You are on page 1of 18

electronics

Article
Capacity Performance Analysis for Terrestrial THz Channels
George K. Varotsos 1 , Konstantinos Aidinis 2,3 , Athanassios Katsis 4 and Hector E. Nistazakis 1, *

1 Section of Electronic Physics and Systems, Department of Physics, National and Kapodistrian University of
Athens, 15784 Athens, Greece; georgevar@phys.uoa.gr
2 Department of Electrical and Computer Engineering, Ajman University,
Ajman P.O. Box 346, United Arab Emirates; k.aidinis@ajman.ac.ae
3 Centre of Medical and Bio-allied Health Sciences Research (CMBHSR), Ajman University,
Ajman P.O. Box 346, United Arab Emirates
4 Department of Social and Educational Policy, University of the Peloponnese, 20100 Korinthos, Greece;
katsis@uop.gr
* Correspondence: enistaz@phys.uoa.gr; Tel.: +30-210-7276710

Abstract: The outdoor terrestrial terahertz (THz) communication links have recently attracted great
research and commercial interest in response to the emerging bandwidth-hungry demands for
extremely high-speed wireless data transmissions. However, their development is hindered by the
random behavior of the atmospheric channel due to the molecular attenuation, adverse weather
effects, and atmospheric turbulence (along with free space path loss (FSPL) and pointing errors) due
to the stochastic misalignments between the transmitter and the receiver. Thus, in this work, we
investigate the joint influence of these detrimental effects on both capacities, i.e., average (ergodic)
and outage, of such a typical line of sight (LOS) THz communication link. Specifically, atmospheric
turbulence-induced intensity fluctuations can be modeled by using either the suitable gamma or the
well-known gamma–gamma distribution for weak and moderate to strong turbulence conditions,
respectively. Additionally, weak to strong stochastic misalignment-induced intensity fluctuations,
due to generalized pointing errors with non-zero boresight (NZB), are emulated by the appropriate
Beckman distribution. Taking into additional consideration the unavoidable presence of FSPL and the
different but realistic water vapor concentration values along with the influence of weather conditions,
an outage performance analysis has been conducted. Considering the abovementioned significant
effects, novel analytical mathematical expressions have been extracted for both average (ergodic)
Citation: Varotsos, G.K.; Aidinis, K.; and outage capacity, which are critical metrics that first incorporate the total influence of all of the
Katsis, A.; Nistazakis, H.E. Capacity above significant effects on the THz links’ performance. Through the derived expressions, proper
Performance Analysis for Terrestrial analytical results verified by simulations are presented and demonstrate the validity of our analysis.
THz Channels. Electronics 2023, 12, It is notable that the derived expressions can accommodate realistic parameter values involved in all
1336. https://doi.org/10.3390/ the above-mentioned major effects and link characteristics. In this context, they provide encouraging
electronics12061336
quantitative results and outcomes for both capacity metrics under investigation. The latter enables
Academic Editor: Reza K. Amineh the design and the establishment of modern and future high-speed THz links, which are expected to
cover longer propagation distances and thus become even more vulnerable to atmospheric turbulence
Received: 14 February 2023
effect. This is modeled and incorporated in our analysis and expressions contrary to most of the
Revised: 8 March 2023
Accepted: 10 March 2023
previous works in the open technical literature.
Published: 11 March 2023
Keywords: THz links; atmospheric turbulence; attenuation; weather conditions; path losses; generalized
pointing errors; average capacity; outage capacity

Copyright: © 2023 by the authors.


Licensee MDPI, Basel, Switzerland.
This article is an open access article 1. Introduction
distributed under the terms and
Over the last few years, the development of terahertz (THz) wireless communication
conditions of the Creative Commons
Attribution (CC BY) license (https://
links has been the focus of significant research and commercial efforts. This is mainly
creativecommons.org/licenses/by/
due to their potential to meet the increasing bandwidth-consuming demands for higher
4.0/). data rate wireless transmissions at a higher security level as well as the extremely large

Electronics 2023, 12, 1336. https://doi.org/10.3390/electronics12061336 https://www.mdpi.com/journal/electronics


