License: CC BY 4.0
arXiv:2311.12628v2 [cs.IT] 01 Dec 2023

Empirical Validation of the Impedance-Based RIS Channel Model in an Indoor Scattering Environment

Placido Mursia1, Taghrid Mazloum2, Frédéric Munoz2, Vincenzo Sciancalepore1, Gabriele Gradoni3,
Raffaele D’Errico2, Marco Di Renzo4, Xavier Costa-Pérez15, Antonio Clemente2, Geoffroy Lerosey6
1NEC Laboratories Europe GmbH, Heidelberg, Germany, Email: [email protected] 2CEA-LETI, Univ. Grenoble Alpes, Grenoble, France 3University of Surrey, Guildford, United Kingdom 4Université Paris-Saclay, CNRS, Centrale Supélec, Gif-sur-Yvette, France 5i2CAT Foundation and ICREA, Barcelona, Spain 6Greenerwave, Paris, France
Abstract

Ensuring the precision of channel modeling plays a pivotal role in the development of wireless communication systems, and this requirement remains a persistent challenge within the realm of networks supported by Reconfigurable Intelligent Surfaces (RIS). Achieving a comprehensive and reliable understanding of channel behavior in RIS-aided networks is an ongoing and complex issue that demands further exploration. In this paper, we empirically validate a recently-proposed impedance-based RIS channel model that accounts for the mutual coupling at the antenna array and precisely models the presence of scattering objects within the environment as a discrete array of loaded dipoles. To this end, we exploit real-life channel measurements collected in an office environment to demonstrate the validity of such a model and its applicability in a practical scenario. Finally, we provide numerical results demonstrating that designing the RIS configuration based upon such model leads to superior performance as compared to reference schemes.

Index Terms:
RIS, mutual coupling, antennas, electromagnetics, propagation, measurements.

I Introduction

Reconfigurable intelligent surfaces (RISs) are widely considered as one of the ongoing revolutions in the network design due to their property to programmatically alter the propagation properties of the radio environment while keeping the overall manufacturing cost affordable. As such, RISs are a candidate technology for the next-generation wireless networks [1]. In this regard, one of the main research topics has been on exploring and further developing accurate channel modelling, which is essential to fully unlock the potentials of this technology.

Recently, the authors in [2] have proposed a mutual coupling and unit cell-aware electromagnetically-consistent channel model based on mutual impedances. Therein, the unit cells of the RIS are modelled as arbitrarily-spaced wire dipoles, whose loads can be tuned to alter the propagation conditions. Such a model has been exploited in [3] to derive the optimal RIS configuration that maximizes the received power in a single-user and single-antenna setup. Moreover, the authors in [4] have generalized this optimization framework for the case of multi-input multiple-output (MIMO) systems. Therein, the multipath components originated by scattering objects in the environment are modelled as an additive statistical component. On the other hand, the authors of [5] exploit the discrete dipole approximation (DDA) [6] to model the scattering objects present in the environment as loaded wire dipoles, thus effectively capturing its interaction with the transmitted signals. Under this setting, the authors of [7] developed a provably convergent and nearly-optimal RIS optimization algorithm based on Gram-Schmidt’s orthogonalization method.

In this paper, we go one-step beyond and provide an empirical validation of the model in [5] by showcasing how its parameters can be tuned to effectively recreate a real-life environment. To this purpose, we exploit a set of channel measurements [8] performed in an office environment wherein a reflective RIS operating at 28282828 GHz [9] aids in the communication between a transmitter [10] and a single-antenna user equipment (UE), in the presence of several scattering objects. Additionally, we demonstrate how the considered RIS channel model effectively captures such physical propagation scenario with few parameters. Finally, numerical results demonstrate that performing the RIS optimization using such a tailored channel model can bring significant gains in terms of received power at the UE.

Notation. We use bold font lower and upper case for vectors and matrices, respectively. ()HsuperscriptH(\cdot)^{\mathrm{H}}( ⋅ ) start_POSTSUPERSCRIPT roman_H end_POSTSUPERSCRIPT denotes the hermitian transpose operator, while j=1𝑗1j=\sqrt{-1}italic_j = square-root start_ARG - 1 end_ARG is the imaginary number.

II System model

We consider the scenario depicted in Fig. 1, wherein a transmitter (TX) equipped with M=400𝑀400M=400italic_M = 400 antenna elements is located in the origin of the reference system [10], an RIS equipped with N=1600𝑁1600N=1600italic_N = 1600 elements is placed (i.e., its center-point) at coordinates [230]matrix230\begin{bmatrix}-2&-3&0\end{bmatrix}[ start_ARG start_ROW start_CELL - 2 end_CELL start_CELL - 3 end_CELL start_CELL 0 end_CELL end_ROW end_ARG ] m [9], a single-antenna UE is placed at coordinates [13.60]matrix13.60\begin{bmatrix}1&-3.6&0\end{bmatrix}[ start_ARG start_ROW start_CELL 1 end_CELL start_CELL - 3.6 end_CELL start_CELL 0 end_CELL end_ROW end_ARG ] m, and a number of scattering objects are present in the environment.Moreover, we employ the impedance-based RIS channel model in [5], such that the scattering objects in the environment are modelled as discrete arrays of half-wavelength loaded dipoles. Hence, the 1×M1𝑀1\times M1 × italic_M channel vector is given by

