1 Wavenumber-Division in Line-of-Sight Holographic MIMO Communications

Luca Sanguinetti, Senior Member, IEEE, Antonio Alberto D’Amico, Merouane Debbah, Fellow, IEEE

Abstract The ultimate performance of any communication system is limited by electromagnetic principles and mechanisms. Motivated by this, we start from the first principles of wave propagation and consider a multiple-input multiple-output (MIMO) representation of a communication system between two spatially-continuous volumes of arbitrary shape and position. This is the concept of holographic MIMO communications. The analysis takes into account the electromagnetic noise field, generated by external sources, and the constraint on the physical radiated power. The electromagnetic MIMO model is particularized for a system with parallel linear sources and receivers in line-of-sight conditions. Inspired by orthogonal-frequency division-multiplexing, we assume that the spatially-continuous transmit currents and received fields are represented using the Fourier basis functions. In doing so, a wavenumber-division multiplexing (WDM) scheme is obtained whose properties are studied with the conventional tools of linear systems theory. Particularly, the interplay among the different system parameters (e.g., transmission range, wavelength, and sizes of source and receiver) in terms of number of communication modes and level of interference is studied. Due to the non-finite support of the electromagnetic channel, we prove that the interference-free condition can only be achieved when the receiver size grows to infinity. The spectral efficiency of WDM is evaluated via the singular-value decomposition architecture with water- filling and compared to that of a simplified architecture, which uses linear processing at the receiver and suboptimal power allocation.

Index Terms High-frequency communications, wavenumber-domain multiplexing, holographic MIMO commu- arXiv:2106.12531v1 [eess.SP] 23 Jun 2021 nications, electromagnetic channels, free-space line-of-sight propagation, degrees-of-freedom, spectral efficiency.

I.INTRODUCTION

In the quest for ever-increasing data rates, it is natural to continue looking for more bandwidth, which in turn pushes the operation towards higher frequencies than in the past. The current

L. Sanguinetti and A. A. D’Amico are with the Dipartimento di Ingegneria dell’Informazione, University of Pisa, Pisa, Italy. M. Debbah is with the Mathematical and Algorithmic Sciences Lab, Huawei Technologies Co., Ltd., France. 2

frontier is in bands up to 86 GHz. However, there is at least 50 GHz of suitable spectrum in the range 90–200 GHz and another 100 GHz in the range 220–320 GHz. These bands are commonly referred to as sub-THz bands and are currently receiving attention for a large variety of emerging applications [1]. As we move up towards high frequencies, the transmission range necessarily shrinks and line-of-sight (LoS) propagation becomes predominant. Classical results in the multiple-input multiple-output (MIMO) literature teach that spatial multiplexing is inevitably compromised in LoS propagation [2]. To understand why this is not the case, recall that such classical results rely on the planar wavefront approximation [3]; that is, the antenna elements experience roughly the same propagation pathloss. This approximation becomes inadequate at high frequencies. Since the wavelength reduces dramatically and the transmission range tends to be short, the wave curvature over the array is no longer negligible and the LoS MIMO channel matrix becomes of high rank [3]–[7]. By properly modeling the wave curvature over the array, one can compute the effective number η of degrees-of-freedom (DoF), which represents the number of independent data streams that can be reliably sent over the channel in the high

signal-to-noise regime. For the two parallel line segments illustrated in Fig. 1, if Ls,Lr  d we have that (e.g., [8]) L L η = s r (1) λd which can also be expressed in terms of the solid angles Ωr = Lr/d and Ωs = Ls/d. This result is derived by using the so-called paraxial approximation [8], which follows from elementary diffraction theory. Motivated by the need of capturing the fundamental properties of wave propagation in wireless communications, we introduce the MIMO representation of a communication system between two spatially-continuous volumes of arbitrary shape and position. The model is developed on the basis of prior analyses in electromagnetic theory (e.g., [8]–[11]). Particularly, starting from the Maxwell’s equations describing electromagnetic fields in any known (i.e., deterministic) channel medium, the MIMO model is obtained by representing the spatially-continuous transmit currents and received fields with arbitrary basis sets of functions. This is the concept of holographic communications [12]–[14]; that is, the capability to generate any current density distribution (in order to obtain the maximum flexibility in the design of the radiated electromagnetic field) and to weight the impinging electric field according to a desired function. The analysis takes into account the electromagnetic nature of noise fields generated by external sources [10], and 3 emphasizes the relation between the energy of transmit currents and the physical radiated power. The ultimate performance limit of such MIMO model with spatially-uncorrelated Gaussian noise is known [8]–[10] and can be achieved by using the eigenfunctions of channel operators as basis sets. In analogy to classical analysis of time- and band-limited systems, the optimal eigenfunctions take the form of prolate spheroidal wave functions [15], whose computation and implementation are in general prohibitive [16]. Our contribution here is thus not to look at the fundamental limits, which are known (under the aforementioned conditions), but to provide a unified representation of the electromagnetic wave communication problem, which can be directly used by communication theorists to design and study implementable (from a technological point of view) communication schemes. This is exactly at the basis of the major contribution of this paper, which consists in particularizing the developed model for the LoS system in Fig. 1 and in proposing a novel wavenumber-division multiplexing (WDM) scheme, inspired by orthogonal-frequency division-multiplexing (OFDM). This is achieved by assuming that the spatially-continuous transmit currents and received fields are represented using the Fourier basis functions, which can be efficiently implemented at the electromagnetic level by means of lens antenna arrays or metasurfaces (e.g., [17], [18]). The properties of WDM (in terms of number of communication modes and level of interference) and the functional dependence on the different system parameters (e.g., transmission range, wavelength, and sizes of source and receiver) are studied using the conventional tools of linear systems theory. Unlike OFDM, we show that WDM cannot convert the MIMO channel into a set of parallel and independent sub-channels due to the non-finite support (in spatial domain) of the electromagnetic channel response. The orthogonality among the communication modes is achieved only when the size of the receiver grows infinitely large. The singular-value decomposition (SVD) architecture with water-filling is then used to decompose the channel into non-interfering and independent single-input single- output channels. Comparisons are made with simplified architectures, which use linear processing at the receiver only to mitigate the interference in the wavenumber domain.

A. Paper outline and notation

This paper is organized as follows. Section II revisits the very general system model and capacity achieving scheme of MIMO communications, with particular emphasis on LoS propaga- tions. In Section III, first principles of wave propagation are used to introduce an electromagnetic MIMO representation of the communication problem between two volumes of arbitrary shape and 4

L /2 Spherical wavefront r

zˆ Ls Solid angle Ωs = Ls/2 d © yˆ xˆ Distance d d −Ls/2  Lr Solid angle Ωr = d

−Lr/2

Fig. 1: Illustration of the approach to estimating the number of degrees-of-freedom in the communication between two linear segments of size Ls and Lr, at a distance d. The solid

angle Ωr subtended by the receive line segment of length Lr at a perpendicular distance d is

Ωr = Lr/d. The solid angle Ωs is similarly given by Ωs = Ls/d. position. In Section IV, the developed representation is applied to two parallel one-dimensional electromagnetic apertures and used to propose a WDM scheme, which can be seen as the spatial-counterpart of OFDM. The spectral efficiency of WDM is evaluated in Section V with different transceiver architectures. Some conclusions and future research directions are discussed in Section VI. We use rank(A) to denote the rank of matrix A. For a given vector p, pˆ is a unit vector along its direction and ||p|| denotes its magnitude. ∇× denotes the curl operation. We denote δ(x) the Dirac delta function and bxc the greatest integer less than or equal to x.

