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

Spatial Multiplexing Oriented Channel Reconfiguration in Multi-IRS Aided MIMO Systems

Yuxuan Chen, Qingqing Wu, Senior Member, IEEE, Guangji Chen, Wen Chen, Senior Member, IEEE Yuxuan Chen, Qingqing Wu, and Wen Chen are with the Department of Electronic Engineering, Shanghai Jiao Tong University, Shanghai 200240, China (e-mail: [email protected]; [email protected]; [email protected]). Guangji Chen is with Nanjing University of Science and Technology, Nanjing 210094, China (e-mail: [email protected]).
Abstract

Spatial multiplexing plays a significant role in improving the capacity of multiple-input multiple-output (MIMO) communication systems. To improve the spectral efficiency (SE) of a point-to-point MIMO system, we exploit the channel reconfiguration capabilities provided by multiple intelligent reflecting surfaces (IRSs) to enhance the spatial multiplexing. Unlike most existing works, we address both the issues of the IRSs placement and elements allocation. To this end, we first introduce an orthogonal placement strategy to mitigate channel correlation, thereby enabling interference-free multi-stream transmission. Subsequently, we propose a successive convex approximation (SCA)-based approach to jointly optimize the IRS elements and power allocation. Our theoretical analysis unveils that equal IRS elements/power allocation scheme becomes asymptotically optimal as the number of IRS elements and transmit power tend to be infinite. Numerical results demonstrate that when the total number of IRS elements or the power exceeds a certain threshold, a multi-IRS assisted system outperforms a single IRS configuration.

Index Terms:
IRS, elements allocation, deployment.

I Introduction

Over the past decade, multiple-input multiple-output (MIMO) systems have significantly enhanced network throughput by simultaneously transmitting multi-stream via the spatial domain. Particularly, the capacity of MIMO systems increases linearly with the number of sub-channels, which is mainly determined by the rank of the channel matrix[1]. However, the number of available sub-channels is constrained by the number of independent propagation paths and antennas.

Recently, intelligent reflecting surface (IRS) has emerged as a promising technology for future wireless systems due to its ability to create customizable propagation environments with lower hardware cost and power consumption compared to traditional antenna arrays [2, 3, 4]. Additionally, IRS can be densely deployed to facilitate data transmission. In addition to passive beamforming, the optimization of IRS deployment provides another degree of freedom for realizing channel customization. For a point-to-point link, the seminal work[5] unveiled that the passive IRS should be deployed near the transmitter/receiver (Tx/Rx) to minimize the cascaded channel path loss. While the above work focused on applying the deployment of an IRS to increase the received power, it is also appealing to make use of IRS deployment to enhance spatial multiplexing gains of MIMO systems. In terms of IRS deployment in MIMO systems,[6] and [7] demonstrated that the deployment of IRSs is able to create favorable channels with controllable rank and favorable condition numbers. For a multi-user setup, [8] rigorously proved that distributed IRS configurations outperform centralized IRS when the total number of IRS elements exceeds a certain threshold. Unfortunately, no existing works have investigated how to configure IRS positions and their corresponding elements in multi-IRS aided MIMO systems to achieve a balance between spatial multiplexing gain and passive beamforming gain, which thus motivates our work.

Refer to caption
Figure 1: A MIMO wireless communication system aided by KK IRSs.

In this paper, we focus on a multi-IRS assisted point-to-point MIMO system, as illustrated in Fig. 1. We aim to configure a favorable wireless propagation environment for supporting a multi-stream transmission. To this end, the positions and beamforming of the IRSs have to be designed to maximize the exploitation of the capability of spatial multiplexing while simultaneously minimizing the link path loss. Furthermore, the number of elements and transmit power for each data stream should be carefully allocated to maximize the spectral efficiency (SE). Aiming to address these issues, the main contributions of this work are summarized as follows. First, we propose an IRSs placement scheme to create the orthogonal sub-channels and thereby the decoupled spatial correlation can be fulfilled. Under the orthogonal IRS placement, we unveil that the optimal IRS phase shifts are set to maximize the power gain of each individual sub-channel. Second, we propose an efficient algorithm that jointly optimizes the IRS elements allocation and power to maximize the SE. Moreover, we analytically characterize the scaling orders of the system SE with respect to the number of reflecting elements and power. Finally, numerical results are provided to compare the performance of the multi-IRS aided wireless system with various benchmark systems. Compared to a single large IRS, the potential benefits of multiple IRSs are fully unleashed as the increase of the total number of elements or transmit power.

II System Model and Problem Formulation

II-A System Model

As shown in Fig. 1, we consider a MIMO wireless communication system assisted by KK IRSs, where the Tx and the Rx are equipped with NtN_{t} and NrN_{r} antennas, respectively. We assume that the direct Tx-Rx link is obstructed. The positions of the Tx and the Rx in a two-dimensional (2D) Cartesian coordinate system are denoted by 𝐮t2×1\mathbf{u}_{t}\in\mathbb{R}^{2\times 1} and 𝐮r2×1\mathbf{u}_{r}\in\mathbb{R}^{2\times 1}, respectively. The Tx and the Rx employ uniform linear arrays (ULAs) with element spacings of dtd_{t} and drd_{r}, respectively. The kkth IRS is equipped with an Mk(Mv,k×Mh,k)M_{k}\,(M_{\mathrm{v,k}}\times M_{\mathrm{h,k}}) elements uniform planar array (UPA), where k𝒦=Δ{1,2,,K}k\in{\mathcal{K}}\buildrel\Delta\over{=}\{1,2,\ldots,K\}. The UPA on the kkth IRS consists of Mv,kM_{\mathrm{v,k}} rows and Mh,kM_{\mathrm{h,k}} columns, with all spacings equal to dsd_{s}. The total available number of IRSs is MM and thus k=1KMkM\sum_{k=1}^{K}{M_{k}}\leqslant M. The position of the kkth IRS is denoted by 𝐮k2×1\mathbf{u}_{k}\in\mathbb{R}^{2\times 1}.