𝐡H=ZRL[𝐳ROT𝐳ROS(𝐙SS+𝐙SOS+𝐙RIS)1𝐙SOT]𝐙TG,superscript𝐡Hsubscript𝑍RLdelimited-[]subscript𝐳ROTsubscript𝐳ROSsuperscriptsubscript𝐙SSsubscript𝐙SOSsubscript𝐙RIS1subscript𝐙SOTsubscript𝐙TG\displaystyle\mathbf{h}^{\mathrm{H}}=Z_{\scriptscriptstyle\mathrm{RL}}\Big{[}% \mathbf{z}_{\scriptscriptstyle\mathrm{ROT}}-\mathbf{z}_{\scriptscriptstyle% \mathrm{ROS}}\Big{(}\mathbf{Z}_{\scriptscriptstyle\mathrm{SS}}+\mathbf{Z}_{% \scriptscriptstyle\mathrm{SOS}}+\mathbf{Z}_{\scriptscriptstyle\mathrm{RIS}}% \Big{)}^{-1}\mathbf{Z}_{\scriptscriptstyle\mathrm{SOT}}\Big{]}\mathbf{Z}_{% \scriptscriptstyle\mathrm{TG}},bold_h start_POSTSUPERSCRIPT roman_H end_POSTSUPERSCRIPT = italic_Z start_POSTSUBSCRIPT roman_RL end_POSTSUBSCRIPT [ bold_z start_POSTSUBSCRIPT roman_ROT end_POSTSUBSCRIPT - bold_z start_POSTSUBSCRIPT roman_ROS end_POSTSUBSCRIPT ( bold_Z start_POSTSUBSCRIPT roman_SS end_POSTSUBSCRIPT + bold_Z start_POSTSUBSCRIPT roman_SOS end_POSTSUBSCRIPT + bold_Z start_POSTSUBSCRIPT roman_RIS end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT bold_Z start_POSTSUBSCRIPT roman_SOT end_POSTSUBSCRIPT ] bold_Z start_POSTSUBSCRIPT roman_TG end_POSTSUBSCRIPT , (1)