II.MIMOCOMMUNICATIONS

A narrowband time-invariant communication system equipped with Ns antennas at the source and Nr antennas at the receiver can be described, at any arbitrary symbol time, by the following discrete-time input-output relation [2]:

y = Hx + z (2) 5

where y ∈ CNr and x ∈ CNs denote the received and transmitted signal vectors, respectively. The

H 2 vector x must satisfy E{x x} ≤ P to constrain the total transmit power. Also, z ∼ NC(0, σ R)

Nr×Ns is the thermal noise and H ∈ C is the MIMO channel matrix. Here, the entry Hnm represents the gain between the mth source antenna and the nth receive antenna.

A. Capacity and number of degrees-of-freedom

The capacity of the MIMO channel (2) is computed as follows [2]. The received vector y is premultiplied by L−1 with R = LLH to obtain L−1y. Let He = Ue ΛeVe H be the SVD of

−1 H −1 H −1 He = L H. Hence, ye = Ue L y reduces to ye = Λexe + ez where x = Ue xe and ez = Ue L z has 2 independent and identically distributed Gaussian entries, i.e., zen ∼ N (0, σ ). In scalar form, we have

yen = λenxen + zen, n = 1,..., rank(He ) (3)

where λen are the non-zero singular values. The MIMO channel is decomposed into rank(He ) parallel (non-interfering) and independent single-input single-output channels. The capacity of (3) is thus given by [19]

rank(He ) ! X λe2 C = log 1 + p? n (4) 2 n σ2 n=1 ? where pn are the waterfilling solutions: + 2 ! ? σ pn = µ − (5) 2 λen

Prank(He ) ? with µ being such that the total transmit power constraint is satisfied, i.e., n=1 pn = P .

Each non-zero λen corresponds to a communication mode of the channel and can potentially be used to transmit a data stream. The number of communication modes determines the channel DoF [2].

B. Line-of-sight channels

The simplest channel model for H in high-frequency MIMO communications relies on LoS propagation and assume that the source array is located in the far-field of the receive array. This condition refers to the propagation range at which the direction of arrival of the signal and the channel gain are approximately the same for all the elements in the array. In this case, H has 6

√ rank one with λ1 = βNrNs where β is the propagation path loss. The capacity in bit/s/Hz is thus [2]:  βP  C = log 1 + N N (6) 2 r s σ2 which states that, under the far-field LoS assumption, a MIMO system achieves the full array

gain NrNs but can only transmit a single data stream. This is a direct consequence of the far- field assumption and holds true for any arbitrary array geometry, e.g., linear and planar arrays. When the distance between the transmitter and receiver is of the same order as the size of the arrays, the antenna elements experience large differences in the channel gain due to the varying propagation distances and angles. The far-field assumption breaks down and spherical wave modelling is needed [4]–[7]. In this case, uniform linear arrays can give rise to a channel

with N all-equal singular values, provided that Nr = Ns = N and the antenna spacing is ∆ = pλd/N [6]. Similar conditions for a variety of other array geometries can also be found, e.g., [20]. To capture the essence of MIMO communications in different electromagnetic regimes, in the next section we introduce a general MIMO representation of a communication system between two volumes of arbitrary shape and position that is valid for any known channel medium.

III.ELECTROMAGNETIC WAVE COMMUNICATION PROBLEM

Consider two volumes of arbitrary shape and position that communicate through an infinite and homogeneous medium, which is characterized by the scalar frequency- and position-independent −7 permeability µ0 = 4π × 10 [H/m]. An electric current density j(s, t) at the source generates an electric field e(r, t) in [V/m] at a generic location r of the receive volume. We consider only monochromatic sources and electric fields of the form

j(s, t) = <{j(s)e−jωt} and e(r, t) = <{e(r)e−jωt}. (7)

In this case, Maxwell’s equations can be written in terms of the phasor amplitudes j(s) and e(r). The sources are assumed to be bounded, so that the current density satisfies Z kj(s)k2ds < ∞ (8) Vs which guarantees a bounded radiated power, as shown in Section III-E. 7

The electric field y(r) observed inside volume Vr is the sum of the information-carrying electric field e(r) and a random noise field n(r), i.e.,

y(r) = e(r) + n(r) r ∈ Vr (9)

where n(r) is produced by electromagnetic waves that are not generated by the input source. In

Cartesian coordinates, e(r) = ex(r)xˆ + ey(r)yˆ + ex(r)zˆ and n(r) = nx(r)xˆ + ny(r)yˆ + nx(r)zˆ where xˆ, yˆ and zˆ are the unit vectors, independent of position r.

A. The Green function and signal-carrying electric field

The electric current density j(s) at any arbitrary source point s inside volume Vs generates, at any arbitrary point r0, an electric field e(r0) that locally obeys the vector wave equation [21, Eq. (1.3.44)] 0 2 0 0 0 ∇ × ∇ × e(r ) − κ e(r ) = jωµ0j(r ) = jκZ0j(r ) (10)

where κ = ω/c = 2π/λ is the wavenumber (with c being the speed of light) and Z0 =

µ0c = 376.73 [Ohm] is the free-space intrinsic impedance. The solution to (10) is given by [21, Eq. (1.3.53)] Z e(r) = jκZ0 g(r, s)j(s) ds, r ∈ Vr (11) Vs where [21, Eq. (1.3.51)] 1 ejκkr−sk  ∇ ∇H  g(r, s) = I + r r (12) 4π kr − sk κ2 is the Green’s function for the electric field in an unbounded, homogeneous medium. In free- space, under the far-field approximation1 (i.e. kr − sk  λ), the Green’s function in (12) can be approximated as [22] 1 ejκkr−sk g(r, s) ≈ (I − ppH) (13) 4π kr − sk bb where pb = p/||p|| and p = r − s.

1The far-field approximation in electromagnetic propagation corresponds to the radiated field that falls off inversely as the distance apart kr − sk. Hence, its power follows the inverse square law. 8

Fig. 2: Optimal transceiver architecture for any choice of vectors {φm(s); m = 1,...,N} and

{ψn(r); n = 1,...,N}.

B. The noise field

The noise field n(r) is produced by the incoming electromagnetic waves that are not generated by the source. Suppose that they are generated in the far-field of the receiving volume. Each electromagnetic wave can thus be modelled as a plane wave that reaches the receiving volume from an arbitrary elevation angle θr ∈ [0, π) and an arbitrary azimuth angle ϕr ∈ [−π, π). The noise field n(r) is thus Z π Z π n(r) = n(r, θr, ϕr)dθrdϕr (14) −π 0 where n(r, θr, ϕr) is defined as

T jκ (θr,ϕr)r n(r, θr, ϕr) = a(θr, ϕr)e . (15)

Here, 2π κ(θ , ϕ ) = [cos ϕ sin θ , sin ϕ sin θ , cos θ ]T (16) r r λ r r r r r is the wave vector that describes the phase variation of the plane wave with respect to the

three Cartesian coordinates at the receiving volume, a(θr, ϕr) is a zero-mean, complex-Gaussian random process with

H 0 0 2 0 0 E{a(θr, ϕr)a (θr, ϕr)} = σ f(θr, ϕr)I3δ(ϕr − ϕr)δ(θr − θr) (17)