We denote the equivalent channel from the Tx to the kkth IRS and from the kkth IRS to the Rx as 𝐓kMk×Nt\mathbf{T}_{k}\in\mathbb{C}^{M_{k}\times N_{t}} and 𝐑kNr×Mk\mathbf{R}_{k}\in\mathbb{C}^{N_{r}\times M_{k}}, respectively. We assume that the distributed IRSs possess line-of-sight (LoS) paths between the Tx and the Rx. The channel from the Tx to the kkth IRS can be expressed as

𝐓k=ρT,k𝐚S,k(ΦT,kAOA,ΘT,kAOA)𝐚NtH(ΘT,kAOD),\displaystyle\mathbf{T}_{k}=\rho_{\mathrm{T},k}\mathbf{a}_{\mathrm{S},k}\left(\Phi_{\mathrm{T},k}^{\mathrm{AOA}},\Theta_{\mathrm{T},k}^{\mathrm{AOA}}\right)\mathbf{a}_{N_{t}}^{H}\left(\Theta_{\mathrm{T},k}^{\mathrm{AOD}}\right), (1)

where ρT,k\rho_{\mathrm{T},k} denotes the complex channel gain of the Tx to the kkth IRS link. In the array response vector, ΦT,kAOA=2πdScosϕT,kAOA/λ{\Phi_{{\mathrm{T}},k}^{\mathrm{AOA}}}=2\pi d_{\mathrm{S}}\cos{\phi_{{\mathrm{T}},k}^{\mathrm{AOA}}}/\lambda, ΘT,kAOA=2πdSsinϕT,kAOAsinθT,kAOA/λ{\Theta_{{\mathrm{T}},k}^{\mathrm{AOA}}}=2\pi d_{\mathrm{S}}\sin{\phi_{{\mathrm{T}},k}^{\mathrm{AOA}}}\sin{\theta_{{\mathrm{T}},k}^{\mathrm{AOA}}}/\lambda, and ΘT,kAOD=2πdtsinθT,kAOD/λ{\Theta_{{\mathrm{T}},k}^{\mathrm{AOD}}}=2\pi d_{{t}}\sin{\theta_{{\mathrm{T}},k}^{\mathrm{AOD}}}/\lambda, where θT,kAOA{\theta_{{\rm{T}},k}^{\rm{AOA}}}, ϕT,kAOA{\phi_{{\rm{T}},k}^{\rm{AOA}}}, and θT,kAOD{\theta_{{\rm{T}},k}^{\rm{AOD}}} denote the horizontal angle of arrival (AoA), the vertical AoA, and the angle of departure (AoD) of the Tx to kkth IRS link, respectively. Furthermore, 𝐚Nt(){{\bf{a}}_{N_{t}}}\left(\cdot\right) and 𝐚S,k(){{\bf{a}}_{{\rm{S}},k}}\left(\cdot\right) represent the array response vectors at the Tx and the kkth IRS, respectively. Hence, the array response vector of ULA can be unified by

𝐚N(X)=[1,ejX,,ejX(N1)]T.\displaystyle\mathbf{a}_{N}\left(X\right)=\left[1,e^{jX},...,e^{jX\left(N-1\right)}\right]^{T}. (2)

It is worth noting that the array response for UPA can be decomposed into that of ULA as 𝐚S,k(X,Y)=𝐚Mv,k(X)𝐚Mh,k(Y){{\bf{a}}_{{\rm{S}},k}}\left({X,Y}\right)={{\bf{a}}_{{M_{v,k}}}}\left(X\right)\otimes{{\bf{a}}_{{M_{h,k}}}}\left(Y\right), where \otimes is the Kronecker product.

Similar to the Tx to the kkth IRS link, the channel matrix from the kkth IRS to the Rx can be expressed as

𝐑k=ρR,k𝐚Nr(ΘR,kAOA)𝐚S,kH(ΦR,kAOD,ΘR,kAOD),\displaystyle\mathbf{R}_{k}=\rho_{\mathrm{R},k}\mathbf{a}_{N_{r}}\left(\Theta_{\mathrm{R},k}^{\mathrm{AOA}}\right)\mathbf{a}_{\mathrm{S},k}^{H}\left(\Phi_{\mathrm{R},\mathrm{k}}^{\mathrm{AOD}},\Theta_{\mathrm{R},k}^{\mathrm{AOD}}\right), (3)