where ZRLsubscript𝑍RLZ_{\scriptscriptstyle\mathrm{RL}}\in\mbox{$\mathbb{C}$}italic_Z start_POSTSUBSCRIPT roman_RL end_POSTSUBSCRIPT ∈ blackboard_C (𝐙TGM×Msubscript𝐙TGsuperscript𝑀𝑀\mathbf{Z}_{\scriptscriptstyle\mathrm{TG}}\in\mbox{$\mathbb{C}$}^{M\times M}bold_Z start_POSTSUBSCRIPT roman_TG end_POSTSUBSCRIPT ∈ blackboard_C start_POSTSUPERSCRIPT italic_M × italic_M end_POSTSUPERSCRIPT) accounts for the load and self-impedance at the UE (transmitter); 𝐳ROT1×Msubscript𝐳ROTsuperscript1𝑀\mathbf{z}_{\scriptscriptstyle\mathrm{ROT}}\in\mbox{$\mathbb{C}$}^{1\times M}bold_z start_POSTSUBSCRIPT roman_ROT end_POSTSUBSCRIPT ∈ blackboard_C start_POSTSUPERSCRIPT 1 × italic_M end_POSTSUPERSCRIPT includes the mutual impedances between the transmitter and the UE, both directly and indirectly through the scattering objects; 𝐳ROS1×Nsubscript𝐳ROSsuperscript1𝑁\mathbf{z}_{\scriptscriptstyle\mathrm{ROS}}\in\mbox{$\mathbb{C}$}^{1\times N}bold_z start_POSTSUBSCRIPT roman_ROS end_POSTSUBSCRIPT ∈ blackboard_C start_POSTSUPERSCRIPT 1 × italic_N end_POSTSUPERSCRIPT (𝐙SOTN×Msubscript𝐙SOTsuperscript𝑁𝑀\mathbf{Z}_{\scriptscriptstyle\mathrm{SOT}}\in\mbox{$\mathbb{C}$}^{N\times M}bold_Z start_POSTSUBSCRIPT roman_SOT end_POSTSUBSCRIPT ∈ blackboard_C start_POSTSUPERSCRIPT italic_N × italic_M end_POSTSUPERSCRIPT) models the mutual impedance between the RIS and the UE (transmitter), including the effect due to the scattering objects; 𝐙SSN×Nsubscript𝐙SSsuperscript𝑁𝑁\mathbf{Z}_{\scriptscriptstyle\mathrm{SS}}\in\mbox{$\mathbb{C}$}^{N\times N}bold_Z start_POSTSUBSCRIPT roman_SS end_POSTSUBSCRIPT ∈ blackboard_C start_POSTSUPERSCRIPT italic_N × italic_N end_POSTSUPERSCRIPT and 𝐙SOSN×Nsubscript𝐙SOSsuperscript𝑁𝑁\mathbf{Z}_{\scriptscriptstyle\mathrm{SOS}}\in\mbox{$\mathbb{C}$}^{N\times N}bold_Z start_POSTSUBSCRIPT roman_SOS end_POSTSUBSCRIPT ∈ blackboard_C start_POSTSUPERSCRIPT italic_N × italic_N end_POSTSUPERSCRIPT represent the self-impedance among the RIS elements and the self and mutual impedance among the scattering objects and the RIS, respectively; lastly, 𝐙RISN×Nsubscript𝐙RISsuperscript𝑁𝑁\mathbf{Z}_{\scriptscriptstyle\mathrm{RIS}}\in\mbox{$\mathbb{C}$}^{N\times N}bold_Z start_POSTSUBSCRIPT roman_RIS end_POSTSUBSCRIPT ∈ blackboard_C start_POSTSUPERSCRIPT italic_N × italic_N end_POSTSUPERSCRIPT stands for the diagonal matrix of tunable RIS impedances modelled as 𝐙RIS=diag(R0+jx),=1,,Nformulae-sequencesubscript𝐙RISdiagsubscript𝑅0𝑗subscript𝑥1𝑁\mathbf{Z}_{\scriptscriptstyle\mathrm{RIS}}=\mathrm{diag}(R_{0}+jx_{\ell}),% \quad\ell=1,\ldots,Nbold_Z start_POSTSUBSCRIPT roman_RIS end_POSTSUBSCRIPT = roman_diag ( italic_R start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT + italic_j italic_x start_POSTSUBSCRIPT roman_ℓ end_POSTSUBSCRIPT ) , roman_ℓ = 1 , … , italic_N, where R00subscript𝑅00R_{0}\geq 0italic_R start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ≥ 0 is the constant resistance of each RIS load and xsubscript𝑥x_{\ell}\in\mbox{$\mathbb{R}$}italic_x start_POSTSUBSCRIPT roman_ℓ end_POSTSUBSCRIPT ∈ blackboard_R is the tunable reactance at the \ellroman_ℓ-th element. The precoding vector at the transmitter is denoted by 𝐰M×1𝐰superscript𝑀1\mathbf{w}\in\mbox{$\mathbb{C}$}^{M\times 1}bold_w ∈ blackboard_C start_POSTSUPERSCRIPT italic_M × 1 end_POSTSUPERSCRIPT. As depicted in Fig. 1, such vector is defined in order to produce a beam pointing at 35superscript35-35^{\circ}- 35 start_POSTSUPERSCRIPT ∘ end_POSTSUPERSCRIPT in azimuth and 0superscript00^{\circ}0 start_POSTSUPERSCRIPT ∘ end_POSTSUPERSCRIPT in elevation. Specifically, its beampattern is given in [10]. We remark that the considered measurements collected in [8] involve arrays of patch antennas with a uniform gain of 6666 dBi, while the adopted RIS channel model in [5] assumes dipole antennas. Hence, to compensate for the reduced antenna gain, we design the precoder vector such that 𝐰2MGsuperscriptnorm𝐰2𝑀𝐺\|\mathbf{w}\|^{2}\leq M\,G∥ bold_w ∥ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ≤ italic_M italic_G, with G=6𝐺6G=6italic_G = 6 dBi. Moreover, in order to reproduce the setting in [8], we fix the RIS configuration in 𝐙RISsubscript𝐙RIS\mathbf{Z}_{\scriptscriptstyle\mathrm{RIS}}bold_Z start_POSTSUBSCRIPT roman_RIS end_POSTSUBSCRIPT by feeding the channel vector in (1) in the absence of both the scattered paths and the direct link, to the SARIS algorithm in [5] (see Section III-B2 for details). The resulting beam at the RIS exhibits a main lobe pointing at 10superscript10-10^{\circ}- 10 start_POSTSUPERSCRIPT ∘ end_POSTSUPERSCRIPT in azimuth and 0superscript00^{\circ}0 start_POSTSUPERSCRIPT ∘ end_POSTSUPERSCRIPT in elevation. As previously done for the precoder vector, we then compensate for the patch antenna gain at the RIS by multiplying the channel coefficient corresponding to the path reflected upon the RIS by G𝐺\sqrt{G}square-root start_ARG italic_G end_ARG. Lastly, the receive signal at the UE is thus given by

y=𝐡H𝐰x+n,𝑦superscript𝐡H𝐰𝑥𝑛\displaystyle y=\mathbf{h}^{\mathrm{H}}\mathbf{w}x+n\in\mbox{$\mathbb{C}$},italic_y = bold_h start_POSTSUPERSCRIPT roman_H end_POSTSUPERSCRIPT bold_w italic_x + italic_n ∈ blackboard_C , (2)

where x𝑥x\in\mbox{$\mathbb{C}$}italic_x ∈ blackboard_C is the (known) transmit signal, and n𝑛nitalic_n is the additive white Gaussian noise coefficient.