2 2 2 2 and σ f(θr, ϕr) denotes the noise power angular density with σ being measured in V /m . From (14), using (15) and (17) yields

0 H 0 2 E{n(r )n (r + r )} = σ ρ(r)I3 (18) 9

with π π Z Z T −jκ (θr,ϕr)r ρ(r) = f(θr, ϕr)e dθrdϕr. (19) −π 0

The above model is valid for any f(θr, ϕr). In the case of electromagnetic waves not generated by the source, we can reasonably assume that they are uniformly distributed in the angular domain. This corresponds to an isotropic propagation condition, characterized by sin θ f(θ , ϕ ) = r θ ∈ [0, π), ϕ ∈ [−π, π). (20) r r 4π r r Plugging (20) into (19) yields π π   Z Z T sin θr −jκ (θr,ϕr)r 2||r|| ρ(r) = e dθrdϕr = sinc . (21) −π 0 4π λ Remark 1. In MIMO communications, the noise samples in (2), taken at arbitrary points r0 and r + r0, are typically modelled as independent zero-mean and circularly-symmetric Gaussian random variables, i.e., ρ(r) = σ2δ(r). This is not generally the case for noise samples of electromagnetic nature as it follows from (21). It happens only when ||r|| = iλ/2, with i ∈ Z, i.e., all samples of n(r) must be taken along a straight line at a spacing of an integer multiple of λ/2.

C. The electromagnetic MIMO channel

Assume that the current density j(s) is expanded using a set of orthonormal vectors {φm(s); m = 1,...,N} with N being the dimension of the input signal space. Without loss of generality, we assume that N is an odd number. Accordingly, we can write

N X j(s) = ξmφm(s) (22) m=1

where the coefficient ξm satisfies Z H ξm = φm(s)j(s)ds. (23) Vs The received field y(r) in (9) is projected onto an output space of dimension N, spanned by

a set of receive orthogonal basis functions {ψn(r); n = 1,...,N}. We can thus represent the projected y(r) by the coefficients Z H yn = ψn(r)y(r)dr. (24) Vr 10

Using (22) – (24), the input-output relationship takes the form N X yn = Hnmxm + zn n = 1,...,N (25) m=1 where

xm = jκZ0ξm (26) are the effective input samples while Z H zn = ψn(r)n(r)dr (27) Vr and Z Z H Hnm = ψn(r)g(r, s)φm(s)drds. (28) Vr Vs

T T Letting y = [y1, . . . , yN ] and x = [x1, . . . , xN ] , we may rewrite (25) in the same matrix form

N×N T 2 of (2) where H ∈ C is the channel matrix and z = [z1, . . . , zN ] ∼ NC(0N , σ R) with ZZ 0 H 0 0 [R]nm = ρ(r − r )ψn(r)ψm(r )drdr (29) Vr as it follows from (27), by using (18) and (21). Unlike (2), where the entries of H represent the channel gains from transmit to receive antennas, here they represent the coupling coefficients between the source mode m and the reception mode n [8]. The capacity of the described electromagnetic MIMO system is thus given by (4) where the communication modes are obtained by computing the singular value decomposition of He = L−1H with the entries of H and R given by (28) and (29), respectively. The capacity achieving transceiver architecture for any choice of vector basis functions {φm(s); m = 1,...,N} and {ψn(r); n = 1,...,N} is reported in Fig. 2, where we have called x(s) = jκZ0j(s). Notice that the transceiver operates in two stages. Particularly, the inner stage operates directly at the electromagnetic level since it is responsible for generating at the source the current density distribution j(s) through the use of functions

{φm(s); m = 1,...,N} and for weighting at the receiver the electric field y(r) according to the functions {ψn(r); n = 1,...,N}.

D. The electromagnetic radiated power

In signal processing, the quantity on the left-hand-side of (8) represents the energy of the continuous-space signal j(s). Notice that this is different from the electromagnetic radiated power

Prad, which can be computed using the Poynting theorem and integrating the radial component 11 of the Poynting vector over a sphere of radius r → ∞ [23, Ch. 14]. In Appendix A, it is shown that Prad can be upper-bounded as follows

Prad ≤ QEs (30) where E = R kj(s)k2ds and s Vs sZZ κZ0 2 Q = |ρ(s1 − s2)| ds1ds2 (31) 4λ Vs with ρ(·) being defined in (21). The upper-bound in (30) shows that imposing the constraint (8) on the L2 norm of the source j(s) amounts to limit the radiated power. On the contrary, 2 bounding Prad does not necessarily lead to a bounded source L norm in (8); e.g., [9].

IV. WAVENUMBER-DIVISION MULTIPLEXING

The optimal approach for the design of the communication system depicted in Fig. 2 consists in selecting the elements of the input basis {φn(s); n = 1,...,N} as the eigenfunctions of the channel operator (e.g., [8]– [10]) Z 0 H 0 Ks(s, s ) = g (r, s)g(r, s )dr. (32) Vr This means that Z 0 0 0 γnφn(s) = Ks(s, s )φn(s )ds (33) Vs where γn is the eigenvalue associated to the eigenfunction φn(s). The elements of the output basis {ψn(r); n = 1,...,N} are simply the channel responses to the functions of the input basis, i.e., Z ψn(r) = g(r, s)φn(s)ds. (34) Vs The above approach is optimal under the assumption that the electromagnetic noise can be modelled as a spatially-uncorrelated Gaussian random field. This may not be the case as it follows from the model described in Section III-B. In addition, we stress that the numerical solution of the eigenfunction problem (33) is in general computationally demanding. Moreover, its implementation is prohibitive not only at the electromagnetic (i.e., analog) level but also in the digital domain. We can reasonably ask ourselves whether there is a pair of basis for which the communication system becomes simple to implement and with a clear physical interpretation. To answer this question, we consider the LoS scenario in Fig. 1 and take inspiration from 12

a communication system operating over a time-domain dispersive (i.e., frequency selective) channel. The communication over this channel is mathematically equivalent to a time-domain MIMO system. By using the OFDM technology, it can be converted into a frequency-domain communication system that transmits over multiple non-interfering frequency-flat channels. This is simply achieved by using Fourier transform operations and a cyclic prefix technique. The objective of this section is to show that a communication scheme sharing these advantages can be obtained from the electromagnetic MIMO channel in the wavenumber domain. We refer to it as WDM (wavenumber-division multiplexing)

A. Review of orthogonal-frequency-division multiplexing

Consider a system that employs N subcarriers with frequency spacing 1/T and indices in the set {−(N − 1)/2, ..., (N − 1)/2} with N being an odd integer. The base-band equivalent

time-domain signal transmitted over the interval 0 ≤ t ≤ Ts with Ts ≥ T is  N  j 2π m−1− N−1 t  √1 P T 2  T xme , 0 ≤ t ≤ Ts x(t) = s m=1 (35)  0, elsewhere

m−1−(N−1)/2 where xm is the data associated to the frequency fm = T . The signal x(t) is sent over a wireless channel with impulse response g(t). Wireless channel responses are causal and incur

a finite maximum delay. We thus assume that the support of g(t) is the time interval 0 ≤ t ≤ Tg. The received signal y(t), given by

y(t) = g(t) ⊗ x(t) + z(t) (36) where z(t) is the thermal noise, is observed in Tg ≤ t ≤ Tg + Tr. Accordingly, its Fourier transform, evaluated at fn, is written as:

Z Tg+Tr −j 2π (n−1− N−1 )t yn = y(t)e T 2 dt. (37) Tg If

Ts ≥ T + Tg and Tr = T (38) then (37) takes the form [2]:

yn = Hnxn + zn (39) 13

√ where Hn = GnTr/ Ts, with Gn being the channel response at frequency fn, and zn the corresponding noise. The expression in (39) describes a set of N non-interfering (orthogo- nal) parallel transmissions with different complex-valued attenuation factors. Orthogonality is achieved by setting Ts ≥ T + Tg in (38), which amounts to append the so-called cyclic prefix to the transmitted signal so that at the receiver (36) can be viewed as a cyclic convolution in the interval [Tg,Tg + Tr). In the frequency-domain, the cyclic convolution becomes the product of the corresponding Fourier transforms and (39) follows easily. It is worth noticing that the same 2 result is achieved if the received signal is observed in the interval [0,Tr), and we set

Ts = T and Tr ≥ T + Tg. (40)

In this case, no cyclic prefix is appended at the transmitter but the received signal is observed over an extended time interval including the support of the channel response. The input-output relationship is analogous to (39) but with p Hn = TsGn. (41)

Although this is not the way OFDM operates, the implication of the conditions (40) are in- strumental for the design and understanding of the WDM communication scheme described below.

B. System and signal model

To simplify the analysis and better highlight the similarities with time-domain communication systems and, in particular, with OFDM, we consider the one-dimensional setup illustrated in Fig. 1, which relies on the following assumption.

Assumption 1. The source is a segment with the following coordinates:

{(sx, sy, sz): sx = 0, sy = 0, |sz| ≤ Ls/2}. (42)

The receiver is a segment occupying the following region:

{(rx, ry, rz): rx = d, ry = 0, |rz| ≤ Lr/2, }. (43)

3 Also, Lr = `Ls with ` ≥ 1 being an integer.

2This result will be proved in Appendix C for the proposed scheme. 3Notice that this assumption is needed for the output basis functions to be orthogonal, but it is not strictly necessary. 14

0.5

0.4

0.3

0.2

0.1

0 -6 -4 -2 0 2 4 6

Fig. 3: Amplitude of I(κz) for Ls = 0.2 m and λ = 0.01 m when N = 3 and ξm = 1 for m = 1, 2, 3.

Under Assumption 1, the electric current density is

j(s) = i(sz)δ(sx)δ(sy)zˆ (44)

where N X i(sz) = ξmφm(sz) (45) m=1 is a current (measured in amperes). In analogy with an OFDM system, we assume that the basis

functions {φm(sz); m = 1,...,N} take the form

 2π N−1  1 j m−1− sz Ls  √ e Ls 2 , |sz| ≤  Ls 2 φm(sz) = (46)  0, elsewhere such that Ls Z 2 ∗ ξm = i(sz)φm(sz)dsz. (47) Ls − 2

The wavenumber Fourier transform of i(sz) is

Ls N Z 2 −jkzsz X I(κz) = i(sz)e dsz = ξmΦm(κz) (48) Ls − 2 m=1 where Ls Z 2    −jkzsz p κz m − 1 − (N − 1)/2 Φm(κz) = φm(sz)e dsz = Lssinc − Ls . (49) Ls 2π Ls − 2 15

0 -40

-50 -20

-60 -40 -70

-60 -100 -50 0 50 100 -80 -2 -1 0 1 2

(a) Amplitude of |gz(x, d)|. 2 (b) Value of |Gz(κz, d)| .

2 Fig. 4: Behaviour of |gz(z, d)| and |Gz(κz, d)| for d = 5, 10 and 25 m when λ = 0.01 m (i.e., carrier frequency of 30 GHz).

Fig. 3 plots the amplitude of I(κz) in (48) for Ls = 0.2 m, λ = 0.01 m and N = 3. The distance

between the first two nulls of each sinc function is 4π/Ls. Fig. 3 shows that Φm(κz) has most of its energy in the vicinity of 2π (m − 1 − (N − 1)/2). Hence, I(κ ) is practically limited to Ls z  N − 1 2π N − 1 2π  − , (50) 2 Ls 2 Ls and thus has an approximate bandwidth of Ω = πN . Ls

At the receiver, the electric field y(r) is projected onto the space spanned by the vectors ψn(r) given by

ψn(r) = ψn(rz)δ(rx − d)δ(ry)zˆ (51)

where, in analogy with OFDM,

 2π N−1  j n−1− rz  e Ls 2 , |rz| ≤ Lr/2 ψn(rz) = (52)  0, elsewhere.

The output samples yn in (24) are expressed in volts. In the wavenumber domain, we have    κz n − 1 − (N − 1)/2 Ψn(κz) = Lrsinc − Lr (53) 2π Ls where the distance between the first two nulls is now 4π/Lr and thus inversely proportional to

Lr, rather than Ls. From (24), by using (52) we obtain Lr Lr Z 2 Z 2 2π N−1  ∗ −j L n−1− 2 rz yn = ψn(rz)y(d, 0, rz)drz = y(d, 0, rz)e s drz (54) Lr Lr − 2 − 2 16

which is simply the nth coefficient of the wavenumber Fourier series expansion of y(d, 0, rz) taken at 1/Ls. Clearly, the projection of y(r) onto the space spanned by (52) allows to only capture the electric field along the z−axis. Notice that (52) are exponential functions whose fundamental spatial frequency is 1/Ls.

C. The coupling coefficients

Under the system model above, in free-space propagation (28) reduces to

Lr Ls Z 2 Z 2 ∗ Hnm = gz(rz − sz, d)ψn(rz)φm(sz)dszdrz (55) Lr Ls − 2 − 2

where gz(z, d) denotes the (3, 3)th element of g(r, s) evaluated at rx = d and sx = sy = ry = 0.

From (13), we obtain √ d2 ejκ z2+d2 gz(z, d) = . (56) 4π (z2 + d2)3/2 In Appendix B, the double integral (55) is rewritten as a single integral to facilitate its evaluation. We call Z ∞ −jκzz Gz(κz, d) = gz(z, d)e dz (57) −∞

the Fourier transform in the wavenumber-domain of gz(z, d), and hence we can write Z ∞ 1 jκzz gz(z, d) = Gz(κz, d)e dκz. (58) 2π −∞ 2 The behaviour of |gz(z, d)| and |Gz(κz, d)| in dB is illustrated in Fig. 4 for d = 1, 5, 10 and

25 m. Unlike time-domain channels, gz(z, d) is not causal and has an unbounded support. Notice

also that Gz(κz, d) is band-limited in the interval |κz| ≤ κ = 2π/λ. The lower the wavelength λ, the larger the spatial-frequency bandwidth. We notice that, for the investigated scenarios, the

3−dB bandwidth of Gz(κz, d) is approximately ακ with α = 0.6075. To quantify the impact of the propagation channel, we reformulate (55) as follows

Lr Z 2 ∗ Hnm = ψn(rz)θm(rz)drz (59) Lr − 2 with Ls Z 2 θm(rz) = gz(rz − sz, d)φm(sz)dsz (60) Ls − 2 being the mth element of the basis spanning the electric field at the receiver, i.e., N X e(d, 0, rz) = xmθm(rz). (61) m=1 17

-3 10-3 10 2.5 2

2 1.5

1.5 1 1 0.5 0.5

0 0 -6 -4 -2 0 2 4 6 -6 -4 -2 0 2 4 6

L = 0.4 (a) r m. (b) Lr = 5 m.