where ρR,k\rho_{\mathrm{R},k} denotes the complex channel gain of the kkth IRS to the Rx link. For notational convenience, in the sequel we substitute 𝐚S,k,T\mathbf{a}_{\mathrm{S},k,\mathrm{T}} and 𝐚S,k,R\mathbf{a}_{\mathrm{S},k,\mathrm{R}} for 𝐚S,k(ΦT,kAOA,ΘT,kAOA)\mathbf{a}_{\mathrm{S},k}\left(\Phi_{\mathrm{T},k}^{\mathrm{AOA}},\Theta_{\mathrm{T},k}^{\mathrm{AOA}}\right) and 𝐚S,k(ΦR,kAOD,ΘR,kAOD)\mathbf{a}_{\mathrm{S},k}\left(\Phi_{\mathrm{R},\mathrm{k}}^{\mathrm{AOD}},\Theta_{\mathrm{R},k}^{\mathrm{AOD}}\right), respectively. Besides, we define ρkρT,kρR,k\rho_{k}\triangleq\rho_{\mathrm{T},k}\rho_{\mathrm{R},k} and we denote 𝚽k=diag(ejϕk,1,ejϕk,2,,ejϕk,Mk)\mathbf{\Phi}_{k}=\mathrm{diag}(e^{j\phi_{k,1}},e^{j\phi_{k,2}},\cdots,e^{j\phi_{k,M_{k}}}) as the passive beamforming matrix of the kkth IRS, where ϕk,mk\phi_{k,m_{k}} is the phase of the mkm_{k}th element on the kkth IRS, mkk{1,2,,Mk}m_{k}\in\mathcal{M}_{k}\triangleq\left\{1,2,…,M_{k}\right\}, k𝒦k\in\mathcal{K}, and ϕk,mk[0,2π)\phi_{k,m_{k}}\in\left[0,2\pi\right). As such, the effective Tx-IRS-Rx MIMO channel aided by KK IRSs is given by 𝐇=k=1K𝐑k𝚽k𝐓k\mathbf{H}=\sum_{k=1}^{K}{\mathbf{R}_{k}\mathbf{\Phi}_{k}\mathbf{T}_{k}}.

Let f(𝚽k)𝐚S,k,RH𝚽k𝐚S,k,Tf\left(\mathbf{\Phi}_{k}\right)\triangleq\mathbf{a}_{\mathrm{S},k,\mathrm{R}}^{H}\mathbf{\Phi}_{k}\mathbf{a}_{\mathrm{S},k,\mathrm{T}}, ωk=(ρkf(𝚽k))\omega_{k}=\angle(\rho_{k}f\left(\mathbf{\Phi}_{k}\right)), 𝐀T=1Nt[ejω1𝐚Nt(ΘT,1AOD),,ejωK𝐚Nt(ΘT,KAOD)]\mathbf{A}_{\mathrm{T}}=\frac{1}{\sqrt{N_{t}}}[e^{j\omega_{1}}\mathbf{a}_{N_{t}}\left(\Theta_{\mathrm{T},1}^{\mathrm{AOD}}\right),...,e^{j\omega_{K}}\mathbf{a}_{N_{t}}\left(\Theta_{\mathrm{T},K}^{\mathrm{AOD}}\right)], and 𝐀R=1Nr[𝐚Nr(ΘR,1AOA),,𝐚Nr(ΘR,KAOA)]\mathbf{A}_{\mathrm{R}}=\frac{1}{\sqrt{N_{r}}}[\mathbf{a}_{N_{r}}\left(\Theta_{\mathrm{R},1}^{\mathrm{AOA}}\right),...,\mathbf{a}_{N_{r}}\left(\Theta_{\mathrm{R},K}^{\mathrm{AOA}}\right)]. Thus, the composite channel consisting of KK IRSs can be simplified as

𝐇=k=1Kρk𝐚Nr(ΘR,kAOA)f(𝚽k)𝐚NtH(ΘT,kAOD)=𝐀R𝚺𝐀TH,\displaystyle\!\!\!\mathbf{H}\!=\!\sum_{k=1}^{K}{\rho_{k}\mathbf{a}_{N_{r}}\left(\Theta_{\mathrm{R},k}^{\mathrm{AOA}}\right)f\left(\mathbf{\Phi}_{k}\right)}\mathbf{a}_{N_{t}}^{H}\left(\Theta_{\mathrm{T},k}^{\mathrm{AOD}}\right)\!\!=\!\mathbf{A}_{\mathrm{R}}\mathbf{\Sigma A}_{\mathrm{T}}^{H}, (4)

where 𝚺=NtNrdiag(|ρ1f(𝚽1)|,,|ρKf(𝚽K)|)\mathbf{\Sigma}=\sqrt{N_{t}N_{r}}\mathrm{diag}\left(|\rho_{1}f\left(\mathbf{\Phi}_{1}\right)|,...,|\rho_{K}f\left(\mathbf{\Phi}_{K}\right)|\right). As such, the received signal 𝐲Nr×1\mathbf{y}\in\mathbb{C}^{N_{r}\times 1} at the Rx is given by 𝐲=𝐇𝐱+𝐳\mathbf{y}=\mathbf{Hx}+\mathbf{z}, where 𝐱Nt×1\mathbf{x}\in\mathbb{C}^{N_{t}\times 1} denotes the transmitted signal vector and 𝐳𝒞𝒩(0,σ2𝐈Nr)\mathbf{z}\sim\mathcal{CN}(0,\sigma^{2}\mathbf{I}_{N_{r}}) is the additive white Gaussian noise at the Rx with power σ2\sigma^{2}.

The introduction of IRSs enables the creation of a controllable scattering channel environment, potentially establishing favorable conditions for multi-stream transmission. By properly positioning IRSs so that 𝐀T\mathbf{A}_{\mathrm{T}} and 𝐀R\mathbf{A}_{\mathrm{R}} are orthogonal matrices (i.e., 𝐀RH𝐀R=𝐈\mathbf{A}_{\mathrm{R}}^{H}\mathbf{A}_{\mathrm{R}}=\mathbf{I} and 𝐀TH𝐀T=𝐈\mathbf{A}_{\mathrm{T}}^{H}\mathbf{A}_{\mathrm{T}}=\mathbf{I}), the equation (4) naturally admits a form of the singular value decomposition (SVD). In this scenario, the structures of the transmitter’s precoding and the receiver’s combiner can be explicitly derived as 𝐀T\mathbf{A}_{\mathrm{T}} and 𝐀RH\mathbf{A}_{\mathrm{R}}^{H}, respectively. This configuration allows for the minimization of inter-stream interference while simplifying the transceiver design.