Refer to caption
Figure 1: System model

III Empirical validation procedure

In this section, we firstly describe the considered measurements collected in the real-life environment described in Fig. 1. Subsequently, focusing on the channel model in (1), we describe the procedure to fit the associated relevant parameters in order to match the underlying physical propagation environment.

III-A Channel measurements

In the indoor environment depicted in Fig. 1, the propagation channel was measured using a vector network analyzer (VNA) operating within the frequency range of 25252525-35353535 GHz. The setup included the use of an RIS as an extender, a TX, and a wideband monopole antenna serving as UE. All of these terminals were placed at a height of about 1.61.61.61.6 m above the ground, corresponding to the azimuth plane. Further details on the measurement setup and the analysis of the channel characteristics are found in [8]. The TX was configured to steer the beam towards the RIS, which, in turn, was configured to redirect the received beam towards the intended UE. The UE, placed on a 2D positioner, was moved on a 3×3333\times 33 × 3 spatial grid of positions, with a step equivalent to half a wavelength, thus emulating a 3×3333\times 33 × 3 antenna array. Across a bandwidth of 2222 GHz, this spatial grid allows the extraction of the multipath components (MPCs) in both the time delay (or alternatively the distance) and the angular domains, where the latter corresponds specifically to the azimuth angle-of-arrival (AoA) at the UE-side. This is achieved by applying a wideband high resolution algorithm such as the Space-Alternating Generalized Expectation-Maximization algorithm (SAGE) [11].

The TX and the RIS operate within the Ka band at the resonance frequency of 28282828 GHz. The linearly-polarized TX [10] consists of 20×20202020\times 2020 × 20 unit cells (UCs) and is illuminated with a 10101010-dBi horn antenna. The employed RIS [9] is an assembly of four rectangular lattices of 20×20202020\times 2020 × 20 UCs each, resulting on a total of 40×40404040\times 4040 × 40 UCs. The one-bit UCs of the TX and the RIS are arranged at regular intervals of half a wavelength and the configuration is realized through electronic control of integrated PIN diodes. This enables ON and OFF states with a relative phase-shift of approximately 180superscript180180^{\circ}180 start_POSTSUPERSCRIPT ∘ end_POSTSUPERSCRIPT at the chosen operating frequency. Further details on the 1111-bit UC architecture and performance are provided in [10] for the TX and in [9] for the RIS.

The heatmap of the MPCs as perceived by the UE, extracted using the SAGE algorithm, is illustrated in Fig. 2, where each path is graphically represented by a point on the angle-distance axes, and its color is related to its amplitude (in dB). The identification of the MPCs on the heatmap is simplified with geometrical calculations based on the dimension of the floor plan. Accordingly, the MPC reflected by the RIS appears at 7.27.27.27.2 m with an AoA of 76superscript7676^{\circ}76 start_POSTSUPERSCRIPT ∘ end_POSTSUPERSCRIPT, and the direct TX to UE path is located at 4444 m with an AoA of 14superscript1414^{\circ}14 start_POSTSUPERSCRIPT ∘ end_POSTSUPERSCRIPT. Their powers are respectively 72.472.4-72.4- 72.4 dB and 68.768.7-68.7- 68.7 dB. Also, the strongest scattered path due to the environmental objects is represented by the one corresponding to a distance of 5.695.695.695.69 m, AoA of 50superscript50-50^{\circ}- 50 start_POSTSUPERSCRIPT ∘ end_POSTSUPERSCRIPT and a power of 79.879.8-79.8- 79.8 dB. All other MPCs are due to scatterers that are far-away from the UE or to measurement noise. In this paper for model validation we limit our analysis to the main considered MPCs summarized in Table I.

TABLE I: Main MPCs in terms of distance travelled, AoA at the UE and measured power.
MPC Distance AoA Power
Direct path (TX/UE LoS) 4444 m 14superscript1414^{\circ}14 start_POSTSUPERSCRIPT ∘ end_POSTSUPERSCRIPT 68.768.7-68.7- 68.7 dB
Reflected path (TX/RIS/UE) 7.27.27.27.2 m 76superscript7676^{\circ}76 start_POSTSUPERSCRIPT ∘ end_POSTSUPERSCRIPT 72.472.4-72.4- 72.4 dB
Secondary Cluster 5.695.695.695.69 m 50superscript50-50^{\circ}- 50 start_POSTSUPERSCRIPT ∘ end_POSTSUPERSCRIPT 79.879.8-79.8- 79.8 dB

III-B Channel model fitting

In order to tune the parameters in the model described in Section II and replicate the setup in [8], we distinguish three cases, namely i) the direct path from the transmitter to the receiver i.e. the TX/UE Line of Sight (LoS) path, ii) the reflected path from the transmitter to the UE via the RIS, i.e. the TX/RIS/UE path, and iii) the scattered path from the transmitter to the receiver via a cluster of reflectors in the vicinity of the UE.