Fig. 5: Amplitude of E(κz) and Θm(κz) for m = 1, 2, 3 in the same setup of Fig. 3 when d = 5 m and Lr = 0.4 and 5 m.

We call E(κz) and Θm(κz) the Fourier transforms in the wavenumber domain of e(d, 0, rz) and

θm(rz), respectively. Fig. 5 plots their behaviors in the same setup of Fig. 3 when d = 5 m and

Lr = 0.4 and 5 m. From Fig. 5a, we see that when Lr = 0.4 the spectrum of the electric field is highly distorted compared to that in Fig. 3. Moreover, the orthogonality condition among the three communication modes is destroyed. The situation is much different when Lr = 5 m. In this case, the electric field is received undistorted over the considered wavenumber interval and the three communication modes are not affected by interference. By using (58), we may rewrite (55) as 1 Z H = G (κ , d)Ψ∗ (κ )Φ (κ )dκ (62) nm 2π z z n z m z z ∗ where Φm(κz) and Ψn(κz) are given by (49) and (53), respectively. Hence, we obtain √ Z ∞    LsLr κz n − 1 − (N − 1)/2 Hnm = Gz(κz, d)sinc − Lr 2π −∞ 2π Ls    κz m − 1 − (N − 1)/2 × sinc − Ls dκz (63) 2π Ls √ 2 In Fig. 6, we illustrate |Hnn/ Ls| for the same setup of Fig. 4, when Lr = 1 and 10 m. As seen, √ 2 2 the behaviour of |Hnn/ Ls| is much different from that of |Gz (κz, d) | in Fig. 4. The number of significant coefficients reduces as d increases. This is exemplified in Fig. 7 where we plot the

number N of coupling coefficients Hnn that resides in the 3−dB bandwidth of Gz (κz, d) for

Lr = 1 and 5 m. We see that if d = 5 m then N = 3 and 17 for Lr = 1 m and 5 m, respectively. 18

-40 -40

-50 -50

-60 -60

-70 -70

-80 -80 -2 -1 0 1 2 -2 -1 0 1 2

(a) L = 1 m r (b) Lr = 5 m √ 2 Fig. 6: Behaviour of |Hnn/ Ls| in (63) in dB in the same setup of Fig. 4 when Lr = 1 and 5 m.

These numbers reduce to N = 1 and 9 when d = 10 m. The solid curves are obtained from (1) as (in order to have an odd number of communication modes) 1 L L  2 s r + 1 (64) 2 λd and represent a good approximation of N only when Lr is of the same order of d. This is in agreement with the fact that (1) is valid only in that regime (e.g., [8]). A comparison between √ 2 Fig. 6a and Fig. 6b shows that, for a given distance d, the coupling coefficient |Hnn/ Ls| tends 2 to the sampled version of |Gz (κz, d) | as Lr increases. In fact, the following result holds true

when Lr grows unboundedly.

Corollary 1. If Lr → ∞ then  √  LsGn, n = m Hnm → (65)  0 n 6= m with 2π  N − 1  Gn = Gz n − 1 − , d . (66) Ls 2

Proof: If Lr → ∞ then 1  n − 1 − (N − 1)/2 Ψn(κz) → δ κz − 2π . (67) 2π Ls By using the above result into (63) yields p  Hnm → LsGnsinc n − m (68) 19

where G in (66) is obtained from (63) by taking κ = 2π (n−1−(N −1)/2). Since sincn−m n z Ls is 1 for n = m and 0 otherwise, the proof is completed. √ The above corollary shows that asymptotically Hnn/ Ls tends to Gn. Since Gz(κz, d) is band-limited with bandwidth Ω = 2π/λ, the maximum dimension of the input signal space depends on the wavenumber spectrum and is given by L  N = 2 s + 1 (69) max λ which corresponds to the maximum DoF of the electromagnetic channel in the 1D setup in Assumption 1; e.g., [24]. If the 3−dB bandwidth is considered, then the approximate dimension reduces to bαNmaxc with α = 0.6075. Recalling that Hnm represents the coupling coefficient between source mode m and reception mode n, the above corollary shows that the communication modes are decoupled only when Lr grows to infinity, i.e., when an infinitely large receiving segment is used. This is a direct consequence of the fact that gz(z, d) is not limited in the spatial-domain. However, the following result can be proved.

Corollary 2. If |gz(z, d)| is much smaller than |gz(0, d)| for |z| ≥ zg, i.e.,

|gz(z, d)|  |gz(0, d)| |z| ≥ zg (70)

then

Lr ≥ Ls + zg (71)

guarantees (65).

Proof: See Appendix C.

The above corollary shows that the orthogonality condition (65) can be achieved if Lr is larger than a predefined value zg for which |gz(z, d)| in (56) is negligible for |z| ≥ zg. Notice that (71) can be viewed as the spatial-counterpart of the time-domain condition (40) in OFDM systems.

In other words, the extra length Lr −Ls ≥ zg allows to observe the received signal over its entire support in the spatial-domain. As proved in Appendix C, this restores the orthogonality among the communication modes and basically converts the spatial-domain dispersive channel into a set of flat channels in the wavenumber-domain. Numerical results will show that the asymptotic

orthogonality condition (65) is sufficiently accurate starting from a receiver size Lr of the same order of the distance d from the source. 20

40

30

20

10

1 5 10 20 30 40 50

Fig. 7: Number N of coupling coefficients Hnn that resides in the 3−dB bandwidth of Gz (κz, d).

The solid curves are obtained by using (90) with Ls = 0.2 m and λ = 0.01 m.

Remark 2. Fig. 7 shows that N tends to 1 as d increases for any size Lr. Notice that when d

grows large and Lr,Ls, λ are kept fixed, we can approximate gz(z, d) as 1 ejκd g (z, d) ≈ (72) z 4π d which corresponds to the Green’s function in the far-field of source and receiver. This is valid for the propagation range at which the direction and channel gain are approximately the same. In these circumstances, (55) reduces to

Lr Ls jκd Z 2 Z 2 1 e ∗ Hnm = ψn(rx)drx φm(sx)dsx (73) 4π d Lr Ls − 2 − 2 from which it follows  1 ejκd √  4π d Lr Ls, n = m = (N − 1)/2 + 1 Hnm = (74)  0 otherwise. In line with classical MIMO results, a single degree-of-freedom is available in the far-field case.

D. The noise samples

From (27), the noise samples are

Lr Z 2 ∗ zn = ψn(rz)n(d, 0, rz)drz (75) Lr − 2 and the matrix R in (29) has entries

Lr ZZ 2 0 ∗ 0 0 [R]nm= ρ (rz − rz) ψn(rz)ψm(rz)drzdrz (76) Lr − 2 21

10 10

0 0

-10 -10

-20 -20

-30 -30

-40 -40

-50 -50 1 2 3 4 0 10 20 30 40

(a) λ = 0.1 m (b) λ = 0.01 m

Fig. 8: Normalized eigenvalues in dB of R when Ls = 0.2 m and Lr = 0.1, 0.2 and 0.4 m with λ = 0.1 and 0.01 m.

 0  0 rz−rz where ρ(rz − rz) = sinc 2 λ as it follows from (21) under Assumption 1. The double- integral in (76) can be computed by following the same steps used in Appendix B for the computation of (55). As done for (55), we may rewrite (76) as Z ∞ 1 ∗ [R]nm = Sz(κz)Ψn(κz)Ψm(κz)dκz (77) 2π −∞ where π  κ  λ  κ  S (κ ) = rect z = rect z (78) z z κ 2κ 2 2κ is the wavenumber Fourier transform of ρ(z). This is a rectangular function of bandwidth κ.