II-B Problem Formulation

We aim to maximize the SE of the multi-IRS aided system by optimizing the IRS beamforming, IRS placement, elements allocation, and transmit covariance matrix. The corresponding optimization problem is formulated as

max{uk},𝐐{ϕk,mk},{Mk}\displaystyle\mathop{\mathrm{max}}\limits_{{\mathop{\{u_{k}\},\mathbf{Q}}\limits^{\{\phi_{k,m_{k}}\},\{M_{k}\}}}} log2det(𝐈Nr+1σ2𝐇𝐐𝐇H)\displaystyle\log_{2}\det\left(\mathbf{I}_{N_{r}}+\frac{1}{\sigma^{2}}\mathbf{HQH}^{H}\right) (5)
s.t. ϕk,mk[0,2π),mkk,k𝒦,\displaystyle\phi_{k,m_{k}}\in\left[0,2\pi\right),m_{k}\in\mathcal{M}_{k},k\in\mathcal{K}, (5a)
tr(𝐐)P,𝐐𝟎,\displaystyle\mathrm{tr(}\mathbf{Q})\leq P,\mathbf{Q}\succeq\mathbf{0}, (5b)
𝐀RH𝐀R=𝐈,𝐀TH𝐀T=𝐈,\displaystyle\mathbf{A}_{\mathrm{R}}^{H}\mathbf{A}_{\mathrm{R}}=\mathbf{I},\mathbf{A}_{\mathrm{T}}^{H}\mathbf{A}_{\mathrm{T}}=\mathbf{I}, (5c)
k=1KMkM,Mk0,k𝒦,\displaystyle\sum\nolimits_{k=1}^{K}{M_{k}}\leqslant M,M_{k}\in\mathbb{Z}_{\geqslant 0},k\in\mathcal{K}, (5d)

where 𝐐=𝔼{𝐱𝐱H}\mathbf{Q}=\mathbb{E}\{\mathbf{x}\mathbf{x}^{H}\} denotes the transmit covariance matrix and PP denotes the maximum transmit power of the Tx.

III Proposed Solution

Problem (5) is non-convex due to its non-concave objective function, IRS position constraints, and integer elements allocation constraints. Generally, there are no standard methods to solve it optimally. To address this issue, we first propose an efficient solution to determine the IRS placement. By exploiting the particular structure provided by the IRS placement, we derive the optimal IRS beamforming in a closed form expression. Subsequently, the IRS elements allocation and the power allocation are jointly optimized.

III-A IRS Placement and Beamforming Design

First, to satisfy the constraint (5c), IRSs should be positioned along the discrete fourier transform (DFT) directions of both the Tx and the Rx. This arrangement satisfies the requirement for the AoA and AoD discretization:

{ΘT,kAOD𝒜1{2πiNtπ}i=1Nt,k𝒦,ΘR,kAOA𝒜2{2πiNrπ}i=1Nr,k𝒦,ΘT,iAODΘT,jAOD,ΘR,iAOAΘR,jAOA,i,j𝒦,ij.\displaystyle\begin{cases}\Theta_{\mathrm{T},k}^{\mathrm{AOD}}\in\mathcal{A}_{1}\triangleq\left\{\frac{2\pi i}{N_{t}}-\pi\right\}_{i=1}^{N_{t}},k\in\mathcal{K},\\ \Theta_{\mathrm{R},k}^{\mathrm{AOA}}\in\mathcal{A}_{2}\triangleq\left\{\frac{2\pi i}{N_{r}}-\pi\right\}_{i=1}^{N_{r}},k\in\mathcal{K},\\ \Theta_{\mathrm{T},i}^{\mathrm{AOD}}\neq\Theta_{\mathrm{T},j}^{\mathrm{AOD}},\Theta_{\mathrm{R},i}^{\mathrm{AOA}}\neq\Theta_{\mathrm{R},j}^{\mathrm{AOA}},i,j\in\mathcal{K},i\neq j.\\ \end{cases} (6)

When the positions of the Tx and the Rx as well as (ΘT,kAOD,ΘR,kAOA)(\Theta_{\mathrm{T},k}^{\mathrm{AOD}},\Theta_{\mathrm{R},k}^{\mathrm{AOA}}) are given, we can determine the corresponding IRS position 𝐮k\mathbf{u}_{k}. However, obtaining the optimal IRS positions requires an exhaustive enumeration of all combinations due to (6), which results in high computation complexity. To address this problem, we employ a greedy algorithm that searches for the position with the maximum channel gain at each iteration. Besides, we define 𝒫1{𝒮1,𝒮2,,𝒮L}\mathcal{P}_{1}\triangleq\left\{\mathcal{S}_{1},\mathcal{S}_{2},…,\mathcal{S}_{L}\right\} that collects all the possible IRS candidate positions and their corresponding Tx-IRS-Rx channel gain, where LL denotes the maximum number of different sets in 𝒫1\mathcal{P}_{1}, 𝒮l(θl,φl,τl)\mathcal{S}_{l}\triangleq\left(\theta_{l},\varphi_{l},\tau_{l}\right) denotes the IRS position where the AoD at the Tx is θl\theta_{l} and the AoA at the Rx is φl\varphi_{l} with the corresponding Tx-IRS-Rx channel gain τl\tau_{l}, θl𝒜1\theta_{l}\in\mathcal{A}_{1}, φl𝒜2\varphi_{l}\in\mathcal{A}_{2}, l{1,2,,L}l\in\mathcal{L}\triangleq\left\{1,2,…,L\right\}, and 𝒮i𝒮j,ij,i,j\mathcal{S}_{i}\neq\mathcal{S}_{j},i\neq j,i,j\in\mathcal{L}. Then, the process of the greedy search can be described by