III-B1 Direct path

Regarding the direct path from the transmitter to the UE, we modify the channel vector as

𝐡dH=ZRL𝐳RT𝐙TG,superscriptsubscript𝐡𝑑Hsubscript𝑍RLsubscript𝐳RTsubscript𝐙TG\displaystyle\mathbf{h}_{d}^{\mathrm{H}}=Z_{\scriptscriptstyle\mathrm{RL}}% \mathbf{z}_{\scriptscriptstyle\mathrm{RT}}\mathbf{Z}_{\scriptscriptstyle% \mathrm{TG}},bold_h start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT start_POSTSUPERSCRIPT roman_H end_POSTSUPERSCRIPT = italic_Z start_POSTSUBSCRIPT roman_RL end_POSTSUBSCRIPT bold_z start_POSTSUBSCRIPT roman_RT end_POSTSUBSCRIPT bold_Z start_POSTSUBSCRIPT roman_TG end_POSTSUBSCRIPT , (3)

where 𝐳RTsubscript𝐳RT\mathbf{z}_{\scriptscriptstyle\mathrm{RT}}bold_z start_POSTSUBSCRIPT roman_RT end_POSTSUBSCRIPT is the mutual impedance between the transmitter and the UE. As stated above, the transmitter configuration is set as described in [10].

III-B2 Reflected path

In this case, we consider only the path that reaches the UE from the transmitter via the RIS. Hence, we modify the channel vector in (1) as

𝐡RISH=ZRL[𝐳RS(𝐙SS+𝐙RIS)1𝐙ST]𝐙TG,superscriptsubscript𝐡RISHsubscript𝑍RLdelimited-[]subscript𝐳RSsuperscriptsubscript𝐙SSsubscript𝐙RIS1subscript𝐙STsubscript𝐙TG\displaystyle\mathbf{h}_{\scriptscriptstyle\mathrm{RIS}}^{\mathrm{H}}=-Z_{% \scriptscriptstyle\mathrm{RL}}\Big{[}\mathbf{z}_{\scriptscriptstyle\mathrm{RS}% }\Big{(}\mathbf{Z}_{\scriptscriptstyle\mathrm{SS}}+\mathbf{Z}_{% \scriptscriptstyle\mathrm{RIS}}\Big{)}^{-1}\mathbf{Z}_{\scriptscriptstyle% \mathrm{ST}}\Big{]}\mathbf{Z}_{\scriptscriptstyle\mathrm{TG}},bold_h start_POSTSUBSCRIPT roman_RIS end_POSTSUBSCRIPT start_POSTSUPERSCRIPT roman_H end_POSTSUPERSCRIPT = - italic_Z start_POSTSUBSCRIPT roman_RL end_POSTSUBSCRIPT [ bold_z start_POSTSUBSCRIPT roman_RS end_POSTSUBSCRIPT ( bold_Z start_POSTSUBSCRIPT roman_SS end_POSTSUBSCRIPT + bold_Z start_POSTSUBSCRIPT roman_RIS end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT bold_Z start_POSTSUBSCRIPT roman_ST end_POSTSUBSCRIPT ] bold_Z start_POSTSUBSCRIPT roman_TG end_POSTSUBSCRIPT , (4)

with 𝐳RSsubscript𝐳RS\mathbf{z}_{\scriptscriptstyle\mathrm{RS}}bold_z start_POSTSUBSCRIPT roman_RS end_POSTSUBSCRIPT and 𝐙STsubscript𝐙ST\mathbf{Z}_{\scriptscriptstyle\mathrm{ST}}bold_Z start_POSTSUBSCRIPT roman_ST end_POSTSUBSCRIPT the mutual impedance between the receiver and the RIS and between the latter and the transmitter in the absence of scattering objects, respectively. Here, all parameters can be computed given the (fixed) location of the transmitter, the RIS and the receiver [2]. The TX configurations is set as described in Section II, while the RIS configuration is obtained by feeding (4) to the SARIS algorithm [5].

III-B3 Secondary cluster

Lastly, we focus on the path that reaches the UE from the TX via a cluster of scatterers, which is located in the vicinity of the UE. In this case, the channel vector is given by

