h}_{u,g}^{H}[k]\mathbf{x}[k]+n_{u,g}[k]= bold_h start_POSTSUBSCRIPT italic_u , italic_g end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_H end_POSTSUPERSCRIPT [ italic_k ] bold_x [ italic_k ] + italic_n start_POSTSUBSCRIPT italic_u , italic_g end_POSTSUBSCRIPT [ italic_k ]
=𝐡u,gH[k]𝐩csc[k]+𝐡u,gH𝐩gsg[k]+𝐡u,gH𝐩gsg[k]absentsuperscriptsubscript𝐡𝑢𝑔𝐻delimited-[]𝑘subscript𝐩𝑐subscript𝑠𝑐delimited-[]𝑘superscriptsubscript𝐡𝑢𝑔𝐻subscript𝐩𝑔subscript𝑠𝑔delimited-[]𝑘superscriptsubscript𝐡𝑢𝑔𝐻subscript𝐩superscript𝑔subscript𝑠superscript𝑔delimited-[]𝑘\displaystyle=\mathbf{h}_{u,g}^{H}[k]\mathbf{p}_{c}s_{c}[k]+\mathbf{h}_{u,g}^{% H}\mathbf{p}_{g}s_{g}[k]+\mathbf{h}_{u,g}^{H}\mathbf{p}_{g^{\prime}}s_{g^{% \prime}}[k]= bold_h start_POSTSUBSCRIPT italic_u , italic_g end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_H end_POSTSUPERSCRIPT [ italic_k ] bold_p start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT italic_s start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT [ italic_k ] + bold_h start_POSTSUBSCRIPT italic_u , italic_g end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_H end_POSTSUPERSCRIPT bold_p start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT italic_s start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT [ italic_k ] + bold_h start_POSTSUBSCRIPT italic_u , italic_g end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_H end_POSTSUPERSCRIPT bold_p start_POSTSUBSCRIPT italic_g start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_POSTSUBSCRIPT italic_s start_POSTSUBSCRIPT italic_g start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_POSTSUBSCRIPT [ italic_k ]
+nu,g[k].subscript𝑛𝑢𝑔delimited-[]𝑘\displaystyle~{}+n_{u,g}[k].+ italic_n start_POSTSUBSCRIPT italic_u , italic_g end_POSTSUBSCRIPT [ italic_k ] . (3)

From the perspective of (u,g)𝑢𝑔(u,g)( italic_u , italic_g ), the first and second terms in (II-B) both contain useful information – the former represents the common stream component, a part of which is meant solely for group g𝑔gitalic_g users, while the latter represents the private stream component meant solely for group g𝑔gitalic_g users. In contrast, the third term does not contain useful information, as it captures the private stream component meant for the other group ggsuperscript𝑔𝑔g^{\prime}\neq gitalic_g start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ≠ italic_g. Finally, nu,g[k]𝒞𝒩(0,σ2)subscript𝑛𝑢𝑔delimited-[]𝑘𝒞𝒩0superscript𝜎2n_{u,g}[k]\thicksim\mathcal{CN}(0,\sigma^{2})italic_n start_POSTSUBSCRIPT italic_u , italic_g end_POSTSUBSCRIPT [ italic_k ] ∼ caligraphic_C caligraphic_N ( 0 , italic_σ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) denotes the receiver noise at (u,g)𝑢𝑔(u,g)( italic_u , italic_g ).

We assume that precoded pilot333These pilots play a role similar to the demodulation reference signals (DM-RS) used in LTE and 5G NR. signals have been interspersed across some subcarriers of 𝐱[k]𝐱delimited-[]𝑘{\mathbf{x}}[k]bold_x [ italic_k ] to help (u,g)𝑢𝑔(u,g)( italic_u , italic_g ) estimate 𝐡u,gH[k]𝐩csuperscriptsubscript𝐡𝑢𝑔𝐻delimited-[]𝑘subscript𝐩𝑐{\mathbf{h}}_{u,g}^{H}[k]{\mathbf{p}}_{c}bold_h start_POSTSUBSCRIPT italic_u , italic_g end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_H end_POSTSUPERSCRIPT [ italic_k ] bold_p start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT and 𝐡u,gH[k]𝐩gsuperscriptsubscript𝐡𝑢𝑔𝐻delimited-[]𝑘subscript𝐩𝑔{\mathbf{h}}_{u,g}^{H}[k]{\mathbf{p}}_{g}bold_h start_POSTSUBSCRIPT italic_u , italic_g end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_H end_POSTSUPERSCRIPT [ italic_k ] bold_p start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT upon receiving yu,g[k]subscript𝑦𝑢𝑔delimited-[]𝑘y_{u,g}[k]italic_y start_POSTSUBSCRIPT italic_u , italic_g end_POSTSUBSCRIPT [ italic_k ], in order to decode the useful information in the first two terms of (II-B). Using its estimate of 𝐡u,gH[k]𝐩csuperscriptsubscript𝐡𝑢𝑔𝐻delimited-[]𝑘subscript𝐩𝑐{\mathbf{h}}_{u,g}^{H}[k]{\mathbf{p}}_{c}bold_h start_POSTSUBSCRIPT italic_u , italic_g end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_H end_POSTSUPERSCRIPT [ italic_k ] bold_p start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT, (u,g)𝑢𝑔(u,g)( italic_u , italic_g ) first equalizes (II-B) to decode sc[k]subscript𝑠𝑐delimited-[]𝑘s_{c}[k]italic_s start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT [ italic_k ] and recover Wcsubscript𝑊𝑐W_{c}italic_W start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT, by treating the second and third (interference) terms in (II-B) as noise. The decoded estimate of Wcsubscript𝑊𝑐W_{c}italic_W start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT is simultaneously sent to:

  • (a)

    the message splitter block to extract W^c,g(u)superscriptsubscript^𝑊𝑐𝑔𝑢\hat{W}_{c,g}^{(u)}over^ start_ARG italic_W end_ARG start_POSTSUBSCRIPT italic_c , italic_g end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( italic_u ) end_POSTSUPERSCRIPT, which is the estimate of Wc,gsubscript𝑊𝑐𝑔W_{c,g}italic_W start_POSTSUBSCRIPT italic_c , italic_g end_POSTSUBSCRIPT at (u,g)𝑢𝑔(u,g)( italic_u , italic_g ), and

  • (b)

    the SIC block, which generates an estimate of 𝐡u,gH[k]𝐩csc[k]superscriptsubscript𝐡𝑢𝑔𝐻delimited-[]𝑘subscript𝐩𝑐subscript𝑠𝑐delimited-[]𝑘{\mathbf{h}}_{u,g}^{H}[k]{\mathbf{p}}_{c}s_{c}[k]bold_h start_POSTSUBSCRIPT italic_u , italic_g end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_H end_POSTSUPERSCRIPT [ italic_k ] bold_p start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT italic_s start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT [ italic_k ] and subtracts it from yu,g[k]subscript𝑦𝑢𝑔delimited-[]𝑘y_{u,g}[k]italic_y start_POSTSUBSCRIPT italic_u , italic_g end_POSTSUBSCRIPT [ italic_k ]. Using its estimate of 𝐡u,gH[k]𝐩gsuperscriptsubscript𝐡𝑢𝑔𝐻delimited-[]𝑘subscript𝐩𝑔{\mathbf{h}}_{u,g}^{H}[k]{\mathbf{p}}_{g}bold_h start_POSTSUBSCRIPT italic_u , italic_g end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_H end_POSTSUPERSCRIPT [ italic_k ] bold_p start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT, (u,g)𝑢𝑔(u,g)( italic_u , italic_g ) then equalizes the resulting residue to decode sg[]subscript𝑠𝑔delimited-[]s_{g}[\cdot]italic_s start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT [ ⋅ ] and recover Wp,gsubscript𝑊𝑝𝑔W_{p,g}italic_W start_POSTSUBSCRIPT italic_p , italic_g end_POSTSUBSCRIPT, while treating the interference from the third term of (II-B) as noise. Let W^p,g(u)superscriptsubscript^𝑊𝑝𝑔𝑢\hat{W}_{p,g}^{(u)}over^ start_ARG italic_W end_ARG start_POSTSUBSCRIPT italic_p , italic_g end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( italic_u ) end_POSTSUPERSCRIPT denote the decoded estimate of Wp,gsubscript𝑊𝑝𝑔W_{p,g}italic_W start_POSTSUBSCRIPT italic_p , italic_g end_POSTSUBSCRIPT at u𝑢uitalic_u. Hence, the estimate of Wgsubscript𝑊𝑔W_{g}italic_W start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT at (u,g)𝑢𝑔(u,g)( italic_u , italic_g ), denoted by W^g(u)superscriptsubscript^𝑊𝑔𝑢\hat{W}_{g}^{(u)}over^ start_ARG italic_W end_ARG start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( italic_u ) end_POSTSUPERSCRIPT, is given by W^g(u):=W^c,g(u)W^p,g(u)assignsuperscriptsubscript^𝑊𝑔𝑢superscriptsubscript^𝑊𝑐𝑔𝑢superscriptsubscript^𝑊𝑝𝑔𝑢\hat{W}_{g}^{(u)}:=\hat{W}_{c,g}^{(u)}\cup\hat{W}_{p,g}^{(u)}over^ start_ARG italic_W end_ARG start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( italic_u ) end_POSTSUPERSCRIPT := over^ start_ARG italic_W end_ARG start_POSTSUBSCRIPT italic_c , italic_g end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( italic_u ) end_POSTSUPERSCRIPT ∪ over^ start_ARG italic_W end_ARG start_POSTSUBSCRIPT italic_p , italic_g end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( italic_u ) end_POSTSUPERSCRIPT.

Let Rc,g(𝐏)subscript𝑅𝑐𝑔𝐏R_{c,g}(\mathbf{P})italic_R start_POSTSUBSCRIPT italic_c , italic_g end_POSTSUBSCRIPT ( bold_P ) denote the highest (instantaneous) rate (in bits/s/Hz) supporting error-free decoding of sc[]subscript𝑠𝑐delimited-[]s_{c}[\cdot]italic_s start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT [ ⋅ ] among the users in group g𝑔gitalic_g, under the assumption that 𝐡^u,gsubscript^𝐡𝑢𝑔\hat{{\mathbf{h}}}_{u,g}over^ start_ARG bold_h end_ARG start_POSTSUBSCRIPT italic_u , italic_g end_POSTSUBSCRIPT in (1) constitutes perfect CSIT. It has the following expression:

Rc,g(𝐏):=minulog2(1+|𝐡^u,gH𝐩c|2σ2+g=12|𝐡^u,gH𝐩g|2),assignsubscript𝑅𝑐𝑔𝐏subscript𝑢subscript21superscriptsuperscriptsubscript^𝐡𝑢𝑔𝐻subscript𝐩𝑐2superscript𝜎2superscriptsubscript𝑔12superscriptsuperscriptsubscript^𝐡𝑢𝑔𝐻subscript𝐩𝑔2\displaystyle R_{c,g}(\mathbf{P}):=\min_{u}\log_{2}\left(1+\frac{|\hat{\mathbf% {h}}_{u,g}^{H}\mathbf{p}_{c}|^{2}}{\sigma^{2}+\sum_{g=1}^{2}|\hat{\mathbf{h}}_% {u,g}^{H}\mathbf{p}_{g}|^{2}}\right),italic_R start_POSTSUBSCRIPT italic_c , italic_g end_POSTSUBSCRIPT ( bold_P ) := roman_min start_POSTSUBSCRIPT italic_u end_POSTSUBSCRIPT roman_log start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ( 1 + divide start_ARG | over^ start_ARG bold_h end_ARG start_POSTSUBSCRIPT italic_u , italic_g end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_H end_POSTSUPERSCRIPT bold_p start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG italic_σ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + ∑ start_POSTSUBSCRIPT italic_g = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT | over^ start_ARG bold_h end_ARG start_POSTSUBSCRIPT italic_u , italic_g end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_H end_POSTSUPERSCRIPT bold_p start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG ) , (4)

where the minimum reflects the fact that every user in group g𝑔gitalic_g must decode sc[]subscript𝑠𝑐delimited-[]s_{c}[\cdot]italic_s start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT [ ⋅ ].