(ΘT,k+1AOD,ΘR,k+1AOA,ρk+1)=argmax𝒮l𝒫k+1τls.t.𝒫k+1=𝒫k\k,k1,k={𝒮i𝒫k|θi=ΘT,kAOD or φi=ΘR,kAOA}.\displaystyle\begin{aligned} &\left(\Theta_{\mathrm{T},k+1}^{\mathrm{AOD}},\Theta_{\mathrm{R},k+1}^{\mathrm{AOA}},\rho_{k+1}\right)=\mathop{\mathrm{arg}\max}_{\mathcal{S}_{l}\in\mathcal{P}_{k+1}}\tau_{l}\\ &\text{s.t.}\ \begin{aligned} &\mathcal{P}_{k+1}=\mathcal{P}_{k}\backslash\mathcal{B}_{k},k\geqslant 1,\\ &\mathcal{B}_{k}=\left\{\mathcal{S}_{i}\in\mathcal{P}_{k}\,|\,\theta_{i}=\Theta_{\mathrm{T},k}^{\mathrm{AOD}}\text{ or }\varphi_{i}=\Theta_{\mathrm{R},k}^{\mathrm{AOA}}\right\}.\end{aligned}\end{aligned} (7)

Therefore, we execute this greedy search process from k=0k=0 to K1K-1 and successfully obtain the positions for KK IRSs.

By employing the orthogonal placement, the kkth singular value of 𝐇\mathbf{H} is the non-zero singular value of 𝐑k𝚽k𝐓k\mathbf{R}_{k}\mathbf{\Phi}_{k}\mathbf{T}_{k}. And 𝐐\mathbf{Q} in (5) can be derived as 𝐐=𝐀Tdiag(p1,p2,,pK)𝐀TH\mathbf{Q}=\mathbf{A}_{\mathrm{T}}\mathrm{diag}\left(p_{1},p_{2},…,p_{K}\right)\mathbf{A}_{\mathrm{T}}^{H}, where pkp_{k} is the transmit power allocated to the kkth sub-channel that will be optimized in subsection B. Thus, the SE of the KK IRS system can be rewritten as

R=k=1Klog2(1+pkχk(f(𝚽k))2),\displaystyle R=\sum\nolimits_{k=1}^{K}{\log_{2}}\left(1+p_{k}\chi_{k}\left(f\left(\mathbf{\Phi}_{k}\right)\right)^{2}\right), (8)

where χk=|ρk|2/σ2\chi_{k}=|\rho_{k}|^{2}/\sigma^{2}. Due to the fact that maximizing the SE can be achieved by independently maximizing each f(𝚽k)f(\mathbf{\Phi}_{k}), we employ the following passive beamforming structure to optimize each f(𝚽k)f(\mathbf{\Phi}_{k}):

ϕk,mk=arg((𝐭k)mk(𝐫k)mk),mkk,k𝒦,\displaystyle\phi_{k,m_{k}}=\mathrm{arg}\left((\mathbf{t}_{k})_{m_{k}}^{\ast}(\mathbf{r}_{k})_{m_{k}}^{\ast}\right),m_{k}\in\mathcal{M}_{k},k\in\mathcal{K}, (9)

where (𝐭k)mk(\mathbf{t}_{k})_{m_{k}} is the mkm_{k}th element of 𝐚NtH(ΘT,kAOD)\mathbf{a}_{N_{t}}^{H}\left(\Theta_{\mathrm{T},k}^{\mathrm{AOD}}\right) and (𝐫k)mk(\mathbf{r}_{k})_{m_{k}} is the mkm_{k}th element of 𝐚Nr(ΘR,kAOA)\mathbf{a}_{N_{r}}\left(\Theta_{\mathrm{R},k}^{\mathrm{AOA}}\right). By employing the above configuration, f(𝚽k)f\left(\mathbf{\Phi}_{k}\right) can be maximized. Under the optimal 𝚽k\mathbf{\Phi}_{k}^{*}, we have f(𝚽k)=Mkf\left(\mathbf{\Phi}_{k}^{*}\right)=M_{k}.

III-B Joint IRS Elements Allocation and Power Optimization

To better illustrate the spatial multiplexing gain offered by multiple IRSs, we unveil the condition that double-IRS outperforms single-IRS in the following proposition. Besides, we define R1=log2(1+χPM2)R_{1}=\log_{2}\left(1+\chi PM^{2}\right) is the maximum achievable rate with a single IRS comprising MM elements and R2=maxp1,p2(log2(1+p1η1)+log2(1+p2η2))R_{2}=\underset{p_{1},p_{2}}{\max}\left(\log_{2}\left(1+p_{1}\eta_{1}\right)+\log_{2}\left(1+p_{2}\eta_{2}\right)\right) with p1+p2=Pp_{1}+p_{2}=P and ηk=χkM2/4\eta_{k}=\chi_{k}M^{2}/4 is the maximum achievable rate with two IRSs, each comprising M1=M2=M/2M_{1}=M_{2}=M/2 elements under orthogonal placement conditions.