𝐡refH=ZRL𝐳RO(𝐙OO+𝐙US)1𝐳RO𝐙TG.superscriptsubscript𝐡𝑟𝑒𝑓Hsubscript𝑍RLsubscript𝐳ROsuperscriptsubscript𝐙OOsubscript𝐙US1subscript𝐳ROsubscript𝐙TG\displaystyle\mathbf{h}_{ref}^{\mathrm{H}}=-Z_{\scriptscriptstyle\mathrm{RL}}% \mathbf{z}_{\scriptscriptstyle\mathrm{RO}}(\mathbf{Z}_{\scriptscriptstyle% \mathrm{OO}}+\mathbf{Z}_{\scriptscriptstyle\mathrm{US}})^{-1}\mathbf{z}_{% \scriptscriptstyle\mathrm{RO}}\mathbf{Z}_{\scriptscriptstyle\mathrm{TG}}.bold_h start_POSTSUBSCRIPT italic_r italic_e italic_f end_POSTSUBSCRIPT start_POSTSUPERSCRIPT roman_H end_POSTSUPERSCRIPT = - italic_Z start_POSTSUBSCRIPT roman_RL end_POSTSUBSCRIPT bold_z start_POSTSUBSCRIPT roman_RO end_POSTSUBSCRIPT ( bold_Z start_POSTSUBSCRIPT roman_OO end_POSTSUBSCRIPT + bold_Z start_POSTSUBSCRIPT roman_US end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT bold_z start_POSTSUBSCRIPT roman_RO end_POSTSUBSCRIPT bold_Z start_POSTSUBSCRIPT roman_TG end_POSTSUBSCRIPT . (5)

As described in [5], 𝐳ROsubscript𝐳RO\mathbf{z}_{\scriptscriptstyle\mathrm{RO}}bold_z start_POSTSUBSCRIPT roman_RO end_POSTSUBSCRIPT represents the mutual impedance between the UE and the cluster of scatterers, 𝐙OOsubscript𝐙OO\mathbf{Z}_{\scriptscriptstyle\mathrm{OO}}bold_Z start_POSTSUBSCRIPT roman_OO end_POSTSUBSCRIPT the self and mutual impedances among the scatterers, 𝐙USsubscript𝐙US\mathbf{Z}_{\scriptscriptstyle\mathrm{US}}bold_Z start_POSTSUBSCRIPT roman_US end_POSTSUBSCRIPT the diagonal matrix modelling the load attached to each scattering dipole, and 𝐳ROsubscript𝐳RO\mathbf{z}_{\scriptscriptstyle\mathrm{RO}}bold_z start_POSTSUBSCRIPT roman_RO end_POSTSUBSCRIPT stands for the mutual impedance between the cluster of scatterers and the TX. Note that the dimension of such quantities depends on Ncsubscript𝑁𝑐N_{c}italic_N start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT, i.e., the number of scattering dipoles, which needs to be fitted to accturately represent the cluster. In particular, based upon the floor plan of the channel measurements described in Section III-A, such cluster consists of a cylindrical pillar with a width of approximately 1111 m and a height of approximately 3333 m. However, given the narrow shape in 3D of the beam pattern at the transmitter, it is sufficient to model only a small section of 16×816816\times 816 × 8 cm. Specifically, by applying the DDA [6], we model such object as a cylindrical array of radius 0.50.50.50.5 m, consisting of Nc=Nx×Nysubscript𝑁𝑐subscript𝑁𝑥subscript𝑁𝑦N_{c}=N_{x}\times N_{y}italic_N start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT = italic_N start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT × italic_N start_POSTSUBSCRIPT italic_y end_POSTSUBSCRIPT loaded dipoles of length h=λ/4𝜆4h=\lambda/4italic_h = italic_λ / 4 whose inter-element distance is given by d𝑑ditalic_d, with Nxsubscript𝑁𝑥N_{x}italic_N start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT the number of dipoles on the horizontal plane, Nysubscript𝑁𝑦N_{y}italic_N start_POSTSUBSCRIPT italic_y end_POSTSUBSCRIPT the number of dipoles on the vertical plane, and λ𝜆\lambdaitalic_λ the signal wavelength. The load on each t𝑡titalic_t-th dipole is given by ZUSt=y0+jxtsubscript𝑍US𝑡subscript𝑦0𝑗subscript𝑥𝑡Z_{\scriptscriptstyle\mathrm{US}t}=y_{0}+jx_{t}italic_Z start_POSTSUBSCRIPT roman_US italic_t end_POSTSUBSCRIPT = italic_y start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT + italic_j italic_x start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT, where y0subscript𝑦0y_{0}\in\mbox{$\mathbb{R}$}italic_y start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ∈ blackboard_R is constant over the array, and it depends on the specific absorption and reflection properties of the material, while xtsubscript𝑥𝑡x_{t}\in\mbox{$\mathbb{R}$}italic_x start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT ∈ blackboard_R models the scattering properties of the object. To this end, the values of Nxsubscript𝑁𝑥N_{x}italic_N start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT, Nysubscript𝑁𝑦N_{y}italic_N start_POSTSUBSCRIPT italic_y end_POSTSUBSCRIPT, d𝑑ditalic_d, {xt}t=1Nx×Nysuperscriptsubscriptsubscript𝑥𝑡𝑡1subscript𝑁𝑥subscript𝑁𝑦\{x_{t}\}_{t=1}^{N_{x}\times N_{y}}{ italic_x start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT } start_POSTSUBSCRIPT italic_t = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_N start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT × italic_N start_POSTSUBSCRIPT italic_y end_POSTSUBSCRIPT end_POSTSUPERSCRIPT, as well as the location of the object need to be fitted to reproduce the results in Table I.