Hence, the noise samples {zn} are correlated despite the isotropic nature of the electromagnetic

noise n(r). This is a direct consequence of the finite size of Lr that limits the observation interval of n(r) in the spatial domain. To quantify the spatial correlation properties of noise samples

{zn}, Fig. 8 plots the normalized eigenvalues in dB of R when Ls = 0.2 m. We consider either

λ = 0.1 m or λ = 0.01 m corresponding to Nmax = 4 and Nmax = 40, respectively. In both cases,

we see that the eigenvalues are different only when Lr is small while they become all equal as

Lr ≥ 0.4 m. The following asymptotic result is obtained.

Corollary 3. If Lr → ∞, then

 2 [R]  σ , n = m σ2 nm → 2κ (79) Lr  0 n 6= m 22

for n ≤ 2Ls/λ.

Proof: It follows from (77) by using (78) and the same steps in the proof of Corollary 1.

The implication of the above corollary is that, as Lr grows large, the covariance matrix

Lr R becomes diagonal with elements given by 2κ ; that is, the noise samples {zn} become uncorrelated with a variance that increases unboundedly. In practice, this unbounded behaviour of the noise variance is a consequence of the model in Section III-B, which becomes not physically meaningful in the asymptotic regime as it results into noise samples of infinite power. Moreover, it would asymptotically lead to a communication system characterized by a null spectral efficiency

since Corollary 1 showed that the coupling coefficients tend to a finite quantity as Lr → ∞. However, we anticipate that all this is not an issue since the asymptotic orthogonality condition in Corollary 1 is sufficiently accurate for Lr ≥ d. Hence, in scenarios of practical relevance, the noise variance will be large but always finite, and the noise model from Section III-B can be safely used.

V.

There is a clear analogy between the WDM scheme described above and OFDM. In both cases, Fourier transform operations are applied to convert a matrix channel into a set of parallel independent sub-channels. In OFDM, this is achieved for any finite observation interval Tr that satisfies (40). On the other hand, the orthogonality and independence conditions among communication modes is achieved with WDM only when Lr → ∞. This implies that, for finite values of Lr, the SVD architecture is needed to decompose the channel into non-interfering and independent single-input single-output channels. Next, we consider two different WDM schemes, i.e., with and without the SVD of He , and make comparisons in terms of spectral efficiency. In doing so, we impose the following constraint

Ls Z 2 1 2 |i(sz)| dsz ≤ Ps (80) Ls Ls − 2 and define the system SNR as N N 2 (b) 1 1 X (a) (κZ0) 1 X P snr = |x |2 = |ξ |2 ≤ (81) σ2 L n σ2 L n σ2 s n=1 s n=1 PN 2 where (a) follows from (26) and (b) from n=1 |ξn| /Ls ≤ Ps, as obtained by plugging (45) 2 2 2 2 into (80) and defining P = (κZ0) Ps. Notice that Ps is measured in [A ] and P in [V /m ]. 23

A. With Singular-Value Decomposition

As any MIMO system, the capacity of WDM is given by (4) and it can be achieved via the transceiver architecture in Fig. 2. This requires the SVD of He = L−1H where H is the electromagnetic MIMO channel with coupling coefficients as in (55) or (62), and L is obtained

T 2 from the Cholesky decomposition of the covariance matrix of z = [z1, . . . , zN ] ∼ NC(0N , σ R) where the elements of R are given by (76). From Fig. 8, it follows that R can be well- approximated by a diagonal matrix since the noise samples become uncorrelated for practical

values of Lr; that is, the whitening transformation in Fig. 2 reduces to a simple scaling operation. Notice also that H can be precomputed in the considered free-space propagation conditions. Particularly, H is a deterministic matrix, whose computation requires knowledge of the Green’s

function and the Fourier basis functions {φm(s); m = 1,...,N} and {ψn(r); n = 1,...,N} given by (46) and (52), respectively. They all depend on the system parameters, e.g., wavelength

λ, distance d, sizes Ls and Lr. The same is true for the computation of R, which requires knowledge of the noise power density σ2. Notice also that H and R are square matrices of size N, i.e., the dimension of the input signal space or, equivalently, the number of communication modes. Since N ≤ Nmax = 2Ls/λ, it follows that the maximum number of complex operations 2 required for pre-processing x and post-processing y (through Ve and Ue , respectively), is 2Nmax. This is an affordable complexity for scenarios of practical interest. Particularly, if λ = 0.01 m, then Nmax = 40 and Nmax = 100 for Ls = 0.2 m and Ls = 0.5 m. It increases to Nmax = 400

and Nmax = 1000 if λ = 0.001 m.

B. With Linear Processing at the Receiver only

A different scheme is considered here, which does not perform any pre-processing at the

H H H source, i.e., Ve = IN , while it still applies linear processing at the receiver with Ue = [ue1 ,..., ueN ] . −1 −1 In this case, ye = ULe y, x = xe, and ez = ULe z has independent and identically distributed 2 2 2 2 Gaussian entries with zen ∼ N (0, σn) and σn = σ ||uen|| . In scalar form, we obtain N H X H yen = ue1 henxn + ue1 hem xm + zen (82) | {z } m=1,m6=n desired signal | {z } interference

H H H where He = [he1 ,..., heN ] . Assume that xn ∼ CN (0, pn) and that the decoding is performed

by treating the interference as Gaussian noise. For any fixed power set {pn; n = 1,...,N}, the 24

3 103 10 Optimal scheme Optimal scheme © 102 102 ©

101 101

100 100 1 2 3 4 5 1 2 3 4 5

(a) Distance d = 5 m. (b) Distance d = 10 m.

Fig. 9: Spectral efficiency of WDM in bit per channel use. The number of communication modes is (90), as obtained with the paraxial approximation. The asymptotic curve refers to a system in

which Lr → ∞. spectral efficiency of this communication system is (measured in bit per channel use)

N X SE = log2(1 + SINRn) (83) n=1 where

H 2 |u hen| pn SINR = en (84) n N P H 2 2 H |uenhem| pm + σ uenuen m=1,m6=n is the signal-to-interference-plus-noise ratio (SINR). This is a generalized Rayleigh quotient with respect to uen and thus is maximized by the MMSE combiner, i.e., N !−1 MMSE X H 2 uen = pmhemhem + σ eIN hen. (85) m=1

From Corollary 1, we know that the interference in (84) vanishes as Lr grows large. In this MR case, MMSE reduces to maximum ratio (MR) combining, i.e., uen = hen, which has lower computational complexity and maximizes the power of the desired signal; that is, it is optimal in a noise-limited regime. The powers that maximize (84) can be obtained by means of the iterative waterfilling algo- rithm, whose convergence is not always guaranteed as it depends on the amount of interference

(e.g., [25]). To overcome this problem, we exploit Corollary 1 and assume that Lr is sufficiently 25

large such that the interference in (84) is reasonably smaller than the desired signal. In this case, the optimal powers can be computed as + 2 ! ? σ pn = µ − (86) 2 ||hen||

PN ? with n=1 pn = P . Unlike (5), the waterfilling solutions (86) are computed with respect to 2 the coupling coefficients {||hen|| ; n = 1,...,N}. In addition, if the communications modes are transmitted over a wavenumber interval for which the absolute values of the coupling coefficients

{Hnn : n = 1,...,N} are nearly constant (as shown in Fig. 6), then (86) reduces to a uniform power allocation.