Proposition 1

When χ1=χ2=χ\chi_{1}=\chi_{2}=\chi, we have R2R1R_{2}\geqslant R_{1} if

χPM248.\displaystyle\chi PM^{2}\geqslant 48. (10)
Proof:

According to the water-filling power allocation[9], for R2R_{2} the optimal pkp_{k}^{*} can be derived as pk=max(u1/ηk,0),k{1,2}p_{k}^{*}=\max\left(u-1/\eta_{k},0\right),k\in\left\{1,2\right\}. Since η1=η2\eta_{1}=\eta_{2}, we further have pk=P/2p_{k}^{*}=P/2 and R2=2log2(1+χPM2/8)R_{2}=2\log_{2}\left(1+\chi PM^{2}/8\right). To satisfy R2R1R_{2}\geqslant R_{1}, we can obtain (10), which completes the proof. ∎

Proposition 1 theoretically illustrates that, under specific conditions and with appropriate allocation strategies, the performance of a double IRS system can surpass that of a single IRS configuration. Furthermore, we aim to fully exploit the spatial multiplexing potential of multi-IRS architectures through appropriate resource allocation strategies, which motivates us to investigate the joint power and elements allocation algorithm. Next, we jointly optimize the IRS elements and power allocation under the obtained IRS placement and beamforming. To this end, problem (5) is reduced to

(P1):max{pk},{Mk}\displaystyle(\mathrm{P1}):\ \underset{\left\{p_{k}\right\},\left\{M_{k}\right\}}{\max} k=1Klog2(1+pkMk2χk)\displaystyle\sum\nolimits_{k=1}^{K}{\log_{2}}\left(1+p_{k}M_{k}^{2}\chi_{k}\right) (11)
s.t. k=1KpkP,pk0,k𝒦,\displaystyle\sum\nolimits_{k=1}^{K}{p_{k}}\leqslant P,p_{k}\geqslant 0,k\in\mathcal{K}, (11a)
(5d).\displaystyle\eqref{Element Constraints}.

First, to address the integer constraint in (5d), we relax the discrete value MkM_{k} to its continuous counterpart M~k\tilde{M}_{k}. Consequently, the objective function of problem (11) can be relaxed to

k=1Klog2(1+pkM~k2χk).\displaystyle\sum\nolimits_{k=1}^{K}{\log_{2}}\left(1+p_{k}\tilde{M}_{k}^{2}\chi_{k}\right). (12)

However, (12) is non-convex due to the coupling of pkp_{k} and M~k\tilde{M}_{k}. To tackle this difficulty, we introduce auxiliary variables lkl_{k}. Problem (P1)(\mathrm{P1}) can be reformulated as the following optimization problem:

max{pk},{M~k},{lk}\displaystyle\underset{\left\{p_{k}\right\},\left\{\tilde{M}_{k}\right\},\left\{l_{k}\right\}}{\max} k=1Klog2(1+lk)\displaystyle\sum\nolimits_{k=1}^{K}{\log_{2}}\left(1+l_{k}\right) (13)
s.t. lkχkpkM~k2,k𝒦,\displaystyle l_{k}\leqslant\chi_{k}p_{k}\tilde{M}_{k}^{2},k\in\mathcal{K}, (13a)
k=1KM~kM,M~k0,k𝒦,\displaystyle\sum\nolimits_{k=1}^{K}{\tilde{M}_{k}}\leqslant M,\tilde{M}_{k}\geqslant 0,k\in\mathcal{K}, (13b)
(11a).\displaystyle\eqref{Power Constraints}.

Problem (13) is still intractable hard to solve due to (13a). To address the non-convexity of constraint (13a), we employ the successive convex approximation (SCA) technique with appropriate variable substitutions. Thus, we introduce auxiliary variables xkx_{k} and yky_{k}, k𝒦k\in\mathcal{K}. Consequently, the constraints in (13a) can be reformulated as follows:

lkχkexk+yk,eykM~k2,k𝒦,\displaystyle l_{k}\leqslant\chi_{k}e^{x_{k}+y_{k}},e^{y_{k}}\leqslant\tilde{M}_{k}^{2},k\in\mathcal{K}, (14)
exkpk,k𝒦.\displaystyle e^{x_{k}}\leqslant p_{k},k\in\mathcal{K}. (15)

Although constraints (14) exhibit non-convex forms, their right-hand sides, specifically χkexk+yk\chi_{k}e^{x_{k}+y_{k}} and M~k2\tilde{M}_{k}^{2}, are convex functions with respect to their respective variables. This motivates us to apply the first-order Taylor expansion to linearize them as convex constraints given by

lkχkex^k+y^k(1+xk+ykx^ky^k),k𝒦,\displaystyle l_{k}\leqslant\chi_{k}e^{\hat{x}_{k}+\hat{y}_{k}}\left(1+x_{k}+y_{k}-\hat{x}_{k}-\hat{y}_{k}\right),k\in\mathcal{K}, (16)
eyk2M~kM^kM^k2,k𝒦,\displaystyle e^{y_{k}}\leqslant 2\tilde{M}_{k}\hat{M}_{k}-\hat{M}_{k}^{2},k\in\mathcal{K}, (17)

where x^k\hat{x}_{k}, y^k\hat{y}_{k}, and M^k\hat{M}_{k} are the given local points of xkx_{k} ,yky_{k}, and M~k\tilde{M}_{k}, respectively.

As a result, problem (13) is approximated as

(P2):max{pk},{M~k},{lk}\displaystyle(\mathrm{P2}):\ \underset{\left\{p_{k}\right\},\left\{\tilde{M}_{k}\right\},\left\{l_{k}\right\}}{\max} k=1Klog2(1+lk)\displaystyle\sum\nolimits_{k=1}^{K}{\log_{2}}\left(1+l_{k}\right) (18)
s.t. (11a),(13b),(15),(16),(17).\displaystyle\eqref{Power Constraints},\eqref{Continuous Element},\eqref{SCA3},\eqref{NewSCA1},\eqref{NewSCA2}.