IV Numerical results and discussion

In order to validate the effectiveness of the channel model in (1) and [5], we evaluate the (normalized) power received at the UE-side in dB, which is given by

P=10log10(|𝐡H𝐰|2),𝑃10subscript10superscriptsuperscript𝐡H𝐰2\displaystyle P=10\log_{10}(|\mathbf{h}^{\mathrm{H}}\mathbf{w}|^{2}),italic_P = 10 roman_log start_POSTSUBSCRIPT 10 end_POSTSUBSCRIPT ( | bold_h start_POSTSUPERSCRIPT roman_H end_POSTSUPERSCRIPT bold_w | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) , (6)

in the various cases highlighted in Section III-B. The obtained results for the reflected and direct paths are shown in Table II, demonstrating an error of just 1.16%percent1.161.16\%1.16 % and 0.35%percent0.350.35\%0.35 % for the two cases, respectively. For the case of the path including the cylindrical reflector, we generate 103superscript10310^{3}10 start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT different combinations of the reactance xtsubscript𝑥𝑡x_{t}italic_x start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT for the load of each t𝑡titalic_t-th dipole in the cylindrical array by drawing values from the uniform distribution 𝒰[330,100]𝒰330100\mathcal{U}[-330,100]caligraphic_U [ - 330 , 100 ]. Furthermore, we vary the total number of dipoles Nx×Nysubscript𝑁𝑥subscript𝑁𝑦N_{x}\times N_{y}italic_N start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT × italic_N start_POSTSUBSCRIPT italic_y end_POSTSUBSCRIPT with fixed ratio Ny=2Nxsubscript𝑁𝑦2subscript𝑁𝑥N_{y}=2N_{x}italic_N start_POSTSUBSCRIPT italic_y end_POSTSUBSCRIPT = 2 italic_N start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT, the inter-element distance d𝑑ditalic_d, and the real part of the dipole loads y0subscript𝑦0y_{0}italic_y start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT. For each combination, we select the maximum value obtained over the random reactance.

In Fig. 3, we show the receive power at the UE versus the total number of dipoles for the case of d=λ/4𝑑𝜆4d=\lambda/4italic_d = italic_λ / 4, and for different values of y0subscript𝑦0y_{0}italic_y start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT. We thus conclude that approximately 2450245024502450 dipoles, corresponding to a 35×70357035\times 7035 × 70 configuration, with y0=100Ωsubscript𝑦0100Ωy_{0}=100~{}\Omegaitalic_y start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = 100 roman_Ω is a precise enough fitting of the measured data, resulting in an error of 0.14%percent0.140.14\%0.14 %. Moreover, in Table III, we provide the simulated received power for the case of a 35×70357035\times 7035 × 70 cylindrical array with y0=100Ωsubscript𝑦0100Ωy_{0}=100~{}\Omegaitalic_y start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = 100 roman_Ω for different values of the inter-element d𝑑ditalic_d, thus confirming that a dense array accurately represents the physical propagation properties of a real-life object.

While so far we have successfully emulated the measurements in [8] by configuring the RIS as described in Section II, in the following we demonstrate how the algorithm in [5] can lead to superior performance by designing an RIS configuration that explicitly interacts with the scattering environment and creates constructive interference at the UE. In this regard, we set all channel parameters according to the values that provide the best fitting of the physical propagation environment, as described above, and compare the RIS configuration used during the measurements described in Section III-A versus the SARIS algorithm in [5]. As shown in Table IV, the proposed RIS optimization algorithm achieves significant gains in the received power thanks to an accurate modelling and exploitation of the physical properties of the channel.

TABLE II: Simulated MPCs for the reflected and direct paths versus the measured data.
MPC Measured power Simulated power
Direct path 68.768.7-68.7- 68.7 dB 68.568.5-68.5- 68.5 dB
Reflected path 72.472.4-72.4- 72.4 dB 73.373.3-73.3- 73.3 dB
Refer to caption
Figure 2: Heatmap of the time delay (or alternatively the distance) and azimuth AoA for the multipath in the considered scenario as perceived by the UE.
Refer to caption
Figure 3: Received power versus the number of dipoles in the cylindrical array with inter-element spacing of d=λ/4𝑑𝜆4d=\lambda/4italic_d = italic_λ / 4, and for different values of y0subscript𝑦0y_{0}italic_y start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT.
TABLE III: Simulated MPC involving the close-by reflector versus different inter-element spacings d𝑑ditalic_d in a 35×70357035\times 7035 × 70 cylindrical array and with y0=100Ωsubscript𝑦0100Ωy_{0}=100~{}\Omegaitalic_y start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = 100 roman_Ω.
Simulated power Inter-element spacing d𝑑ditalic_d Simulated power Inter-element spacing d𝑑ditalic_d
113.85113.85-113.85- 113.85 dB 5λ5𝜆5\lambda5 italic_λ 89.4589.45-89.45- 89.45 dB λ𝜆\lambdaitalic_λ
86.6486.64-86.64- 86.64 dB 0.5λ0.5𝜆0.5\lambda0.5 italic_λ 79.9379.93-79.93- 79.93 dB 0.25λ0.25𝜆0.25\lambda0.25 italic_λ
TABLE IV: Received power in the considered scenario under different RIS configurations.
RIS configuration Simulated power
10superscript10-10^{\circ}- 10 start_POSTSUPERSCRIPT ∘ end_POSTSUPERSCRIPT azimuth, 0superscript00^{\circ}0 start_POSTSUPERSCRIPT ∘ end_POSTSUPERSCRIPT elevation 63.2763.27-63.27- 63.27 dB
SARIS [5] 52.3452.34-52.34- 52.34 dB