C. Numerical analysis

Numerical results are used to quantify the spectral efficiency of WDM. We assume Ls =

0.2 [m] and λ = 0.01 [m]. The maximum number of communication modes is Nmax = 2Ls/λ =

40 with spacing in the wavenumber domain of 2π/Ls = 31.41 rad/m. The system SNR given −7 2 by (81) is snr = 70 dB and the source power is Ps = 10 [A ]. Accordingly, from Section III-D −3 we have that the radiated power in (30) is Prad ≤ 3.7 × 10 [W/m] and the noise power density is σ2 = 5.6 × 10−4 [V2/m2]. Comparisons are made with a communication system in which the

basis sets {φm(s); m = 1,...,N} and {ψn(r); n = 1,...,N} are chosen as the eigenfunctions of channel operators. This is reported as a reference. As discussed in Section IV, this requires

to select φn(sz) in (46) as the solution of the following eigenfunction problem:

Z Ls/2 0 0 0 γnφn(sz) = Ks(sz, sz)φn(sz)dsz (87) −Ls/2 with

Z Lr/2 0 ∗ 0 Ks(sz, sz) = gz (rz − sz, d)gz(rz − sz, d)drz. (88) −Lr/2 The output basis function in (52) are then obtained as

Z Lr/2 ψn(rz) = gz(rz − sz, d)φn(sz)dsz. (89) −Lr/2

Fig. 9 plots the spectral efficiency as a function of Lr when d = 5 and 10 m. According to (1), we assume that the number of communication modes is 1 L L  N = 2 s r + 1. (90) 2 λd 26

103

Optimal scheme © 102

101

100 10 20 30 40 50

Fig. 10: Spectral efficiency of WDM in bit per channel use as a function of distance when

Lr = 5 m.

We see that the spectral efficiency of WDM with SVD and MMSE processing is exactly the same and matches well the asymptotic performance (i.e., free-interference case) for any value

of Lr. On the contrary, WDM with MR processing approaches the asymptotic performance only when Lr is large. As a rule of thumb, this is approximately achieved by Lr ≥ d. Compared to the optimal scheme, a substantial loss is observed. However, we recall that the superior performance of the optimal scheme is achieved at the cost of a prohibitively high implementation complexity. Fig. 10 plots the spectral efficiency as a function of the distance. We see that WDM with SVD and MMSE processing performs very close irrespective of the distance. The spectral efficiency reduces as d increases since the received power as well as the number of communication modes decreases.

VI.CONCLUSIONSAND FUTURE RESEARCH DIRECTIONS

Building on prior analyses (e.g., [8]–[10]), we provided a MIMO representation of the elec- tromagnetic wave propagation problem between two spatially-continuous volumes, which can be directly used by communication theorists as a baseline for developing and studying optimal or suboptimal (but implementable) holographic MIMO communication schemes. Particularly, we used it to develop a communication scheme for the system setup in Fig. 1; that is, linear sources and receivers in LoS propagation conditions. Inspired by OFDM, we made use of Fourier basis functions and proposed a WDM communication scheme that operates in the wavenumber domain and makes use of Fourier transform operations directly at the electromagnetic level. Conventional tools of linear systems theory were used to understand the interplay among the different system 27 parameters in terms of number of communication modes and level of interference. Unlike OFDM, the orthogonality among the communication modes (in the wavenumber domain) is achieved with WDM only when the size of the receiver grows infinitely large, due to the unbounded support of the channel response in the spatial domain. Different communication architectures were thus used and designed to deal with the interference. Numerical results showed that, for the investigated scenarios, the optimal SVD architecture with water-filling and the linear MMSE receiver with suboptimal power allocation achieve the same spectral efficiency of the interference-free case (infinitely large receiver), while MR combining processing performs sufficiently well starting from a receiver size of the same order of the distance from the source. Future research directions for the proposed WDM scheme are clearly represented by con- sidering scenarios where some of the underlying assumptions of Fig. 1 are not valid. A first extension is to study the effects of multipath propagation and LoS blockage. Since in practice the source and receiver are not parallel, the impact of arbitrary positions and orientations should be investigated. This would open the door for multi-user communications in which multiple sources (arbitrarily located in the area) transmit to a fixed receiver. A natural extension is also the design of the WDM scheme when two-dimensional (rather one-dimensional) surfaces are used for transmission and reception [16]; this provides additional flexibility in terms of generating the current densities and processing the impinging electric fields.

APPENDIX A

R 2 From [23, Ch. 14], the radiated power Prad is given by Prad = limr→∞ Ω Prr dΩ where

1 H Pr = e (r)e(r) (91) 2Z0 is the radial component of the Poynting vector [23, Ch. 14], e(r) is the radiation electric field with r = ||r||, and Ω is the solid angle of 4π steradians. When r → ∞, the electric field reduces to jκr Z T e H −jκ (θs,ϕs)s e(r) = jκZ0 (I − brbr ) j(s)e ds (92) 4πr Vs since the Green’s function (13) can be approximated as jκr 1 e T g(r, s) ≈ (I − rrH) e−jκ (θs,ϕs)s. (93) 4π r bb Plugging (92) into (91) yields 2 2 Z 1 Z T Z 1 Z T P = 0 (I − rrH) j(s)e−jκ (θs,ϕs)s ds ≤ 0 j(s)e−jκ (θs,ϕs)s ds . (94) r 2 2 bb 2 2 8λ r Vs 8λ r Vs 28

From (94), we thus have that Prad can be upper bounded as 2 Z Z T ZZ Z0 −jκ (θs,ϕs)s Z0 H Prad ≤ 2 j(s)e ds dΩ = 4π 2 j (s1)j(s2)ρ(s1 − s2)ds1ds2 8λ Ω Vs 8λ Vs where ρ(·) is given in (21). By applying the Cauchy-Schwarz inequality, we obtain (30).

APPENDIX B

Rewrite the integral in (55) as Hnm = A1 + A2 + A3 with

−L /2+L /2  L /2+x  rZ s sZ  j2π(m−n)r/Ls  A1 = fm(x)  e dr dx (95)

−Lr/2−Ls/2 −Lr/2

L /2−L /2  L /2+x  r Z s sZ  j2π(m−n)r/Ls  A2 = fm(x)  e dr dx (96)

−Lr/2+Ls/2 −Ls/2+x

L /2+L /2  L /2  r Z s Zr  j2π(m−n)r/Ls  A3 = fm(x)  e dr dx (97)

Lr/2−Ls/2 −Ls/2+x

1 −j2πmx/Ls with fm(x) = √ g(x)e . If n = m, then Ls

−Lr/2+Ls/2 Z L + L  A = f (x) s r + x dx (98) 1 m 2 −Lr/2−Ls/2 L /2−L /2 r Z s A2 = fm(x)Lsdx (99)

−Lr/2+Ls/2

Lr/2+Ls/2 Z L + L  A = f (x) s r − x dx. (100) 3 m 2 Lr/2−Ls/2

If n 6= m, then A2 = 0 and −L /2+L /2 rZ s A1 = bnm fm(x)cnm(x)dx (101)

−Lr/2−Ls/2 L /2+L /2 r Z s A2 = −bnm fm(x)cnm(x)dx (102)