Remark 3 (Achievability).

Strictly speaking, Rc,g(𝐏)subscript𝑅𝑐𝑔𝐏R_{c,g}({\mathbf{P}})italic_R start_POSTSUBSCRIPT italic_c , italic_g end_POSTSUBSCRIPT ( bold_P ) is not achievable as it may involve transmitting at a rate that is not supported by the true CSI {𝐡u,g[k]:k}conditional-setsubscript𝐡𝑢𝑔delimited-[]𝑘for-all𝑘\{{\mathbf{h}}_{u,g}[k]:\forall k\}{ bold_h start_POSTSUBSCRIPT italic_u , italic_g end_POSTSUBSCRIPT [ italic_k ] : ∀ italic_k }. The ergodic rate, 𝔼𝐡,𝐡^[Rc,g(𝐏)]subscript𝔼𝐡^𝐡delimited-[]subscript𝑅𝑐𝑔𝐏{\mathbb{E}}_{{\mathbf{h}},\hat{{\mathbf{h}}}}[R_{c,g}({\mathbf{P}})]blackboard_E start_POSTSUBSCRIPT bold_h , over^ start_ARG bold_h end_ARG end_POSTSUBSCRIPT [ italic_R start_POSTSUBSCRIPT italic_c , italic_g end_POSTSUBSCRIPT ( bold_P ) ], on the other hand is achievable, where the averaging is over the joint distribution of the true CSI and the CSIT [33] (subscripts omitted for simplicity). Since this joint distribution is unknown in practice, the ergodic rate is not a tractable metric for designing 𝐏𝐏{\mathbf{P}}bold_P. Hence, despite not being technically achievable, Rc,g(𝐏)subscript𝑅𝑐𝑔𝐏R_{c,g}({\mathbf{P}})italic_R start_POSTSUBSCRIPT italic_c , italic_g end_POSTSUBSCRIPT ( bold_P ) [and similarly, Rp,g(𝐏)subscript𝑅𝑝𝑔𝐏R_{p,g}({\mathbf{P}})italic_R start_POSTSUBSCRIPT italic_p , italic_g end_POSTSUBSCRIPT ( bold_P ) in (6)] is still useful444The ergodic rate, 𝔼𝐡,𝐡^[Rc,g(𝐏)]subscript𝔼𝐡^𝐡delimited-[]subscript𝑅𝑐𝑔𝐏{\mathbb{E}}_{{\mathbf{h}},\hat{{\mathbf{h}}}}[R_{c,g}({\mathbf{P}})]blackboard_E start_POSTSUBSCRIPT bold_h , over^ start_ARG bold_h end_ARG end_POSTSUBSCRIPT [ italic_R start_POSTSUBSCRIPT italic_c , italic_g end_POSTSUBSCRIPT ( bold_P ) ], can be expressed as 𝔼𝐡^[𝔼𝐡^|𝐡[Rc,g(𝐏)]]subscript𝔼^𝐡delimited-[]subscript𝔼conditional^𝐡𝐡delimited-[]subscript𝑅𝑐𝑔𝐏{\mathbb{E}}_{\hat{{\mathbf{h}}}}[{\mathbb{E}}_{\hat{{\mathbf{h}}}|{\mathbf{h}% }}[R_{c,g}({\mathbf{P}})]]blackboard_E start_POSTSUBSCRIPT over^ start_ARG bold_h end_ARG end_POSTSUBSCRIPT [ blackboard_E start_POSTSUBSCRIPT over^ start_ARG bold_h end_ARG | bold_h end_POSTSUBSCRIPT [ italic_R start_POSTSUBSCRIPT italic_c , italic_g end_POSTSUBSCRIPT ( bold_P ) ] ], where the outer expectation is w.r.t the distribution of the CSIT, and the inner expectation is w.r.t the conditional distribution of the true CSI, given the CSIT. As an alternative to the instantaneous rate Rc,g(𝐏)subscript𝑅𝑐𝑔𝐏R_{c,g}({\mathbf{P}})italic_R start_POSTSUBSCRIPT italic_c , italic_g end_POSTSUBSCRIPT ( bold_P ), the term 𝔼𝐡^|𝐡[Rc,g(𝐏)]subscript𝔼conditional^𝐡𝐡delimited-[]subscript𝑅𝑐𝑔𝐏{\mathbb{E}}_{\hat{{\mathbf{h}}}|{\mathbf{h}}}[R_{c,g}({\mathbf{P}})]blackboard_E start_POSTSUBSCRIPT over^ start_ARG bold_h end_ARG | bold_h end_POSTSUBSCRIPT [ italic_R start_POSTSUBSCRIPT italic_c , italic_g end_POSTSUBSCRIPT ( bold_P ) ] – known as the average rate (and in general, not achievable) – can also be used to design 𝐏𝐏{\mathbf{P}}bold_P as a function of the CSIT [33]. However, like with the ergodic rate, the average rate is also not a tractable metric in practice, since the conditional distribution is unknown and obtaining an empirical estimate at the TX involves considerable overhead. for designing 𝐏𝐏{\mathbf{P}}bold_P [see OPmmf𝑂subscript𝑃mmfOP_{\rm mmf}italic_O italic_P start_POSTSUBSCRIPT roman_mmf end_POSTSUBSCRIPT in (8)-(9)]. For such 𝐏𝐏{\mathbf{P}}bold_P, the empirically achievable rates in our experiments are determined by a suitable choice of modulation and coding scheme (MCS) levels.

Since sc[]subscript𝑠𝑐delimited-[]s_{c}[\cdot]italic_s start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT [ ⋅ ] contains useful information for users in both groups, it follows that both groups must be able to decode sc[]subscript𝑠𝑐delimited-[]s_{c}[\cdot]italic_s start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT [ ⋅ ]. Hence, subject to (4), the highest rate supporting error-free decoding of sc[]subscript𝑠𝑐delimited-[]s_{c}[\cdot]italic_s start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT [ ⋅ ] by both groups – denoted by Rc(𝐏)subscript𝑅𝑐𝐏R_{c}(\mathbf{P})italic_R start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT ( bold_P ) and referred to as the common stream rate – is given by:

Rc(𝐏)=mingRc,g(𝐏).subscript𝑅𝑐𝐏subscript𝑔subscript𝑅𝑐𝑔𝐏R_{c}(\mathbf{P})=\min_{g}R_{c,g}(\mathbf{P}).italic_R start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT ( bold_P ) = roman_min start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT italic_R start_POSTSUBSCRIPT italic_c , italic_g end_POSTSUBSCRIPT ( bold_P ) . (5)

After perfect SIC (in other words, assuming Rc(𝐏)subscript𝑅𝑐𝐏R_{c}({\mathbf{P}})italic_R start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT ( bold_P ) is achievable for all users), let Rp,g(𝐏)subscript𝑅𝑝𝑔𝐏R_{p,g}(\mathbf{P})italic_R start_POSTSUBSCRIPT italic_p , italic_g end_POSTSUBSCRIPT ( bold_P ) denote the highest (instantaneous) rate supporting error-free decoding of sg[]subscript𝑠𝑔delimited-[]s_{g}[\cdot]italic_s start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT [ ⋅ ] among the users in group g𝑔gitalic_g, again under the assumption that (1) constitutes perfect CSIT. This is referred to as group g𝑔gitalic_g’s private stream rate and has an expression similar to (4), given by:

Rp,g(𝐏):=minulog2(1+|𝐡^u,gH𝐩g|2σ2+|𝐡^u,gH𝐩g|2)(gg).assignsubscript𝑅𝑝𝑔𝐏subscript𝑢subscript21superscriptsuperscriptsubscript^𝐡𝑢𝑔𝐻subscript𝐩𝑔2superscript𝜎2superscriptsuperscriptsubscript^𝐡𝑢𝑔𝐻subscript𝐩superscript𝑔2superscript𝑔𝑔R_{p,g}(\mathbf{P}):=\min_{u}\log_{2}\left(1+\frac{|\hat{\mathbf{h}}_{u,g}^{H}% \mathbf{p}_{g}|^{2}}{\sigma^{2}+|\hat{\mathbf{h}}_{u,g}^{H}\mathbf{p}_{g^{% \prime}}|^{2}}\right)~{}(g^{\prime}\neq g).italic_R start_POSTSUBSCRIPT italic_p , italic_g end_POSTSUBSCRIPT ( bold_P ) := roman_min start_POSTSUBSCRIPT italic_u end_POSTSUBSCRIPT roman_log start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ( 1 + divide start_ARG | over^ start_ARG bold_h end_ARG start_POSTSUBSCRIPT italic_u , italic_g end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_H end_POSTSUPERSCRIPT bold_p start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG italic_σ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + | over^ start_ARG bold_h end_ARG start_POSTSUBSCRIPT italic_u , italic_g end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_H end_POSTSUPERSCRIPT bold_p start_POSTSUBSCRIPT italic_g start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_POSTSUBSCRIPT | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG ) ( italic_g start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ≠ italic_g ) . (6)

Subject to (5) and (6), the highest rate supporting error-free recovery of Wgsubscript𝑊𝑔W_{g}italic_W start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT by all users in group g𝑔gitalic_g – denoted by Rg(𝐏)subscript𝑅𝑔𝐏R_{g}({\mathbf{P}})italic_R start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT ( bold_P ) and referred to as the net group g𝑔gitalic_g rate – is given by :

Rg(𝐏)=|Wc,g||Wc|Rc(𝐏)+Rp,g(𝐏),subscript𝑅𝑔𝐏subscript𝑊𝑐𝑔subscript𝑊𝑐subscript𝑅𝑐𝐏subscript𝑅𝑝𝑔𝐏R_{g}(\mathbf{P})=\frac{|W_{c,g}|}{|W_{c}|}R_{c}(\mathbf{P})+R_{p,g}(\mathbf{P% }),italic_R start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT ( bold_P ) = divide start_ARG | italic_W start_POSTSUBSCRIPT italic_c , italic_g end_POSTSUBSCRIPT | end_ARG start_ARG | italic_W start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT | end_ARG italic_R start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT ( bold_P ) + italic_R start_POSTSUBSCRIPT italic_p , italic_g end_POSTSUBSCRIPT ( bold_P ) , (7)

where the first term is the fraction of the common message (Wcsubscript𝑊𝑐W_{c}italic_W start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT) intended for group g𝑔gitalic_g.