Electronics 2023, 12, 1336 2 of 18

amount of unlicensed spectra that accommodate the THz band, which remains one of
the least explored frequency bands for communication applications [1–4]. Contrary to
its immediately below millimeter wave (MMW) frequency band (which has been lately
utilized for fifth generation (5G) mobile communication systems), THz systems offer higher
capacity in links. Indeed, they have a larger atmospheric transmission window and thus
possess the potential to achieve data rates beyond several Gbps and reach capacities
of 100 Gbps to 1Tbps, which is sufficient to cover the growing sixth generation (6G)
future demands [5–7]. The latter is a vital issue, considering that the global mobile traffic
volume was 7.462 EB/month in 2010, and it is predicted to be 5016 EB/month in 2030 [8].
Additionally, the THz frequency spectrum is largely unregulated in contrast with the MMW
frequency spectrum. Furthermore, THz systems outperform their MMW counterparts in
terms of more compact transceiver antenna equipment and inherently more directional
links owing mainly to less free-space diffraction of the waves [9]. On the other hand,
immediately above THz frequency band there is the infrared (IR) frequency band where
optical wireless communication (OWC) links, commonly known as free space optical
(FSO) links, usually operate [4,10]. Contrary to their FSO counterparts, THz have been
experimentally proven to be less susceptible to signal scintillation and attenuation effects
resulting from the presence of atmospheric turbulence, fog, dust, smoke, and/or clouds,
respectively [11–14]. Moreover, contrary to FSO, the emitted power and, as a result, the
performance and the availability of THz links is not degraded by any eye-safety limitations;
plus, THz are immune to ambient and infrared light noise [15]. In view of the above distinct
advantages of THz links, the latter can potentially adapt varying and vital applications such
as high-speed transmission of high-definition television (HDTV) signals, health monitoring,
high-speed sensor networks, secure terabit wireless communication in the military field,
terabit wireless personal area networks, terabit wireless local area networks, solution for
backhaul or bottleneck problem, machine-to-machine communication, and augmented
reality (along with virtual reality [4,6–8]).
Nevertheless, regarding the establishment of outdoor THz communication systems,
there are some major impairments that mainly arise from the random nature of the atmo-
spheric channel. Still, airborne particles that emerge across the atmospheric channel during
adverse weather effects that attenuate the propagating THz signals, with rain droplets
being by far the most significant attenuators. Indeed, conversely to fog droplets, rain
attenuates more THz radiation than IR light [16–18]. Additionally, molecular absorption
that basically stems from oxygen and especially water vapor in the air channel plays a dom-
inant detrimental role in the propagation of THz information-bearing waves. Specifically,
from a communication perspective, molecular absorption adds a margin of noise, which
limits the communication range [19]. Regarding the total attenuation that experiences the
propagating THz signal through the atmospheric channel, there is also an unavoidable loss
that always exists, commonly known as free space path loss (FSPL) [20].
Even in a clear sky, atmospheric turbulence originates from solar radiation absorbed
by Earth’s surface that makes the air around the earth surface to be warmer than that
at higher altitude. This complex stochastic phenomenon causes random refractive index
fluctuations, which in turn create random intensity variations of the received THz signal, the
so-called scintillation effect analogous to fading in RF systems, and results in performance
degradation of the link (especially for long propagation distances [20–27]).
Another significant limitation factor for THz performance is the pointing errors (PE)
effect between transceiver apertures. This effect causes random intensity fluctuations of the
THz signal, which arrives at the receiver’s input [2,28–30]. It should be mentioned here
that this effect includes boresight and jitter components. The boresight is a permanent
displacement between the beam and the detector aperture, while the spatial jitter is the
random offset of the beam center at the receiver [24,31].
Unlike the case with FSO communication systems [28,32–40], the joint impact of most
of the above critical effects on the capacity performance and availability of THz links has
not yet been extensively investigated. In more detail, the feasibility of establishing practical
Electronics 2023, 12, 1336 3 of 18