Problem (P2)(\mathrm{P2}) is a convex optimization problem, which can be solved optimally in an iterative manner by using standard solvers like CVX until convergence is achieved. As the objective value increases with each iteration and is bounded by a finite value, the solution of the original problem (P1)(\mathrm{P1}) is ensured to reach convergence. The integer number of reflecting elements can be determined by rounding the continuous solutions to problem (P2)(\mathrm{P2}). Moreover, the complexity of the greedy search algorithm in (7) is smaller than 𝒪(KNtNr)\mathcal{O}\left(KN_{t}N_{r}\right) and the complexity of solving problem (P2)(\mathrm{P2}) via SCA is 𝒪(JK3.5)\mathcal{O}\left(JK^{3.5}\right) with JJ denoting the number of iteration while that of obtaining the optimal IRS beamforming is negligible due to the closed-form solution given in (9).

IV SE Scaling Order Analysis

In this section, we characterize the SE scaling order with respect to MM and PP for the multi-IRS aided wireless system. Based on the analysis in Section III, we first derive the elements allocation and power allocation scheme for the multi-IRS as MM approaches infinity.

Proposition 2

Under orthogonal placement conditions, as M{M}\to\infty and we assume that MM is a multiple of KK, the asymptotically optimal solution of problem (P1)(\mathrm{P1}), denoted by {pk,Mk}\left\{{p_{k}^{*},M_{k}^{*}}\right\}, is derived as

Mk=MK,pk=PK,k𝒦,\displaystyle M_{k}^{*}=\frac{M}{K},p_{k}^{*}=\frac{P}{K},k\in\mathcal{K}, (19)
Proof:

Note that it can be easily proved that the solution obtained when not using all IRS elements is always suboptimal as M{M}\to\infty. Under any given MkM_{k}, the optimal solution for pkp_{k} is pk=max(u1/(χkMk2),0)p_{k}=\max\left(u-1/\left(\chi_{k}M_{k}^{2}\right),0\right) according to the water-filling algorithm. As MM\to\infty, pk=u1/(χkMk2)p_{k}=u-1/\left(\chi_{k}M_{k}^{2}\right) holds naturally. Thus, the objective function of problem (11) can be simplified as

limMk=1Klog2(1+pkMk2χk)=k=1Klog2(pkMk2χk).\displaystyle\underset{M\rightarrow\infty}{\lim}\sum_{k=1}^{K}{\log_{2}}\left(1+p_{k}M_{k}^{2}\chi_{k}\right)=\sum_{k=1}^{K}{\log_{2}}\left(p_{k}M_{k}^{2}\chi_{k}\right). (20)

According to the Cauchy-Schwarz inequality, we have

Γk(k=1KMk)2Γk(k=1KMk/K)2K,\displaystyle\varGamma_{k}\left(\prod_{k=1}^{K}{M_{k}}\right)^{2}\leqslant\varGamma_{k}\left(\sum_{k=1}^{K}{M_{k}}/K\right)^{2K}, (21)

where Γk=k=1Kpkχk\varGamma_{k}=\prod_{k=1}^{K}{p_{k}\chi_{k}}, k=1KMk=M\sum_{k=1}^{K}{M_{k}}=M and the condition for equality in this case is Mk=M/K,k𝒦{M}_{k}^{*}={M}/K,k\in{\cal K}. Accordingly, the optimal power allocation in this case is P/K,k𝒦P/K,\forall k\in{\cal K} . Thus, we complete the proof. ∎

Proposition 2 reveals that, given a large number of elements, equal elements/power allocation can achieve near-optimal performance. This is because deploying a large number of IRS elements significantly enhances the magnitude of the singular values of the link, thereby artificially creating an equivalent "high-SNR" condition.

Proposition 3

Under orthogonal placement conditions, as M{M}\to\infty, the system SE increases with M{M} according to

limMRlog2M=2K.\displaystyle\lim_{M\rightarrow\infty}\frac{R}{\log_{2}M}=2K. (22)
Proof:

According to Proposition 2, when M{M}\to\infty we have

limMR=limMk=1K(log2(Γk/K)+2log2(M)).\displaystyle\lim_{M\rightarrow\infty}R=\lim_{M\rightarrow\infty}\sum_{k=1}^{K}{\left(\log_{2}\left(\varGamma_{k}/K\right)+2\log_{2}\left(M\right)\right)}. (23)

As such, we have (22), which completes the proof. ∎

Proposition 4

Under orthogonal placement conditions, as P{P}\to\infty, the system SE increases with PP according to

limMRlog2P=K.\displaystyle\lim_{M\rightarrow\infty}\frac{R}{\log_{2}P}=K. (24)
Proof:

Similar to Proposition 3, when P{P}\to\infty we have

limPR=limPk=1K(log2(Mk2χk/K)+log2(P)).\displaystyle\lim_{P\rightarrow\infty}R=\lim_{P\rightarrow\infty}\sum_{k=1}^{K}{\left(\log_{2}\left(M_{k}^{2}\chi_{k}/K\right)+\log_{2}\left(P\right)\right)}. (25)

As such, we have (24), which completes the proof. ∎

Proposition 3 and Proposition 4 demonstrate that compared to the single IRS system (K=1K=1), a KK-fold gain is realized in the SE scaling order of either MM or PP, which implies that we can potentially harvest a spatial multiplexing gain of a factor of KK from the system by deploying KK IRSs.

V Numerical Results

In this section, numerical results are provided to compare the performance of multiple IRSs and single IRS configurations, as well as to draw useful insights. The locations of the Tx and the Rx are set at (0, 0) and (85m, 0) respectively. We deploy a ULA at both the Tx and the Rx, aligning the base directions of the antenna arrays at the Tx and the Rx as 𝐯t=𝐯r=[0,1]T\mathbf{v}_{t}=\mathbf{v}_{r}=[0,-1]^{T}. Other system parameters are configured as follows: Nt=8N_{t}=8, Nr=4N_{r}=4, λ=0.15\lambda=0.15 m, H=5H=5 m, dt=dr=λ/2=0.075d_{t}=d_{r}=\lambda/2=0.075 m, and σ2=80\sigma^{2}=-80 dBm.