In general, the sizes of Wc,1subscript𝑊𝑐1W_{c,1}italic_W start_POSTSUBSCRIPT italic_c , 1 end_POSTSUBSCRIPT and Wc,2subscript𝑊𝑐2W_{c,2}italic_W start_POSTSUBSCRIPT italic_c , 2 end_POSTSUBSCRIPT (in bits) in Fig. 1 need not be equal, which can be used to enable RSMA-based MGM to achieve better fairness. To illustrate this, suppose group 2’s private stream rate is smaller than that of group 1 [i.e., Rp,2(𝐏)<Rp,1(𝐏)subscript𝑅𝑝2𝐏subscript𝑅𝑝1𝐏R_{p,2}({\mathbf{P}})<R_{p,1}({\mathbf{P}})italic_R start_POSTSUBSCRIPT italic_p , 2 end_POSTSUBSCRIPT ( bold_P ) < italic_R start_POSTSUBSCRIPT italic_p , 1 end_POSTSUBSCRIPT ( bold_P )] because group 2 users have weaker channels than group 1 users. In such a scenario, allocating a larger fraction of the common stream to group 2 (|Wc,2|>|Wc,1|subscript𝑊𝑐2subscript𝑊𝑐1|W_{c,2}|>|W_{c,1}|| italic_W start_POSTSUBSCRIPT italic_c , 2 end_POSTSUBSCRIPT | > | italic_W start_POSTSUBSCRIPT italic_c , 1 end_POSTSUBSCRIPT |) has the effect of driving the net group rates – R1(𝐏)subscript𝑅1𝐏R_{1}({\mathbf{P}})italic_R start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ( bold_P ) and R2(𝐏)subscript𝑅2𝐏R_{2}({\mathbf{P}})italic_R start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ( bold_P ) – towards parity, as captured by the RHS of (7). This motivates the following scheme:

  • S1)

    if |Rp,1(𝐏)Rp,2(𝐏)|Rc(𝐏)subscript𝑅𝑝1𝐏subscript𝑅𝑝2𝐏subscript𝑅𝑐𝐏|R_{p,1}({\mathbf{P}})-R_{p,2}({\mathbf{P}})|\leq R_{c}({\mathbf{P}})| italic_R start_POSTSUBSCRIPT italic_p , 1 end_POSTSUBSCRIPT ( bold_P ) - italic_R start_POSTSUBSCRIPT italic_p , 2 end_POSTSUBSCRIPT ( bold_P ) | ≤ italic_R start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT ( bold_P ), then the common stream is allocated such that Rp,1(𝐏)+(|Wc,1|/|Wc|)Rc(𝐏)=Rp,2(𝐏)+(|Wc,2|/|Wc|)Rc(𝐏)subscript𝑅𝑝1𝐏subscript𝑊𝑐1subscript𝑊𝑐subscript𝑅𝑐𝐏subscript𝑅𝑝2𝐏subscript𝑊𝑐2subscript𝑊𝑐subscript𝑅𝑐𝐏R_{p,1}({\mathbf{P}})+(|W_{c,1}|/|W_{c}|)R_{c}({\mathbf{P}})=R_{p,2}({\mathbf{% P}})+(|W_{c,2}|/|W_{c}|)R_{c}({\mathbf{P}})italic_R start_POSTSUBSCRIPT italic_p , 1 end_POSTSUBSCRIPT ( bold_P ) + ( | italic_W start_POSTSUBSCRIPT italic_c , 1 end_POSTSUBSCRIPT | / | italic_W start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT | ) italic_R start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT ( bold_P ) = italic_R start_POSTSUBSCRIPT italic_p , 2 end_POSTSUBSCRIPT ( bold_P ) + ( | italic_W start_POSTSUBSCRIPT italic_c , 2 end_POSTSUBSCRIPT | / | italic_W start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT | ) italic_R start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT ( bold_P ) to ensure that both groups have the same throughput (i.e., achieving max-min fairness);

  • S2)

    otherwise, when |Rp,1(𝐏)Rp,2(𝐏)|>Rc(𝐏)subscript𝑅𝑝1𝐏subscript𝑅𝑝2𝐏subscript𝑅𝑐𝐏|R_{p,1}({\mathbf{P}})-R_{p,2}({\mathbf{P}})|>R_{c}({\mathbf{P}})| italic_R start_POSTSUBSCRIPT italic_p , 1 end_POSTSUBSCRIPT ( bold_P ) - italic_R start_POSTSUBSCRIPT italic_p , 2 end_POSTSUBSCRIPT ( bold_P ) | > italic_R start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT ( bold_P ), the entire common stream is allocated to the group with the smaller private stream rate.

The design of 𝐏𝐏{\mathbf{P}}bold_P to realize max-min rate fairness across the two groups yields the following optimization problem:

OPmmf:max𝐏:𝑂subscript𝑃mmfsubscript𝐏\displaystyle OP_{\rm mmf}:\max_{\mathbf{P}}italic_O italic_P start_POSTSUBSCRIPT roman_mmf end_POSTSUBSCRIPT : roman_max start_POSTSUBSCRIPT bold_P end_POSTSUBSCRIPT min{R1(𝐏),R2(𝐏)}subscript𝑅1𝐏subscript𝑅2𝐏\displaystyle~{}\min\{R_{1}({\mathbf{P}}),R_{2}({\mathbf{P}})\}roman_min { italic_R start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ( bold_P ) , italic_R start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ( bold_P ) } (8)
s.t. tr(𝐏𝐏H)Pt,trsuperscript𝐏𝐏𝐻subscript𝑃𝑡\displaystyle~{}{\mathrm{tr}}(\mathbf{P}\mathbf{P}^{H})\leq P_{t},roman_tr ( bold_PP start_POSTSUPERSCRIPT italic_H end_POSTSUPERSCRIPT ) ≤ italic_P start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT , (9)

where (9) is the TX power constraint. Due to the non-convexity of OPmmf𝑂subscript𝑃mmfOP_{\rm mmf}italic_O italic_P start_POSTSUBSCRIPT roman_mmf end_POSTSUBSCRIPT, a tractable algorithm that converges to its global optimal solution does not exist. However, a locally optimal solution that is widely recognized as the gold-standard can be obtained using the WMMSE method [24], which takes advantage of the well-known relationship between the achievable rate and the post-equalization symbol mean square error [34]. Let 𝐏wm:=[𝐩cwm,𝐩1wm,𝐩2wm]assignsuperscript𝐏wmsuperscriptsubscript𝐩𝑐wmsuperscriptsubscript𝐩1wmsuperscriptsubscript𝐩2wm\mathbf{P}^{\mathrm{wm}}:=[{\mathbf{p}}_{c}^{\rm wm},~{}{\mathbf{p}}_{1}^{\rm wm% },~{}{\mathbf{p}}_{2}^{\rm wm}]bold_P start_POSTSUPERSCRIPT roman_wm end_POSTSUPERSCRIPT := [ bold_p start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT start_POSTSUPERSCRIPT roman_wm end_POSTSUPERSCRIPT , bold_p start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT roman_wm end_POSTSUPERSCRIPT , bold_p start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT roman_wm end_POSTSUPERSCRIPT ] denote the solution to OPmmf𝑂subscript𝑃mmfOP_{\rm mmf}italic_O italic_P start_POSTSUBSCRIPT roman_mmf end_POSTSUBSCRIPT obtained by the WMMSE method. However, the resulting max-min fair rate is difficult to realize in practice because, in addition to Remark 3:

  • 1.

    Data transmission is typically restricted to a finite collection of MCS levels, which caps the achievable rate/throughput, and

  • 2.

    Error-free decoding of the streams at any user cannot be guaranteed due to a combination of finite block length effects and CSI estimation errors, which further reduces the measured throughput.

To account for imperfect CSIT, discrete MCS levels and decoding errors, we consider the MCS-limited throughput as the metric of interest throughout this paper. An MCS level is defined by a pair (m,r)𝑚𝑟(m,r)( italic_m , italic_r ), where positive integer m𝑚mitalic_m denotes the bits per constellation symbol (e.g., 2 for QPSK) and r(0,1]𝑟01r\in(0,1]italic_r ∈ ( 0 , 1 ] denotes the code rate. Let :={(mc,rc),(m1,r1),(m2,r2)}assignsubscript𝑚𝑐subscript𝑟𝑐subscript𝑚1subscript𝑟1subscript𝑚2subscript𝑟2{\mathcal{M}}:=\{(m_{c},r_{c}),(m_{1},r_{1}),(m_{2},r_{2})\}caligraphic_M := { ( italic_m start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT , italic_r start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT ) , ( italic_m start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT , italic_r start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ) , ( italic_m start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT , italic_r start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ) } denote the set of MCS levels chosen for sc[]subscript𝑠𝑐delimited-[]s_{c}[\cdot]italic_s start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT [ ⋅ ], s1[]subscript𝑠1delimited-[]s_{1}[\cdot]italic_s start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT [ ⋅ ] and s2[]subscript𝑠2delimited-[]s_{2}[\cdot]italic_s start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT [ ⋅ ], respectively. Then, the corresponding MCS-limited throughputs (measured in bits/s) with WMMSE precoders are given by:

Tcmcs(𝐏wm,)superscriptsubscript𝑇𝑐mcssuperscript𝐏wm\displaystyle T_{c}^{\rm mcs}({\mathbf{P}}^{\rm wm},{\mathcal{M}})italic_T start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT start_POSTSUPERSCRIPT roman_mcs end_POSTSUPERSCRIPT ( bold_P start_POSTSUPERSCRIPT roman_wm end_POSTSUPERSCRIPT , caligraphic_M ) =Bmcrc×(W^c=Wc),absent𝐵subscript𝑚𝑐subscript𝑟𝑐subscript^𝑊𝑐subscript𝑊𝑐\displaystyle=Bm_{c}r_{c}\times{\mathbb{P}}(\hat{W}_{c}=W_{c}),= italic_B italic_m start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT italic_r start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT × blackboard_P ( over^ start_ARG italic_W end_ARG start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT = italic_W start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT ) , (10)
Tp,1mcs(𝐏wm,)superscriptsubscript𝑇𝑝1mcssuperscript𝐏wm\displaystyle T_{p,1}^{\rm mcs}({\mathbf{P}}^{\rm wm},{\mathcal{M}})italic_T start_POSTSUBSCRIPT italic_p , 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT roman_mcs end_POSTSUPERSCRIPT ( bold_P start_POSTSUPERSCRIPT roman_wm end_POSTSUPERSCRIPT , caligraphic_M ) =Bm1r1×(W^p,1(1)=W^p,1(2)=Wp,1),absent𝐵subscript𝑚1subscript𝑟1superscriptsubscript^𝑊𝑝11superscriptsubscript^𝑊𝑝12subscript𝑊𝑝1\displaystyle=Bm_{1}r_{1}\times{\mathbb{P}}(\hat{W}_{p,1}^{(1)}=\hat{W}_{p,1}^% {(2)}=W_{p,1}),= italic_B italic_m start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT italic_r start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT × blackboard_P ( over^ start_ARG italic_W end_ARG start_POSTSUBSCRIPT italic_p , 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( 1 ) end_POSTSUPERSCRIPT = over^ start_ARG italic_W end_ARG start_POSTSUBSCRIPT italic_p , 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( 2 ) end_POSTSUPERSCRIPT = italic_W start_POSTSUBSCRIPT italic_p , 1 end_POSTSUBSCRIPT ) ,
Tp,2mcs(𝐏wm,)superscriptsubscript𝑇𝑝2mcssuperscript𝐏wm\displaystyle T_{p,2}^{\rm mcs}({\mathbf{P}}^{\rm wm},{\mathcal{M}})italic_T start_POSTSUBSCRIPT italic_p , 2 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT roman_mcs end_POSTSUPERSCRIPT ( bold_P start_POSTSUPERSCRIPT roman_wm end_POSTSUPERSCRIPT , caligraphic_M ) =Bm2r2×(W^p,2(1)=W^p,2(2)=Wp,2).absent𝐵subscript𝑚2subscript𝑟2superscriptsubscript^𝑊𝑝21superscriptsubscript^𝑊𝑝22subscript𝑊𝑝2\displaystyle=Bm_{2}r_{2}\times{\mathbb{P}}(\hat{W}_{p,2}^{(1)}=\hat{W}_{p,2}^% {(2)}=W_{p,2}).= italic_B italic_m start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT italic_r start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT × blackboard_P ( over^ start_ARG italic_W end_ARG start_POSTSUBSCRIPT italic_p , 2 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( 1 ) end_POSTSUPERSCRIPT = over^ start_ARG italic_W end_ARG start_POSTSUBSCRIPT italic_p , 2 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( 2 ) end_POSTSUPERSCRIPT = italic_W start_POSTSUBSCRIPT italic_p , 2 end_POSTSUBSCRIPT ) .

In the above expressions, B𝐵Bitalic_B denotes the effective bandwidth555The effective bandwidth is the portion of the total bandwidth available for data transferring, after accounting for signalling overhead (e.g., the cyclic prefix in OFDM, pilot subcarriers and guard band in IEEE 802.11 standard), and the probability terms capture the loss in throughput due to decoding errors – specifically, the first expression denotes the probability that Wcsubscript𝑊𝑐W_{c}italic_W start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT is correctly decoded by all groups, while the second and third expressions denote the probability that Wp,1subscript𝑊𝑝1W_{p,1}italic_W start_POSTSUBSCRIPT italic_p , 1 end_POSTSUBSCRIPT and Wp,2subscript𝑊𝑝2W_{p,2}italic_W start_POSTSUBSCRIPT italic_p , 2 end_POSTSUBSCRIPT are correctly decoded by groups 1 and 2, respectively.

Remark 4.

The message decoding probabilities in the RHS of (10) are a function of the chosen MCS level, as well as the SINR at each user, which in turn, depends on 𝐏wmsuperscript𝐏wm{\mathbf{P}}^{\rm wm}bold_P start_POSTSUPERSCRIPT roman_wm end_POSTSUPERSCRIPT. This functional dependence is not shown explicitly to avoid messy notation.

Similar to (7), the MCS-limited throughput for group g𝑔gitalic_g with WMMSE precoders is given by :

Tgmcs(𝐏wm,)=|Wc,g||Wc|Tcmcs(𝐏wm,)+Tp,gmcs(𝐏wm,)superscriptsubscript𝑇𝑔mcssuperscript𝐏wmsubscript𝑊𝑐𝑔subscript𝑊𝑐superscriptsubscript𝑇𝑐mcssuperscript𝐏wmsuperscriptsubscript𝑇𝑝𝑔mcssuperscript𝐏wmT_{g}^{\rm mcs}({\mathbf{P}}^{\rm wm},\mathcal{M})=\frac{|W_{c,g}|}{|W_{c}|}T_% {c}^{\rm mcs}({\mathbf{P}}^{\rm wm},{\mathcal{M}})+T_{p,g}^{\rm mcs}({\mathbf{% P}}^{\rm wm},{\mathcal{M}})italic_T start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT start_POSTSUPERSCRIPT roman_mcs end_POSTSUPERSCRIPT ( bold_P start_POSTSUPERSCRIPT roman_wm end_POSTSUPERSCRIPT , caligraphic_M ) = divide start_ARG | italic_W start_POSTSUBSCRIPT italic_c , italic_g end_POSTSUBSCRIPT | end_ARG start_ARG | italic_W start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT | end_ARG italic_T start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT start_POSTSUPERSCRIPT roman_mcs end_POSTSUPERSCRIPT ( bold_P start_POSTSUPERSCRIPT roman_wm end_POSTSUPERSCRIPT , caligraphic_M ) + italic_T start_POSTSUBSCRIPT italic_p , italic_g end_POSTSUBSCRIPT start_POSTSUPERSCRIPT roman_mcs end_POSTSUPERSCRIPT ( bold_P start_POSTSUPERSCRIPT roman_wm end_POSTSUPERSCRIPT , caligraphic_M ) (11)

Thus, similar to OPmmf𝑂subscript𝑃mmfOP_{\rm mmf}italic_O italic_P start_POSTSUBSCRIPT roman_mmf end_POSTSUBSCRIPT in (8)-(9), the MCS-limited max-min throughput fairness problem for RSMA-based MGM, which is of considerable practical relevance, can be defined as follows:

OPmmfmcs:max:𝑂superscriptsubscript𝑃mmfmcssubscript\displaystyle OP_{\rm mmf}^{\rm mcs}:\max_{{\mathcal{M}}}italic_O italic_P start_POSTSUBSCRIPT roman_mmf end_POSTSUBSCRIPT start_POSTSUPERSCRIPT roman_mcs end_POSTSUPERSCRIPT : roman_max start_POSTSUBSCRIPT caligraphic_M end_POSTSUBSCRIPT min{T1mcs(𝐏wm,),T2mcs(𝐏wm,)}subscriptsuperscript𝑇mcs1superscript𝐏wmsubscriptsuperscript𝑇mcs2superscript𝐏wm\displaystyle~{}\min\{T^{\rm mcs}_{1}({\mathbf{P}}^{\rm wm},{\mathcal{M}}),T^{% \rm mcs}_{2}({\mathbf{P}}^{\rm wm},{\mathcal{M}})\}roman_min { italic_T start_POSTSUPERSCRIPT roman_mcs end_POSTSUPERSCRIPT start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ( bold_P start_POSTSUPERSCRIPT roman_wm end_POSTSUPERSCRIPT , caligraphic_M ) , italic_T start_POSTSUPERSCRIPT roman_mcs end_POSTSUPERSCRIPT start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ( bold_P start_POSTSUPERSCRIPT roman_wm end_POSTSUPERSCRIPT , caligraphic_M ) } (12)
s.t. 𝕄,𝕄\displaystyle~{}{\mathcal{M}}\in{\mathbb{M}},caligraphic_M ∈ blackboard_M , (13)

where 𝕄𝕄{\mathbb{M}}blackboard_M denotes the collection of permissible MCS levels for the three streams, which is typically pre-determined through standards (see Table II for the 𝕄𝕄{\mathbb{M}}blackboard_M used in our measurements). The optimal {\mathcal{M}}caligraphic_M for OPmmfmcs𝑂superscriptsubscript𝑃mmfmcsOP_{\rm mmf}^{\rm mcs}italic_O italic_P start_POSTSUBSCRIPT roman_mmf end_POSTSUBSCRIPT start_POSTSUPERSCRIPT roman_mcs end_POSTSUPERSCRIPT is a function of 𝐏wmsuperscript𝐏wm{\mathbf{P}}^{\rm wm}bold_P start_POSTSUPERSCRIPT roman_wm end_POSTSUPERSCRIPT that is difficult to characterize because the message decoding probabilities in (10) do not have closed-form expressions in terms of 𝐏wmsuperscript𝐏wm{\mathbf{P}}^{\rm wm}bold_P start_POSTSUPERSCRIPT roman_wm end_POSTSUPERSCRIPT and {\mathcal{M}}caligraphic_M. However, these probabilities can be empirically evaluated through measurements. This motivates us to empirically solve OPmmfmcs𝑂superscriptsubscript𝑃mmfmcsOP_{\rm mmf}^{\rm mcs}italic_O italic_P start_POSTSUBSCRIPT roman_mmf end_POSTSUBSCRIPT start_POSTSUPERSCRIPT roman_mcs end_POSTSUPERSCRIPT through a brute force search666A more sophisticated empirical approach involves link adaptation, where the most suitable MCS level is determined by an ARQ-based mechanism [35]. This is left for future work. over 𝕄𝕄{\mathbb{M}}blackboard_M, which is the focus of our measurements in Section IV.

We conclude this subsection by noting SDMA- and NOMA-based MGMs are special cases of RSMA-based MGM, corresponding to specific choices of message splitting, as explained in the following remarks.

Remark 5.

SDMA-based MGM is a special case of RSMA-based MGM, where the common stream is turned off – i.e., Wg=Wp,g,(g=1,2)subscript𝑊𝑔subscript𝑊𝑝𝑔𝑔12W_{g}=W_{p,g},(g=1,2)italic_W start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT = italic_W start_POSTSUBSCRIPT italic_p , italic_g end_POSTSUBSCRIPT , ( italic_g = 1 , 2 ) in Fig. 1. The corresponding expressions in (2)-(13) for SDMA-based MGM can be obtained by setting 𝐩c=𝟎subscript𝐩𝑐0{\mathbf{p}}_{c}=\mathbf{0}bold_p start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT = bold_0. Essentially, without a common stream, group g𝑔gitalic_g users treat the interference from Wg(gg)subscript𝑊superscript𝑔superscript𝑔𝑔W_{g^{\prime}}~{}(g^{\prime}\neq g)italic_W start_POSTSUBSCRIPT italic_g start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_POSTSUBSCRIPT ( italic_g start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ≠ italic_g ) as noise.

Remark 6.