terrestrial outdoor line of sight (LOS) THz communication links under various weather and
channel conditions has been recently shown in [41–43]. Note that during the previous years
the poor development of THz devices up to this time dramatically limited the achievable
outdoor THz link lengths from just 0.5 m up to 22 m operating at 300 GHz. However,
owing to the impressive development of THz devices that has been witnessed just over
the last few years, the practical achievable outdoor THz propagation distance has been
extended in acceptable and remarkable ways for communication practical link lengths,
i.e., from 186 m up to 1 km [41–43]. In this respect, we don’t extensively refer here to the
earlier background, i.e., to what happened on the development of THz links before the
publication of papers [41–43]. Furthermore, relying mainly on the molecular absorption
loss model proposed in [44], authors in [45] suggested another THz link incorporating
losses due to the propagation path as well as critical parameters concerning the transceivers
and molecular absorption. According to this model, the capacity up to 0.4 THz is evaluated.
In these papers it is also validated that, for such THz transmissions, attenuation caused by
the water vapor is more important than the attenuation caused by the oxygen, and thus
the focus on the dominant molecular attenuation factor within the frequency regime that
emerging THz systems operate. Attenuation of THz radiation (which is caused by fog,
clouds, or rain) has been investigated by means of laser environmental effects definition
and reference (LEEDR) computation tool in [46], whereas in [7,13,14,16–18,42,43,47], the
same effects have been studied theoretically and/or experimentally. Their findings indicate
a good agreement between the expected results and the measurements and it is shown that
the rain has the most significant negative influence on the THz links’ performance and
availability. In this context, it has been shown that the attenuation caused by the weather
effects should be considered mainly for long THz links, whereas it has been verified that,
under fog conditions, THz outperforms FSO, which makes THz a viable alternative to FSO
(especially in foggy channels). Moreover, atmospheric turbulence-induced scintillation
effects on THz propagating signals have been initially theoretically investigated via the
scintillation index metric in [48], and then experimentally in [14,49]. Their results imply
that, while atmospheric turbulence is a more significant issue for IR than for THz links,
it should be taken into consideration especially for the emerging near the ground THz
links that cover longer propagation distances. Considering that atmospheric turbulence
causes rapid fluctuations in the signal’s intensity, it should be described by using the
appropriate statistical models. In this respect, authors in [20,21], used the well-known
distribution models from the FSO systems for the THz links, i.e., lognormal, gamma–
gamma, etc. Each one of the above-mentioned models is suitable to describe the influence
of the atmospheric turbulence effect depending on its strength. More recently, authors
in [50] proposed gamma distribution for weak turbulence-induced scintillations as a less
complex alternative to the lognormal one to estimate the outage probability of a typical THz
link. The PE effect between the transmitter and the receiver of the THz link devices was
first studied as a part of the shadowing effect in [51], and next in [52,53] by deterministic
models; however, these models are not able to accommodate their stochastic nature. The
stochastic behavior of PE for THZ links has been recently assessed in [1,2,54] relying mainly
on the Rayleigh distribution model [24], which has been used for FSO communication
links. Thus, the ergodic (average) capacity has been evaluated in [2] under the presence of
stochastic zero boresight pointing errors, FSPL, molecular attenuation, multipath fading
along with hardware imperfections for different THz link propagation distances, relative
humidity, operation frequency, signal to noise ratio (SNR), and spatial jitter values. Next,
authors in [1] investigated the total impact of stochastic zero boresight PE along with FSPL,
molecular attenuation, and different THz link and antenna characteristics on the THz
average capacity by utilizing three different movement models. The joint impact on the
average LOS THz capacity of the atmospheric turbulence effect along with zero boresight
stochastic pointing errors (also including FSPL and molecular attenuation) has been first
evaluated in [20]. Still, the outage capacity metric has not been reported for LOS THz links
in the open technical literature as well as the joint influence of atmospheric turbulence,
Electronics 2023, 12, 1336 4 of 18