Fig. 2 and Fig. 3 compare the performance of the multi-IRS and the single-IRS aided systems under orthogonal placement. For the multi-IRS aided MIMO system, we employ our proposed algorithm to maximize the SE. Fig. 2 and Fig. 3 compare the achievable rates for different numbers of KK under the conditions of P=30P=30 dBm and MM = 2400, respectively. "KK=4 Average" strategy refers to the scenario where the power and elements are equally allocated. As shown in Fig. 2 and Fig. 3, the multi-IRS aided system significantly outperforms its single-IRS counterpart as the power and the total number of IRS elements increase. However, the single-IRS aided system remains optimal under the constraints of lower power and fewer total IRS elements.

In Fig. 4, we further investigate the impact of IRS elements number on the overall system performance. We employ the number of the effective rank, as introduced in [10], which is defined as Erank(𝐇)=exp(kδ¯klnδ¯k)\mathrm{Erank}\left(\mathbf{H}\right)=\exp\left(-\sum_{k}{\bar{\delta}_{k}\ln\bar{\delta}_{k}}\right), where δ¯k=δk/iδi\bar{\delta}_{k}=\sqrt{\delta_{k}}/\sum_{i}{\sqrt{\delta_{i}}} and δk\delta_{k} is the kkth singular value of 𝐇\mathbf{H}. Fig. 4 illustrates the variation of effective rank with the total number of IRS elements when P=30P=30 dBm. It is clear that as the number of IRS elements increases, multiple IRSs offer greater potential for rank enhancement compared to a single IRS. The rank improvement reflects the system’s spatial multiplexing capability. In particular, enhanced spatial multiplexing allows multiple IRSs to achieve higher gains when the total number of IRS elements is large.

Refer to caption
Figure 2: Achievable rate versus MM under orthogonal placement.
Refer to caption
Figure 3: Achievable rate versus PP under orthogonal placement.
Refer to caption
Figure 4: Effective rank versus MM.
Refer to caption
Figure 5: Elements allocation versus PP when K=3K=3.

Fig. 5 provides a more detailed view of the IRS elements allocation process for different total numbers of elements MM when K=3K=3. In Fig. 5, M1M_{1}, M2M_{2}, and M3M_{3} denote the number of elements allocated to three IRSs, where M1M2M3M_{1}\geqslant M_{2}\geqslant M_{3} and k=13Mk=M\sum_{k=1}^{3}{M_{k}}=M. As the power increases, the IRS with the largest number of elements will gradually allocate its elements to other IRSs and their elements distribution across individual IRSs tends towards equality, corresponding to our proposition in Section IV.

VI Conclusion

In this work, we investigate the problem of the IRS placement and resource allocation in MIMO communication systems. We propose an orthogonal placement scheme to maximize the spatial multiplexing gain, upon which we optimize IRS beamforming as well as elements and power allocation to maximize the SE. Moreover, we analytically characterize the system’s SE scaling orders with respect to the number of reflecting elements and power. Our numerical results demonstrate that, when the total number of IRS elements or the power exceeds a certain threshold, multi-IRS systems significantly outperform single-IRS systems and equal distribution of elements and power across multiple IRSs is shown to be asymptotically optimal.

References

  • [1] A. Goldsmith, S. A. Jafar, N. Jindal, and S. Vishwanath, “Capacity limits of MIMO channels,” IEEE J. Sel. Areas Commun., vol. 21, no. 5, pp. 684–702, Jun. 2003.
  • [2] Q. Wu, X. Guan, and R. Zhang, “Intelligent reflecting surface-aided wireless energy and information transmission: An overview,” Proceedings of the IEEE, vol. 110, no. 1, pp. 150–170, Nov. 2022.
  • [3] G. Chen, Q. Wu, R. Liu, J. Wu, and C. Fang, “IRS aided MEC systems with binary offloading: A unified framework for dynamic IRS beamforming,” IEEE J. Sel. Areas Commun., vol. 41, no. 2, pp. 349–365, Feb. 2023.
  • [4] G. Chen and Q. Wu, “Fundamental limits of intelligent reflecting surface aided multiuser broadcast channel,” IEEE Transactions on Communications, vol. 71, no. 10, pp. 5904–5919, Oct. 2023.
  • [5] Q. Wu, S. Zhang, B. Zheng, C. You, and R. Zhang, “Intelligent reflecting surface-aided wireless communications: A tutorial,” IEEE Trans. Commun., vol. 69, no. 5, pp. 3313–3351, May 2021.
  • [6] Y. Han, S. Zhang, L. Duan, and R. Zhang, “Double-IRS aided MIMO communication under LoS channels: Capacity maximization and scaling,” IEEE Trans. Commun., vol. 70, no. 4, pp. 2820–2837, Apr. 2022.
  • [7] W. Chen, C.-K. Wen, X. Li, and S. Jin, “Channel customization for joint Tx-RISs-Rx design in hybrid mmWave systems,” IEEE Trans. Wireless Commun., vol. 22, no. 11, pp. 8304–8319, Nov. 2023.
  • [8] G. Chen, Q. Wu, W. Chen, Y. Hou, M. Jian, S. Zhang, J. Li, and L. Hanzo, “Intelligent reflecting surface aided MIMO networks: Distributed or centralized architecture?” [Online] Available: https://arxiv.org/pdf/2310.01742.
  • [9] S. Boyd and L. Vandenberghe, Convex Optimization, Cambridge, U.K.:Cambridge Univ. Press, 2004.
  • [10] O. Roy and M. Vetterli, “The effective rank: A measure of effective dimensionality,” in 2007 15th European Signal Processing Conf., May 2007, pp. 606–610.