NOMA-based MGM is a special case of RSMA-based MGM, where the private stream message of one group (group 2222, say) is turned off (i.e., Wp,2=subscript𝑊𝑝2W_{p,2}=\emptysetitalic_W start_POSTSUBSCRIPT italic_p , 2 end_POSTSUBSCRIPT = ∅), and no portion of the other message is allocated to the common stream (i.e., W1=Wp,1,W2=Wc,2=Wcformulae-sequencesubscript𝑊1subscript𝑊𝑝1subscript𝑊2subscript𝑊𝑐2subscript𝑊𝑐W_{1}=W_{p,1},W_{2}=W_{c,2}=W_{c}italic_W start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT = italic_W start_POSTSUBSCRIPT italic_p , 1 end_POSTSUBSCRIPT , italic_W start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT = italic_W start_POSTSUBSCRIPT italic_c , 2 end_POSTSUBSCRIPT = italic_W start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT in Fig. 1). The corresponding expressions in (2)-(13) can be obtained by setting 𝐩2=𝟎subscript𝐩20{\mathbf{p}}_{2}=\mathbf{0}bold_p start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT = bold_0. For this example, group 1 users fully decode and subtract the interference from W2subscript𝑊2W_{2}italic_W start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT, whereas group 2 users treat the interference from W1subscript𝑊1W_{1}italic_W start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT as noise. In contrast, for RSMA-based MGM, group g𝑔gitalic_g users partially decode the interference from Wgg(Wc,g)subscript𝑊superscript𝑔𝑔subscript𝑊𝑐superscript𝑔W_{g^{\prime}\neq g}~{}(W_{c,g^{\prime}})italic_W start_POSTSUBSCRIPT italic_g start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ≠ italic_g end_POSTSUBSCRIPT ( italic_W start_POSTSUBSCRIPT italic_c , italic_g start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_POSTSUBSCRIPT ) and partially treat it as noise (Wp,gsubscript𝑊𝑝superscript𝑔W_{p,g^{\prime}}italic_W start_POSTSUBSCRIPT italic_p , italic_g start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_POSTSUBSCRIPT).

Remark 7.

The common stream allocation, as per condition S2 on page 4, is similar to NOMA-based MGM, albeit the weaker group also has a private stream.

III RSMA Prototype

III-A Hardware setup

We implement RSMA-based MGM by using our Software Defined Radio (SDR) based RSMA prototype [26]. The TX and RXs are realized using National Instruments’ (NI) USRP 2942 SDR units, which have two antennas/RF chains. Hence, we use three USRP 2942 units, one to realize a two-antenna TX (Nt=2subscript𝑁𝑡2N_{t}=2italic_N start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT = 2), and the other two to realize four single-antenna users that are assigned to two groups of two users each in our measurements in Section IV. In particular, the antennas of group 2 users are connected to their USRP through coaxial cables, which allow them to be moved around to realize different channel environments, as described in Section IV-A. The USRPs share a common timing source (CDA-2990), and are controlled by a workstation running LabVIEW NXG, through which the various blocks in Fig. 1 are realized. All connections (SDRs to workstation, SDRs to timing sources) are through PCIe cables, facilitated by a PCIe bus (CPS-8910). A list of hardware components is provided in Table I.

Name Description
1. Workstation Running LabVIEW NXG
2. NI USRP-2942 (3 units) SDRs used to realize TX and users
3. NI CPS-8910 Provides additional PCIe ports
4. NI CDA-2990 8 Channel, 10 MHz clock distribution device
5. 8dBi RP-SMA Male Wifi Antenna TX antennas
6. RP-SMA Male Wifi Antenna Group 1 users’ antennas
7. Mini-Circuits ZAPD-2-272-S+ Power splitter
8. TP-Link TL-ANT2405C Group 2 users’ antennas
TABLE I: List of hardware components.

III-B RSMA-based MGM Implementation

Refer to caption
Figure 2: Signal structure within the two-stage transmission protocol used to implement RSMA-based MGM. In conventional 802.11g, every DATA (OFDM) symbol for a user contains four pilot subcarriers for FPS. However, for RSMA-based MGM, this approach would cause the pilot subcarriers in the common and private streams to interfere with each other. To avoid this, the three streams carry non-zero pilots orthogonally once every three OFDM symbols, as shown above.

We adopt several features of the IEEE 802.11g physical layer frames to implement the system model described in Section II.

Stage 1

Each TX antenna transmits a pilot signal orthogonally in time comprising a Short Training Field (STF, 8μs8𝜇s8\mu{\rm s}8 italic_μ roman_s in duration) and a Long Training Field (LTF, 8μs8𝜇s8\mu{\rm s}8 italic_μ roman_s in duration), as shown in the top portion of Fig. 2. The STF is used for synchronization and coarse frequency offset estimation, while the LTF is used for CSI estimation at each user (i.e., 𝐡^u,g[k]subscript^𝐡𝑢𝑔delimited-[]𝑘\hat{{\mathbf{h}}}_{u,g}[k]over^ start_ARG bold_h end_ARG start_POSTSUBSCRIPT italic_u , italic_g end_POSTSUBSCRIPT [ italic_k ] in Section II-A).

Stage 2

The transmitted signal consists of a preamble followed by the data payload, as shown in the bottom portion of Fig. 2.

  • Preamble: The preamble consists of one STF and three LTFs. The function of the STF is the same as in Stage 1, while the LTFs are precoded in order to estimate the precoded CSI for equalization at the users. The first LTF is used by each (u,g)𝑢𝑔(u,g)( italic_u , italic_g ) to estimate 𝐡u,gH𝐩csuperscriptsubscript𝐡𝑢𝑔𝐻subscript𝐩𝑐{\mathbf{h}}_{u,g}^{H}{\mathbf{p}}_{c}bold_h start_POSTSUBSCRIPT italic_u , italic_g end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_H end_POSTSUPERSCRIPT bold_p start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT for decoding the common stream. The second LTF is used by each user u𝑢uitalic_u in group 1 to estimate 𝐡u,1H𝐩1superscriptsubscript𝐡𝑢1𝐻subscript𝐩1{\mathbf{h}}_{u,1}^{H}{\mathbf{p}}_{1}bold_h start_POSTSUBSCRIPT italic_u , 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_H end_POSTSUPERSCRIPT bold_p start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT to decode its private stream. Similarly, the third LTF is used by each user usuperscript𝑢u^{\prime}italic_u start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT in group 2 to estimate 𝐡u,2H𝐩2superscriptsubscript𝐡superscript𝑢2𝐻subscript𝐩2{\mathbf{h}}_{u^{\prime},2}^{H}{\mathbf{p}}_{2}bold_h start_POSTSUBSCRIPT italic_u start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT , 2 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_H end_POSTSUPERSCRIPT bold_p start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT to decode its private stream.

  • Data Payload: For the payload, we consider a total bandwidth of 20MHz20MHz20{\rm MHz}20 roman_M roman_H roman_z with Nc=64subscript𝑁𝑐64N_{c}=64italic_N start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT = 64 subcarriers and a cyclic prefix (CP) of 16 samples per OFDM symbol. Aligned with IEEE 802.11 frames, 52525252 subcarriers are used for communications while the rest serve as guard bands. Among these 52525252 subcarriers, 48484848 are used to carry data symbols, with the remaining used to correct the common phase error across all subcarriers in one OFDM symbol [36]. This yields an effective bandwidth of:

    B=20MHz×(64/80)CP overhead×(48/64)Guard bandoverhead=12MHz𝐵20MHzsubscript6480CP overheadsubscript4864Guard bandoverhead12MHz\displaystyle B=20{\rm MHz}\times\underbrace{(64/80)}_{\text{CP overhead}}% \times\underbrace{(48/64)}_{\begin{subarray}{c}\text{Guard band}\\ \text{overhead}\end{subarray}}=12{\rm MHz}italic_B = 20 roman_M roman_H roman_z × under⏟ start_ARG ( 64 / 80 ) end_ARG start_POSTSUBSCRIPT CP overhead end_POSTSUBSCRIPT × under⏟ start_ARG ( 48 / 64 ) end_ARG start_POSTSUBSCRIPT start_ARG start_ROW start_CELL Guard band end_CELL end_ROW start_ROW start_CELL overhead end_CELL end_ROW end_ARG end_POSTSUBSCRIPT = 12 roman_M roman_H roman_z (14)

    The payload consists of three superposed streams (one common, two private), each comprising 50 OFDM symbols.

  • MCS Implementation: Table II lists the MCS levels, 𝕄𝕄{\mathbb{M}}blackboard_M, implemented in our prototype. For channel coding, we implement Polar codes augmented with an 8-bit cyclic redundancy check [37, 38], along with successive cancellation list decoding [39], with a list depth of 2. After the preamble, the first OFDM symbol (labelled SERVICE in the bottom portion of Fig. 2) contains the MCS information of each stream.

MCS Modulation Code Rate Data Rate
Index (m𝑚mitalic_m) r𝑟ritalic_r Bmr𝐵𝑚𝑟Bmritalic_B italic_m italic_r (Mbps)
00 BPSK (1) 1/2121/21 / 2 6666
1111 BPSK (1) 3/4343/43 / 4 9999
2222 QPSK (2) 1/2121/21 / 2 12121212
3333 QPSK (2) 3/4343/43 / 4 18181818
4444 16QAM (4) 1/2121/21 / 2 24242424
5555 16QAM (4) 3/4343/43 / 4 36363636
6666 64QAM (6) 2/3232/32 / 3 48484848
7777 64QAM (6) 3/4343/43 / 4 54545454
8888 256QAM (8) 3/4343/43 / 4 72727272
9999 256QAM (8) 5/6565/65 / 6 80808080
TABLE II: MCS levels (largely based on IEEE 802.11g) implemented in our prototype. The data rate in the last column is equal to Bmr𝐵𝑚𝑟Bmritalic_B italic_m italic_r, where B𝐵Bitalic_B is the effective bandwidth given by (14).