Lr/2−Ls/2

L Ls m−n j2π(m−n) x −jπ(m−n) r with b = and c (x) = (−1) e Ls −e Ls . Notice that the above nm j2π(m − n) nm steps can also be used for the computation of the elements of R in (76), after taking into account that ψn(sx) must be replaced with φm(sx). 29

APPENDIX C

0 Define κz = κz − m−1−(N−1)/2 and rewrite the integral in (62) as follows 2π 2π Ls Z ∞   0 m − 1 − (N − 1)/2 0 0 Gz κz + 2π , d Qn−m(κz)dκz (103) −∞ Ls with  0   0   0 κz κz i Qi(κz) = sinc Ls sinc + Lr . (104) 2π 2π Ls By using Parseval, (103) becomes Z ∞ 0 j2π m−1−(N−1)/2 z0 0 0 gz (z , d) e Ls qn−m(z )dz (105) −∞ with 2 0 0 0 (2π)  z   z  −j 2π iz0 qi(z ) = rect ⊗ rect e Ls . (106) LsLr Ls Lr Assume now that the conditions in (70) and (71) are satisfied. In these circumstances, the integral in (105) reduces to

Z zg 0 j2π m−1−(N−1)/2 z0 0 0 gz (z , d) e Ls qn−m(z )dz . (107) −zg

2 For |z0| ≤ z , it is easy to show that q (x0) = (2π) for i = 0 and q (x0) = 0 for i 6= 0. Plugging g i Lr i this result into (107) produces

 2 n−1−(N−1)/2 (2π) R zg 0 j2π x0 0  gz (z , d) e Ls dz n = m Lr −zg (108)  0 n 6= m. By noticing that, under the condition in (71),

Z zg   0 j2π n−1−(N−1)/2 z0 0 2π gz (z , d) e Ls dz = Gz (n − N/2 − 1) , d (109) −zg Ls the orthogonality condition in (65) easily follows.

REFERENCES

[1] T. S. Rappaport, Y. Xing, O. Kanhere, S. Ju, A. Madanayake, S. Mandal, A. Alkhateeb, and G. C. Trichopoulos, “Wireless communications and applications above 100 ghz: Opportunities and challenges for 6G and beyond,” IEEE Access, vol. 7, pp. 78 729–78 757, 2019. [2] D. Tse and P. Viswanath, Fundamentals of Wireless Communication. Cambridge University Press, 2005. [3] P. Driessen and G. Foschini, “On the capacity formula for multiple input-multiple output wireless channels: a geometric interpretation,” IEEE Trans. Commun., vol. 47, no. 2, pp. 173–176, 1999. [4] J.-S. Jiang and M. Ingram, “Spherical-wave model for short-range MIMO,” IEEE Trans. Commun., vol. 53, no. 9, 2005. 30

[5] F. Bohagen, P. Orten, and G. E. Oien, “On spherical vs. plane wave modeling of line-of-sight MIMO channels,” IEEE Trans. Commun., vol. 57, no. 3, pp. 841–849, 2009. [6] E. Torkildson, U. Madhow, and M. Rodwell, “Indoor millimeter wave MIMO: Feasibility and performance,” IEEE Trans. Wireless Commun., vol. 10, no. 12, pp. 4150–4160, 2011. [7] H. Do, N. Lee, and A. Lozano, “Reconfigurable ULAs for line-of-sight MIMO transmission,” IEEE Trans. Wireless Commun., vol. 20, no. 5, pp. 2933–2947, 2021. [8] D. A. B. Miller, “Communicating with waves between volumes: evaluating orthogonal spatial channels and limits on coupling strengths,” Appl. Opt., vol. 39, no. 11, pp. 1681–1699, Apr 2000. [9] F. K. Gruber and E. A. Marengo, “New aspects of electromagnetic information theory for wireless and antenna systems,” IEEE Trans. Antennas Propag., vol. 56, no. 11, pp. 3470–3484, 2008. [10] M. A. Jensen and J. W. Wallace, “Capacity of the continuous-space electromagnetic channel,” IEEE Transactions on Antennas and Propagation, vol. 56, no. 2, pp. 524–531, 2008. [11] M. Franceschetti, Wave Theory of Information. Cambridge University Press, 2017. [12] A. Pizzo, L. Sanguinetti, and T. L. Marzetta, “Holographic MIMO communications,” CoRR, vol. abs/2105.01535, 2021. [Online]. Available: https://arxiv.org/abs/2105.01535 [13] D. Dardari and N. Decarli, “Holographic Communication using Intelligent Surfaces,” CoRR, vol. abs/2012.01315, 2020. Online: https://arxiv.org/abs/2012.01315. [14] C. Huang, S. Hu, G. C. Alexandropoulos, A. Zappone, C. Yuen, R. Zhang, M. D. Renzo, and M. Debbah, “Holographic MIMO surfaces for 6g wireless networks: Opportunities, challenges, and trends,” IEEE Wireless Communications, vol. 27, no. 5, pp. 118–125, 2020. [15] D. Slepian, “Prolate spheroidal wave functions, Fourier analysis and uncertainty: Extensions to many dimensions; generalized prolate spheroidal functions,” The Bell System Technical Journal, vol. 43, no. 6, pp. 3009–3057, 1964. [16] D. Dardari, “Communicating with large intelligent surfaces: Fundamental limits and models,” IEEE J. Sel. Areas Commun., vol. 38, no. 11, pp. 2526–2537, 2020. [17] S. Tretyakov, “Metasurfaces for general transformations of electromagnetic fields,” Philosoohical Trans. Royal Soc. A: Mathematical, Physical and Engineering Sciences, vol. 373, no. 2049, 2015. [18] Y. B. Li, B. G. Cai, Q. Cheng, and T. J. Cui, “Surface Fourier-transform lens using a metasurface,” Journal of Physics D: Applied Physics, vol. 48, no. 3, p. 035107, jan 2015. [Online]. Available: https://doi.org/10.1088/0022-3727/48/3/035107 [19] E. Telatar, “Capacity of multi-antenna gaussian channels,” European Trans. Telecommun., vol. 10, no. 6, 1999. [20] F. Bohagen, P. Orten, and G. Oien, “Construction and capacity analysis of high-rank line-of-sight MIMO channels,” in IEEE Wireless Commun. Netw. Conf., 2005, vol. 1, 2005, pp. 432–437 Vol. 1. [21] W. C. Chew, Waves and Fields in Inhomogenous Media. Wiley-IEEE Press, 1995. [22] A. S. Y. Poon, R. W. Brodersen, and D. N. C. Tse, “Degrees of freedom in multiple-antenna channels: a signal space approach,” IEEE Trans. Inf. Theory, vol. 51, no. 2, Feb 2005. [23] S. J. Orfanidis, Electromagnetic Waves and Antennas. [Online]. Available: http://www.ece.rutgers.edu/ orfanidi/ewa/, 2008. [24] A. Pizzo, T. L. Marzetta, and L. Sanguinetti, “Degrees of Freedom of Holographic MIMO Channels,” in 2020 IEEE 21st Int. Workshop Signal Process. Adv. Wireless Commun. (SPAWC), 2020, pp. 1–5. [25] W. Yu, W. Rhee, S. Boyd, and J. Cioffi, “Iterative water-filling for gaussian vector multiple-access channels,” IEEE Trans. Inf. Theory, vol. 50, no. 1, pp. 145–152, 2004.