V Conclusion

In this paper, we shed experimentally validate a recently-introduced electromagnetic-compliant channel model for RIS, which takes into account mutual coupling effects and considers the presence of scattering objects. To this end, we have validated its accuracy in representing the physical characteristics of real-world scenarios. Interestingly, we have fine-tuned the parameters of model using a dataset of channel measurements collected from an office environment. By recreating the channel propagation with the DDA, we have demonstrated that the proposed modeling can be leveraged to derive enhanced RIS configurations, leading to substantial improvements in the overall system performance.

Acknowledgment

This work was supported in part by the EU H2020 RISE-6G project under grant 101017011.

References

  • [1] E. Calvanese Strinati, G. C. Alexandropoulos, V. Sciancalepore, M. Di Renzo, H. Wymeersch, D.-T. Phan-huy, M. Crozzoli, R. D’Errico, E. De Carvalho, P. Popovski, P. Di Lorenzo, L. Bastianelli, M. Belouar, J. E. Mascolo, G. Gradoni, S. Phang, G. Lerosey, and B. Denis, “Wireless environment as a service enabled by reconfigurable intelligent surfaces: The RISE-6G perspective,” in IEEE EuCNC/6G Summit, 2021, pp. 562–567.
  • [2] G. Gradoni and M. Di Renzo, “End-to-End Mutual Coupling Aware Communication Model for Reconfigurable Intelligent Surfaces: An Electromagnetic-Compliant Approach Based on Mutual Impedances,” IEEE Wireless Communications Letters, vol. 2337, no. c, pp. 1–5, 2021.
  • [3] X. Qian and M. Di Renzo, “Mutual coupling and unit cell aware optimization for reconfigurable intelligent surfaces,” IEEE Wireless Commun. Lett., vol. 10, no. 6, pp. 1183–1187, 2021.
  • [4] A. Abrardo et al., “MIMO interference channels assisted by reconfigurable intelligent surfaces: Mutual coupling aware sum-rate optimization based on a mutual impedance channel model,” IEEE Wireless Commun. Lett., vol. 10, no. 12, pp. 2624–2628, 2021.
  • [5] P. Mursia, S. Phang, V. Sciancalepore, G. Gradoni, and M. Di Renzo, “Saris: Scattering aware reconfigurable intelligent surface model and optimization for complex propagation channels,” IEEE Wireless Communications Letters, pp. 1–1, 2023.
  • [6] M. A. Yurkin and A. G. Hoekstra, “The discrete dipole approximation: An overview and recent developments,” J. Quantitative Spectroscopy and Radiative Transfer, vol. 106, no. 1, pp. 558–589, 2007.
  • [7] H. E. Hassani, X. Qian, S. Jeong, N. S. Perović, M. Di Renzo, P. Mursia, V. Sciancalepore, and X. Costa-Pérez, “Optimization of RIS-Aided MIMO – A Mutually Coupled Loaded Wire Dipole Model,” 2023. [Online]. Available: http://arxiv.org/abs/2306.09480
  • [8] T. Mazloum, L. Santamaria, F. Munoz, A. Clemente, J.-B. Gros, Y. Nasser, M. Odit, G. Lerosey, and R. D’Errico, “Impact of multiple ris on channel characteristics: An experimental validation in ka band,” in 2023 Joint European Conference on Networks and Communications & 6G Summit (EuCNC/6G Summit), 2023, pp. 13–18.
  • [9] V. Popov, M. Odit, J.-B. Gros, V. Lenets, A. Kumagai, M. Fink, K. Enomoto, and G. Lerosey, “Experimental demonstration of a mmWave passive access point extender based on a binary reconfigurable intelligent surface,” Frontiers in Communications and Networks, vol. 2, 2021.
  • [10] L. Di Palma, A. Clemente, L. Dussopt, R. Sauleau, P. Potier, and P. Pouliguen, “Circularly-polarized reconfigurable transmitarray in ka-band with beam scanning and polarization switching capabilities,” IEEE Transactions on Antennas and Propagation, vol. 65, no. 2, pp. 529–540, 2017.
  • [11] K. Haneda and J.-I. Takada, “An application of sage algorithm for uwb propagation channel estimation,” in IEEE Conference on Ultra Wideband Systems and Technologies, 2003, 2003, pp. 483–487.