An instance of Stage 1 and Stage 2, as described above and illustrated in Fig. 2, constitutes a single measurement run. To empirically solve OPmmfmcs𝑂superscriptsubscript𝑃mmfmcsOP_{\rm mmf}^{\rm mcs}italic_O italic_P start_POSTSUBSCRIPT roman_mmf end_POSTSUBSCRIPT start_POSTSUPERSCRIPT roman_mcs end_POSTSUPERSCRIPT, we conduct 100100100100 measurement runs. Let Dcsubscript𝐷𝑐D_{c}italic_D start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT denote the number of runs in which the common stream is successfully decoded by all users. Similarly, let Dg(g{1,2})subscript𝐷𝑔𝑔12D_{g}~{}(g\in\{1,2\})italic_D start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT ( italic_g ∈ { 1 , 2 } ) denote the number of runs in which both users in group g𝑔gitalic_g successfully decode their private stream. Replacing the message decoding probabilities in (10) with their empirical estimates, the measured common and private stream throughputs for RSMA-based MGM is given by:

Tcmcs(𝐏wm,)subscriptsuperscript𝑇mcs𝑐superscript𝐏wm\displaystyle T^{\rm mcs}_{c}({\mathbf{P}}^{\rm wm},{\mathcal{M}})italic_T start_POSTSUPERSCRIPT roman_mcs end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT ( bold_P start_POSTSUPERSCRIPT roman_wm end_POSTSUPERSCRIPT , caligraphic_M ) =Dc100Bmcrc,absentsubscript𝐷𝑐100𝐵subscript𝑚𝑐subscript𝑟𝑐\displaystyle=\frac{D_{c}}{100}Bm_{c}r_{c},= divide start_ARG italic_D start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT end_ARG start_ARG 100 end_ARG italic_B italic_m start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT italic_r start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT ,
Tp,1mcs(𝐏wm,)subscriptsuperscript𝑇mcs𝑝1superscript𝐏wm\displaystyle T^{\rm mcs}_{p,1}({\mathbf{P}}^{\rm wm},{\mathcal{M}})italic_T start_POSTSUPERSCRIPT roman_mcs end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_p , 1 end_POSTSUBSCRIPT ( bold_P start_POSTSUPERSCRIPT roman_wm end_POSTSUPERSCRIPT , caligraphic_M ) =D1100Bm1r1,absentsubscript𝐷1100𝐵subscript𝑚1subscript𝑟1\displaystyle=\frac{D_{1}}{100}Bm_{1}r_{1},= divide start_ARG italic_D start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT end_ARG start_ARG 100 end_ARG italic_B italic_m start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT italic_r start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ,
Tp,2mcs(𝐏wm,)subscriptsuperscript𝑇mcs𝑝2superscript𝐏wm\displaystyle T^{\rm mcs}_{p,2}({\mathbf{P}}^{\rm wm},{\mathcal{M}})italic_T start_POSTSUPERSCRIPT roman_mcs end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_p , 2 end_POSTSUBSCRIPT ( bold_P start_POSTSUPERSCRIPT roman_wm end_POSTSUPERSCRIPT , caligraphic_M ) =D2100Bm2r2,absentsubscript𝐷2100𝐵subscript𝑚2subscript𝑟2\displaystyle=\frac{D_{2}}{100}Bm_{2}r_{2},= divide start_ARG italic_D start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_ARG start_ARG 100 end_ARG italic_B italic_m start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT italic_r start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT , (15)

where B𝐵Bitalic_B is given by (14) and the MCS levels are chosen from Table II.

Parameter Notation Value
Center frequency fcsubscript𝑓𝑐f_{c}italic_f start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT 2.484GHz2.484GHz2.484{\rm GHz}2.484 roman_GHz777This value corresponds to channel no. 14141414 in the IEEE 802.11 family of standards for the 2.4GHz2.4GHz2.4{\rm GHz}2.4 roman_GHz band. We use this channel to avoid ambient WiFi interference, as it is not commercially used in the UK.
Transmit power Ptsubscript𝑃𝑡P_{t}italic_P start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT 23dBm23dBm23{\rm dBm}23 roman_d roman_B roman_m
No. of groups 2
No. of users per group 2
Total bandwidth 20MHz20MHz20{\rm MHz}20 roman_M roman_H roman_z
Subcarriers Total (Ncsubscript𝑁𝑐N_{c}italic_N start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT) 64646464
Data 48484848
Pilot (FPS) 4444
Guard band 12121212
CP length 16161616
Effective bandwidth B𝐵Bitalic_B 12MHz12MHz12{\rm MHz}12 roman_M roman_H roman_z
OFDM symbols in payload 50505050
Experiment runs (per case) 100100100100
TX antenna spacing 0.13m0.13m0.13{\rm m}0.13 roman_m
Fraunhofer distance 0.28m0.28m0.28{\rm m}0.28 roman_m
Distance between Group-1 RXs and TX 1.00m
Distance between Group-2 RXs and TX Case 1-3 1.00-1.50m
Case 4-6 2.00-2.30m
Case 7-9 3.20-3.50m
TABLE III: List of parameters used in our experiments.

IV Results

IV-A Measurement setup

In MGM, it is well known that a group’s throughput is limited by its weakest user. Hence, in the existing literature, there is an assumption of admission control during the formation of the groups, so that users within a group have similar channel strength/link SNR. This motivates us to focus on the scenario where users within a group are closely situated, which could occur in applications like gaming, video conferencing, etc. In such cases, users within a group are likely to experience similar pathloss, whereas there could be considerable pathloss differences across groups. Furthermore, as the relative performance of SDMA-, NOMA- and RSMA-based MGM is a function of the inter-group interference, it is important to realize a range of environments with varying levels of inter-group interference for a meaningful three-way comparison. This is intrinsically challenging for MGM, as notions of weak/medium/strong interference need to be defined over four pairs of links across the two groups instead of a single pair of links for two-user unicast. Motivated by our unicast measurements [26, Section IV], we rely on geometric notions of channel spatial correlation in line-of-sight environments as a predictor of inter-group interference levels. Thus, similar to [26], we consider nine measurement cases, as illustrated in Fig. 3(a), where moving in the horizontal direction should increase the inter-group pathloss difference, while moving in the vertical direction is should decrease the spatial correlation (and in turn, the interference levels) between the groups.

To quantify the inter-group pathloss difference between user u𝑢uitalic_u in group 1 and user usuperscript𝑢u^{\prime}italic_u start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT in group 2, we define a parameter, αu,usubscript𝛼𝑢superscript𝑢\alpha_{u,u^{\prime}}italic_α start_POSTSUBSCRIPT italic_u , italic_u start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_POSTSUBSCRIPT (in dB scale) as follows:

αu,u[dB]=10log10𝐡^u,2𝐡^u,1,subscript𝛼𝑢superscript𝑢delimited-[]dB10subscript10normsubscript^𝐡superscript𝑢2normsubscript^𝐡𝑢1\alpha_{u,u^{\prime}}~{}[\mathrm{dB}]=10\log_{10}\frac{||\hat{\mathbf{h}}_{u^{% \prime},2}||}{||\hat{\mathbf{h}}_{u,1}||},italic_α start_POSTSUBSCRIPT italic_u , italic_u start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_POSTSUBSCRIPT [ roman_dB ] = 10 roman_log start_POSTSUBSCRIPT 10 end_POSTSUBSCRIPT divide start_ARG | | over^ start_ARG bold_h end_ARG start_POSTSUBSCRIPT italic_u start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT , 2 end_POSTSUBSCRIPT | | end_ARG start_ARG | | over^ start_ARG bold_h end_ARG start_POSTSUBSCRIPT italic_u , 1 end_POSTSUBSCRIPT | | end_ARG , (16)

where 𝐡^u,gsubscript^𝐡𝑢𝑔\hat{\mathbf{h}}_{u,g}over^ start_ARG bold_h end_ARG start_POSTSUBSCRIPT italic_u , italic_g end_POSTSUBSCRIPT is given by (1). Since group 2 is farther from the TX than group 1, a large negative value signifies greater pathloss difference among the two groups. Likewise, let ρu,usubscript𝜌𝑢superscript𝑢\rho_{u,u^{\prime}}italic_ρ start_POSTSUBSCRIPT italic_u , italic_u start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_POSTSUBSCRIPT denote the spatial correlation between users u𝑢uitalic_u in group 1 and usuperscript𝑢u^{\prime}italic_u start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT in group 2, which is defined as follows:

ρu,u=|𝐡^u,1H𝐡^u,2|𝐡^u,1𝐡^u,2.subscript𝜌𝑢superscript𝑢superscriptsubscript^𝐡𝑢1𝐻subscript^𝐡superscript𝑢2normsubscript^𝐡𝑢1normsubscript^𝐡superscript𝑢2\rho_{u,u^{\prime}}=\frac{|\hat{\mathbf{h}}_{u,1}^{H}\hat{\mathbf{h}}_{u^{% \prime},2}|}{||\hat{\mathbf{h}}_{u,1}||\cdot||\hat{\mathbf{h}}_{u^{\prime},2}|% |}.italic_ρ start_POSTSUBSCRIPT italic_u , italic_u start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_POSTSUBSCRIPT = divide start_ARG | over^ start_ARG bold_h end_ARG start_POSTSUBSCRIPT italic_u , 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_H end_POSTSUPERSCRIPT over^ start_ARG bold_h end_ARG start_POSTSUBSCRIPT italic_u start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT , 2 end_POSTSUBSCRIPT | end_ARG start_ARG | | over^ start_ARG bold_h end_ARG start_POSTSUBSCRIPT italic_u , 1 end_POSTSUBSCRIPT | | ⋅ | | over^ start_ARG bold_h end_ARG start_POSTSUBSCRIPT italic_u start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT , 2 end_POSTSUBSCRIPT | | end_ARG . (17)

Thus, ρu,u=1subscript𝜌𝑢superscript𝑢1\rho_{u,u^{\prime}}=1italic_ρ start_POSTSUBSCRIPT italic_u , italic_u start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_POSTSUBSCRIPT = 1 signifies fully correlated (i.e., highly interfering) channels, while ρu,u=0subscript𝜌𝑢superscript𝑢0\rho_{u,u^{\prime}}=0italic_ρ start_POSTSUBSCRIPT italic_u , italic_u start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_POSTSUBSCRIPT = 0 signifies orthogonal (i.e., zero interference) channels. The intuitive directions of decreasing α𝛼\alphaitalic_α and ρ𝜌\rhoitalic_ρ are marked in Fig. 3(a).

To realize Fig. 3(a), group 1 users are fixed on a workbench 1m from the TX, while group 2 users are placed on a trolley (Fig. 3(b)) and moved around to realize the nine cases. The distance between the TX and group 2 is provided in Table III. All users are in the far field. Table IV captures the spread of the four α𝛼\alphaitalic_α and ρ𝜌\rhoitalic_ρ values for the nine cases. While the α𝛼\alphaitalic_α spread is fairly narrow and follows the expected trend shown in Fig. 3a, the ρ𝜌\rhoitalic_ρ spread is much larger for several cases because the intuition behind strong/medium/high levels of inter-group spatial correlation in Fig. 3(a) was based on only the (dominant) line-of-sight component, whereas the measured ρ𝜌\rhoitalic_ρ is also sensitive to multipath, which is uncorrelated for each user. This highlights the difficulty of precisely controlling the extent of inter-group interference. Nevertheless, a shift towards 1 in the ρ𝜌\rhoitalic_ρ spread consistent with increasing spatial correlation can be discerned when looking at the low and medium category of cases. No such insight can be drawn from the medium and high categories.

A full list of parameters used in our experiments is provided in Table III.

Inter-group Pathloss Difference Inter-group Spatial Correlation
Case α𝛼\alphaitalic_α spread [-20,0] Case ρ𝜌\rhoitalic_ρ spread [0,1]
Low 1
3
2
6
3
9
Medium 4
2
5
5
6
8
High 7
1
8
4
9
7
TABLE IV: The rectangles in the third and last columns capture the spread of the measured inter-group pathloss difference (αu,usubscript𝛼𝑢superscript𝑢\alpha_{u,u^{\prime}}italic_α start_POSTSUBSCRIPT italic_u , italic_u start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_POSTSUBSCRIPT) and inter-group spatial correlation (ρu,usubscript𝜌𝑢superscript𝑢\rho_{u,u^{\prime}}italic_ρ start_POSTSUBSCRIPT italic_u , italic_u start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_POSTSUBSCRIPT). The left (right) end of each rectangle corresponds to the smallest (largest) value for each case. As expected, the α𝛼\alphaitalic_α spread becomes more negative as the pathloss difference between the groups increases. But, the ρ𝜌\rhoitalic_ρ spread is large for several cases due to sensitivity to multipath. Despite this, a shift towards 1 consistent with increasing spatial correlation can be discerned when looking at the low and medium categories of cases.
Refer to caption
(a) The layout of the measurement environment, along with the TX and user positions for the nine cases.
Refer to caption
(b) The measurement environment (Case 4 is shown here).
Figure 3: Measurement setup

IV-B Fairness Comparison

Refer to caption
Figure 4: Fairness comparison between SDMA-based, NOMA-based and RSMA-based MGM. The number beside each data point indicates the measurement case. The black line (y=2x𝑦2𝑥y=2xitalic_y = 2 italic_x) corresponds to max-min fairness, and points that are closer (in terms of Euclidean distance) to this line represent fairer outcomes.

To compare the fairness performance of RSMA-, NOMA- and SDMA-based MGM, Fig. 4 plots the sum throughput versus the minimum throughput for each case. The case number is indicated beside each data point. The black y=2x𝑦2𝑥y=2xitalic_y = 2 italic_x line represents max-min fairness, and the region above is feasible for all three multiple access schemes. Thus, points that are closer (in terms of Euclidean distance) to the black line represent fairer outcomes. Furthermore, a point that is either to the north, northeast or east of another point represents a better outcome. Based on these insights, we make the following observations:

  • SDMA: From the spatial geometry of the user locations, cases 1, 4 and 7 have the highest amount of inter-group interference. For these cases, SDMA-based MGM achieves significantly lower minimum and sum throughput than NOMA- and RSMA-based MGM despite achieving near fairness. This is borne by the fact that the SDMA points for these cases lie to the southwest of the corresponding NOMA/RSMA points roughly along the y=2x𝑦2𝑥y=2xitalic_y = 2 italic_x line. This validates the well-known limitation of treating interference as noise in the high interference regime.

  • NOMA: With the exception of Case 9, NOMA-based MGM either achieves or comes close to achieving fairness for the other cases. It also achieves better outcomes than SDMA for cases 1, 4, 5, 7 and 8. Interestingly, it achieves strictly worse outcomes than SDMA for cases 3, 6, and 9, despite SDMA’s limitations on interference suppression in an overloaded scenario. It’s worth noting that these cases are associated with the lowest inter-group interference (due to the relatively lower inter-group spatial correlation). This suggests that even for overloaded scenarios where interference cannot be fully suppressed to noise levels, a flexible mix of (partially) decoding the interference and (partially) treating it as noise, in response to the channel conditions is the ideal interference strategy, something RSMA achieves by design.

  • RSMA: In Section II, when condition S1 (see page 4) is true, the corresponding RSMA-based MGM point lies on the y=2x𝑦2𝑥y=2xitalic_y = 2 italic_x (i.e., achieves max-min fairness). Otherwise, for condition S2, the RSMA point lies as close as possible above the black line. For all cases, RSMA achieves max-min fairness, thereby implying that S1 is satisfied for each case (more on this while discussing Fig. 5). Furthermore, compared to SDMA, RSMA achieves strictly better performance for all cases. Likewise, when compared to NOMA, RSMA achieves strictly better performance for all cases bar one – the exception being case 7, where NOMA achieves a higher sum throughput than RSMA but not fairness (see Section IV-C-c for an explanation). To explain RSMA’s superior performance w.r.t SDMA and NOMA, it’s important to gain some insight into the relative contributions of the common and private streams to each group’s net throughput for each case. We will elaborate on this next.

Refer to caption
(a)
Refer to caption
(b)
Refer to caption
(c)
Figure 5: The throughput performance for both multicast groups of SDMA-, NOMA- and RSMA-based MGM.

Complementing Fig. 4, Fig. 5 plots the throughput of each group. For RSMA-based MGM, a group’s throughput is broken down into common and private stream contributions. The arrows indicate the gain in the minimum throughput due to RSMA-based MGM. We make the following observations:

  • In each subfigure, the contribution of the common stream to each group’s throughput decreases from left to right across the three cases. This pattern was observed for unicast communications, as well [26]. Essentially, when the inter-group interference is high (as in cases 1, 4 and 7), most of the power is allocated to the common stream, which, in turn, contributes substantially to the net throughput. In other words, decoding most of the interference is favoured over suppressing it to noise levels. On the other hand, when the inter-group interference is lower (as in cases 3, 6 and 9), suppressing most of the interference to noise levels is preferred and hence, the power allocation to the private streams increases. As a result, the private streams contribute more to the net throughput.

  • For RSMA-based MGM, there is a non-zero common stream contribution to each group’s throughput in every case. Thus, S1 (see page 4) is satisfied for each case, and the allocation of the common stream to each group is done to ensure that both groups have the same net throughput. In other words, throughput contributions from both the common and the private streams for each group provides the required flexibility in terms of resource allocation to ensure that fairness is realized for each case.

  • To further highlight the benefits of throughput contributions from both the common and the private streams for each group, consider cases 1 and 7 for RSMA-based MGM. Since the private stream contribution to (the weaker) group 2’s throughput is zero, it appears at first glance as though RSMA-based MGM reduces to NOMA-based MGM. But there is a key difference. Unlike NOMA, the common stream can contribute to (the stronger) group 1’s throughput as well, to the extent that most of group 1’s throughput is due to the common stream. This is intuitive, since under high inter-group interference, most of the power is allocated to the common stream. In contrast, for NOMA-based MGM, group 1 needs to rely entirely on its private stream for its throughput. This forces a suboptimal resource allocation, where power has to be diverted from the common stream to group 1’s private stream to achieve fairness. As a result, NOMA-based MGM achieves a lower minimum throughput than RSMA-based MGM for these cases. The flexibility offered by throughput contributions from both the common and the private streams for each group is also why RSMA-based MGM achieves better outcomes than both SDMA- and NOMA-based MGM.

In summary, RSMA-based MGM adapts well to variations in pathloss difference and interference between the groups and achieves superior fairness performance (i.e., fairness at a higher minimum throughput) when compared to SDMA- and NOMA-based MGM. We conclude this section by remarking on some seemingly counter-intuitive results in Fig. 5.

IV-C Counter-intuitive Results

BLER (MCS index from Table II)
Case SDMA NOMA RSMA
Group 1 Group 2 Group 1 Group 2 Common Group 1 pvt. Group 2 pvt.
1 0.02 (0) 0 (0) 0.15 (4) 0.01 (4) 0.01 (6) 0.08 (0)
2 0.01 (4) 0.01 (4) 0.03 (4) 0 (4) 0.10 (5) 0.25 (3) 0.10 (2)
3 0.05 (5) 0.09 (4) 0.01 (4) 0.02 (4) 0.06 (2) 0.04 (4) 0.06 (4)
4 0 (3) 0.09 (3) 0.02 (4) 0 (4) 0 (4) 0.08 (4) 0.05 (0)
5 0 (4) 0.03 (3) 0.11 (4) 0.02 (4) 0 (4) 0.11 (2) 0.02 (3)
6 0 (4) 0.12 (5) 0.01 (4) 0.01 (3) 0 (3) 0 (4) 0.09 (4)
7 0.23 (2) 0 (0) 0 (4) 0 (3) 0 (5) 0.45 (0)
8 0 (3) 0 (3) 0.08 (4) 0 (3) 0 (5) 0.04 (2) 0.06 (0)
9 0 (5) 0.23 (3) 0 (5) 0.15 (1) 0 (2) 0.01 (4) 0.01 (3)
TABLE V: The MCS Index and the BLER corresponding to each throughput bar in Fig. 5.

SDMA, Case 6

Perhaps the most counter-intuitive observation in Fig. 5 is the supposedly weaker group 2 having a higher throughput (31.68Mbps) than the stronger group 1 (24Mbps) for SDMA-based MGM in Case 6. Table V lists the MCS levels and the block error rate (BLER) associated with each bar in Fig. 5, and the values corresponding to this particular case are:

  • Group 1: 16QAM, rate 1/2 (level 4 in Table II) with 0 BLER

  • Group 2: 16QAM, rate 3/4 (level 5 in Table II) with 0.12 BLER

It is clear from above that group 2’s higher throughput stems from its ability to support a higher MCS level than group 1, albeit it is only one level higher and not error-free. The higher MCS level is due to the fact that more power is allocated to the group 2 precoder to achieve fairness. Ideally, under perfect CSIT, one would expect more power to be allocated to the group 2 precoder in such a way that both groups support the same MCS level, thereby achieving the same throughput. However, with imperfect CSIT, which is inherent in our experiments due to the use of wideband CSI in (1), we reckon it is not unusual for group 2 to support an MCS level one higher than group 1. Moreover, if we had imposed strict reliability requirements (i.e., zero BLER in our measurements), then both groups would have supported the same MCS level (16QAM, rate 1/2) and achieved the same throughput.

NOMA, Cases 1, 2, 4 and 5

Again, group 2 has a higher throughput than group 1 for these cases, although the difference is not as stark as above. This is because both groups support the same MCS level (see Table V) which is expected, but the throughput variations are due to block errors that are difficult to control.

NOMA v/s RSMA, Case 7

For NOMA-based MGM, groups 1 and 2 supports MCS levels 4 and 3, respectively (both with zero BLER, see Table V), and the resulting sum throughput is 42Mbps. Due to the high pathloss difference and the high inter-group interference associated with this case, RSMA-based MGM reduces to NOMA-based MGM, as evidenced by the lack of a private stream for group 2 (see Remark 6). But, unlike NOMA, the RSMA common stream can contribute to the throughput of both groups. Hence, nearly all of the transmit power is allocated to the common stream precoder. Consequently, the common stream alone should be able to support an MCS level whose throughput is at least 42Mbps (i.e., the NOMA sum throughput), which would then be equally split between the two groups to achieve fairness. However, from Table II, we see that there is no MCS level close to 42Mbps. Hence, in Table V, we see that the common stream supports MCS level 5 (36Mbps) which is the nearest level lower than 42Mbps, but not MCS level 6 (48Mbps) which is the nearest level above 42Mbps. This is the reason why the NOMA sum throughput exceeds the RSMA sum throughput. However, even with conservative MCS level 5 for the common stream, RSMA-based MGM achieves a higher minimum throughput than NOMA-based MGM, which is the desired objective.

V Conclusion

In this paper, we presented the first-ever experimental evaluation of the fairness performance of RSMA-, SDMA- and NOMA-based MGM. We focused our attention on the overloaded scenario with closely located group members, which is relevant for many applications. For two groups with two users per group, we realized nine cases that captured varying levels of interference and pathloss difference between the two groups. Over these nine cases, we observed that RSMA-based MGM achieved superior fairness performance (i.e., fairness at a higher minimum throughput) than SDMA- and NOMA-based MGM. This is consistent with theoretical predictions. The secret behind these gains stems from the fact that each group’s throughput could, in general, have contributions from both the common and the private streams with the exact amounts of each dictated by the channel conditions.

References

  • [1] B. Elmali, D. M. Soleymani, and S. A. Ashraf, “Empowering public safety services using multicast in 5G-Advanced,” Nokia Blog, Jan. 2024. [Online]. Available: https://www.nokia.com/blog/empowering-public-safety-services-using-multicast-in-5g-advanced/
  • [2] 3GPP TS 26.517, “5G Multicast-Broadcast User Services; Protocols and Formats (rel. 17),” Sep. 2023.
  • [3] M. Á. Vázquez, A. Pérez-Neira, D. Christopoulos, S. Chatzinotas, B. Ottersten, P.-D. Arapoglou, A. Ginesi, and G. Taricco, “Precoding in Multibeam Satellite Communications: Present and Future Challenges,” IEEE Trans. Wireless Commun., vol. 23, no. 6, pp. 88–95, 2016.
  • [4] G. Araniti, M. Condoluci, P. Scopelliti, A. Molinaro, and A. Iera, “Multicasting over Emerging 5G Networks: Challenges and Perspectives,” IEEE Network, vol. 31, no. 2, pp. 80–89, 2017.
  • [5] E. Karipidis, N. D. Sidiropoulos, and Z.-Q. Luo, “Quality of service and max-min fair transmit beamforming to multiple cochannel multicast groups,” IEEE Trans. Signal Process., vol. 56, no. 3, pp. 1268–1279, 2008.
  • [6] T.-H. Chang, Z.-Q. Luo, and C.-Y. Chi, “Approximation bounds for semidefinite relaxation of Max-Min-Fair multicast transmit beamforming problem,” IEEE Trans. Signal Process., vol. 56, no. 8, pp. 3932–3943, 2008.
  • [7] D. Christopoulos, S. Chatzinotas, and B. Ottersten, “Weighted fair multicast multigroup beamforming under per-antenna power constraints,” IEEE Trans. Signal Process., vol. 62, no. 19, pp. 5132–5142, 2014.
  • [8] M. Dong and Q. Wang, “Multi-group multicast beamforming: Optimal structure and efficient algorithms,” IEEE Trans. Signal Process., vol. 68, pp. 3738–3753, 2020.
  • [9] E. Karipidis, N. D. Sidiropoulos, and Z.-Q. Luo, “Far-field multicast beamforming for uniform linear antenna arrays,” IEEE Trans. Signal Process., vol. 55, no. 10, pp. 4916–4927, 2007.
  • [10] W. Wang, A. Liu, Q. Zhang, L. You, X. Gao, and G. Zheng, “Robust multigroup multicast transmission for frame-based multi-beam satellite systems,” IEEE Access, vol. 6, pp. 46 074–46 083, 2018.
  • [11] N. Bornhorst, M. Pesavento, and A. B. Gershman, “Distributed beamforming for multi-group multicasting relay networks,” IEEE Transactions on Signal Processing, vol. 60, no. 1, pp. 221–232, 2012.
  • [12] D. Christopoulos, S. Chatzinotas, and B. Ottersten, “Multicast multigroup precoding and user scheduling for frame-based satellite communications,” IEEE Trans. Wireless Commun., vol. 14, no. 9, pp. 4695–4707, Sep. 2015.
  • [13] X. Zhu, C. Jiang, L. Yin, L. Kuang, N. Ge, and J. Lu, “Cooperative multigroup multicast transmission in integrated terrestrial-satellite networks,” IEEE J. Sel. Areas Commun., vol. 36, no. 5, pp. 981–992, 2018.
  • [14] Y. C. B. Silva and A. Klein, “Linear transmit beamforming techniques for the multigroup multicast scenario,” IEEE Trans. Veh. Technol., vol. 58, no. 8, pp. 4353–4367, Oct 2009.
  • [15] N.-D. Nguyen, R. Knopp, N. Nikaein, and C. Bonnet, “Implementation and validation of multimedia broadcast multicast service for LTE/LTE-advanced in OpenAirInterface platform,” in Proc. of the 38th Annual IEEE Conf. on Local Computer Networks (Workshops), Oct. 2013, pp. 70–76.
  • [16] D. Dujovne and T. Turletti, “Multicast in 802.11 WLANs: an experimental study,” in Proc. of the 9th ACM Intl. Symp. on Modeling, Analysis and Simulation of Wireless and Mobile Systems, 2006, p. 130–138.
  • [17] M. F. Hanif, Z. Ding, T. Ratnarajah, and G. K. Karagiannidis, “A Minorization-Maximization method for optimizing sum rate in the downlink of non-orthogonal multiple access systems,” IEEE Trans. Signal Process., vol. 64, no. 1, pp. 76–88, 2016.
  • [18] Z. Wang, J. Hu, G. Liu, and Z. Ma, “Optimal power allocations for relay-assisted NOMA-based 5G V2X broadcast/multicast communications,” in Proc. of the IEEE/CIC Intl. Conf. on Commun. in China (ICCC), 2018, pp. 688–693.
  • [19] S. M. Ivari, M. Caus, M. A. Vazquez, M. R. Soleymani, Y. R. Shayan, and A. I. Perez-Neira, “Power allocation and user clustering in Multicast NOMA based satellite communication systems,” in Proc. of the IEEE Intl. Conf. on Commun. (ICC), 2020, pp. 1–6.
  • [20] A. Ihsan, W. Chen, S. Zhang, and S. Xu, “Energy-efficient NOMA multicasting system for beyond 5G cellular V2X communications with imperfect CSI,” IEEE Trans. Intell. Transp. Syst., vol. 23, no. 8, pp. 10 721–10 735, 2022.
  • [21] J. Choi, “Minimum power multicast beamforming with superposition coding for multiresolution broadcast and application to NOMA systems,” IEEE Trans. Commun., vol. 63, no. 3, pp. 791–800, 2015.
  • [22] Y. Zhang, T.-X. Zheng, Q. Yang, H.-M. Wang, B. Wang, and Z. Li, “The application of Non-Orthogonal Multiple Access in 5G Physical-Layer Multi-Region geocast,” in 2017 IEEE Wireless Communications and Networking Conference (WCNC), 2017, pp. 1–6.
  • [23] B. Clerckx, Y. Mao, R. Schober, E. A. Jorswieck, D. J. Love, J. Yuan, L. Hanzo, G. Y. Li, E. G. Larsson, and G. Caire, “Is NOMA efficient in multi-antenna networks? A critical look at next generation multiple access techniques,” IEEE Open Journal of the Communications Society, vol. 2, pp. 1310–1343, 2021.
  • [24] H. Joudeh and B. Clerckx, “Rate-Splitting for Max-Min fair multigroup multicast beamforming in overloaded systems,” IEEE Trans. Wireless Commun., vol. 16, no. 11, pp. 7276–7289, 2017.
  • [25] Y. Mao, O. Dizdar, B. Clerckx, R. Schober, P. Popovski, and H. V. Poor, “Rate-Splitting Multiple Access: Fundamentals, Survey, and Future Research Trends,” IEEE Commun. Surveys Tuts., 2022.
  • [26] X. Lyu, S. Aditya, J. Kim, and B. Clerckx, “Rate-Splitting Multiple Access: The First Prototype and Experimental Validation of its Superiority over SDMA and NOMA,” IEEE Trans. Wireless Commun., pp. 1–1, 2024.
  • [27] B. Clerckx, Y. Mao, E. A. Jorswieck, J. Yuan, D. J. Love, E. Erkip, and D. Niyato, “A Primer on Rate-Splitting Multiple Access: Tutorial, Myths, and Frequently Asked Questions,” IEEE J. Sel. Areas Commun., vol. 41, no. 5, pp. 1265–1308, May 2023.
  • [28] A. Z. Yalcin, M. Yuksel, and B. Clerckx, “Rate splitting for multi-group multicasting with a common message,” IEEE Trans. Veh. Technol., vol. 69, no. 10, pp. 12 281–12 285, Oct 2020.
  • [29] L. Yin and B. Clerckx, “Rate-splitting multiple access for multigroup multicast and multibeam satellite systems,” IEEE Trans. Commun., vol. 69, no. 2, pp. 976–990, Feb 2021.
  • [30] H. Chen, D. Mi, T. Wang, Z. Chu, Y. Xu, D. He, and P. Xiao, “Rate-splitting for multicarrier multigroup multicast: Precoder design and error performance,” IEEE Trans. Broadcast., vol. 67, no. 3, pp. 619–630, 2021.
  • [31] L. Yin, O. Dizdar, and B. Clerckx, “Rate-splitting multiple access for multigroup multicast cellular and satellite communications: PHY layer design and link-level simulations,” in Proc. of the IEEE Intl. Conf. on Communications (ICC) Workshops, 2021, pp. 1–6.
  • [32] H. Cui, L. Zhu, Z. Xiao, B. Clerckx, and R. Zhang, “Energy-efficient rsma for multigroup multicast and multibeam satellite communications,” IEEE Wireless Commun. Lett., vol. 12, no. 5, pp. 838–842, May 2023.
  • [33] H. Joudeh and B. Clerckx, “Sum-rate maximization for linearly precoded downlink multiuser MISO systems with partial CSIT: A rate-splitting approach,” IEEE Trans. Commun., vol. 64, no. 11, pp. 4847–4861, 2016.
  • [34] S. S. Christensen, R. Agarwal, E. De Carvalho, and J. M. Cioffi, “Weighted sum-rate maximization using weighted mmse for mimo-bc beamforming design,” IEEE Trans. Wireless Commun., vol. 7, no. 12, pp. 4792–4799, Dec. 2008.
  • [35] C. Mosquera and F. Gómez-Cuba, “Link Adaptation for Rate Splitting Systems with partial CSIT,” IEEE J. Sel. Areas Commun., vol. 41, no. 5, pp. 1336–1350, May 2023.
  • [36] H. Minn, “A robust timing and frequency synchronization for OFDM systems,” IEEE Trans. Wireless Commun., vol. 2, no. 4, pp. 822–839, 2003.
  • [37] P. Trifonov, “Efficient Design and Decoding of Polar Codes,” IEEE Trans. Commun., vol. 60, no. 11, pp. 3221–3227, 2012.
  • [38] H. Li and J. Yuan, “A Practical Construction Method for Polar Codes in AWGN Channels,” in IEEE 2013 Tencon - Spring, 2013, pp. 223–226.
  • [39] I. Tal and A. Vardy, “List Decoding of Polar Codes,” IEEE Trans. Inf. Theory, vol. 61, no. 5, pp. 2213–2226, 2015.