generalized PE with NZB, FSPL, and molecular attenuation along with different weather
effects on their average capacity have not yet been investigated.
Motivated by the above, in this work we investigate, for first time, both outage and
average capacity metrics for a typical LOS THz link under the combined influence of
atmospheric turbulence, generalized NZB PE, FSPL, and molecular attenuation caused by
the water vapor along with clear, fog or rain weather conditions. By emulating turbulence-
induced scintillations by gamma or gamma–gamma distributions for weak, moderate, or
strong turbulence conditions and generalized pointing errors via Beckmann distribution,
an outage and average capacity analysis is performed. In this context, another key novelty
of this work is that, for the first time in the open THz literature, closed form mathematical
expressions are derived either for the outage capacity or for the average capacity of a typical
LOS THz link, including the joint influence of all these above-mentioned major factors that
affect the THz capacity performance and availability. Additionally, for realistic THz link
scenarios, we present proper analytical results obtained via the extracted expressions, which
are further verified by simulations. Specifically, these are Monte Carlo simulations, which
have been performed with the help of well-known mathematical software. Consequently,
insightful observations are provided for the influence of each major factor (much less for
their total impact on the link’s capacity availability and performance), which can be used
for the design of THz links.

2. Signal and Channel Model


2.1. The THz Link
The LOS THz link under investigation is assumed to employ the on-off keying (OOK)
modulation format, which is well-known for industrial and commercial applications mainly
due to its simplicity. Thus, the arriving THz signal at the receiver is expressed as

y = hx + n (1)

with h denoting the channel state due to turbulence, generalized pointing errors, and total
signal attenuation, x = {0 or 2Pt } being the emitted OOK information signal where Pt
represents the average transmitted signal power, and n standing for the corresponding
additive Gaussian white noise with variance σn2 [20].
The above coefficient, h, can be expressed as

h = hl h p h a (2)

where the signal’s intensity of the THz channel is obtained through the total attenuation
(hl ), the pointing errors (h p ), and the turbulence effect (h a ) [25,28,50].

2.2. Total Attenuation


Furthermore, the factor hl , can be written as

hl = h f l hvl hwl (3)

where the total attenuation of the THz channel depends on the FSPL (h f l ), the molecular
attenuation (hνl ), and the weather effects attenuation (hwl ). Note that the hwl is taking
various values fog or rain, while it is equal to one for clear weather conditions as has
been already reported in [2,20,50], where the presence of any adverse weather effect has
been neglected.
Specifically, the attenuation term due to FSPL is described by Friis equation as [54]

c Gt Gr
hfl = (4)
4π f z
Electronics 2023, 12, 1336 5 of 18

with c being the light’s speed in the free space, f being the operation frequency of the
information-bearing THz carrier wave, z standing for the length of the link, and Gt , Gr
representing the transmission and reception antenna gains, respectively.
According to the Beer–Lambert law, the molecular attenuation factor due to the signal’s
absorption along the propagation path is given as [11]

hvl = exp(− av z) (5)

where aν denotes the attenuation coefficient (m−1 ), with the link length of the THz channel,
z, being expressed in meters. Taking into consideration that water vapor is the dominant
attenuation factor for the outdoor LOS THz links at low altitudes [6,9,55], it is assumed
without loss of generality that aν coincides with the attenuation coefficient due to water
vapor, which at T0 = 20 ◦ C surface temperature and f ≤ 350 GHz is given as [[56],
Equation (3.2)]
" #
2.4 7.33 4.4
av (dB/km) = 0.067 + + + f 2 ρ × 10−4 , (6)
( f − 22.3)2 + 6.6 ( f − 183.5)2 + 5 ( f − 323.8)2 + 10
where the frequency f should be now expressed in GHz while ρ represents the water
vapor concentration in g/m3 . It is worth noting here that it is feasible to extend Equation (6)
for different surface temperature values, i.e., the lower the surface temperatures, the higher
the attenuation coefficient. Specifically, for every Celsius degree, the coefficient increases
by about 1% [56].
In the same context, the attenuation term, hwl , is given as

hwl = exp(− aw z), (7)

where aw denotes the experimental coefficient of attenuation due to weather effect (in
km−1 ) and the propagation distance z (in km). Furthermore, it is worth mentioning that the
attenuation due to adverse weather effect can be added with the FSPL and the molecular
attenuation [20]. Here, therefore, by using the above presented expressions, the total
attenuation will be jointly evaluated.

2.3. Atmospheric Turbulence Models


In this work we use two different distribution models to study weak, moderate, or
strong turbulence conditions. For weak THz turbulent channels we use the more compact
gamma distribution, while for moderate to strong turbulence we resort to the more complex
but very accurate gamma–gamma model.
The probability density function (PDF) for the gamma–gamma distribution of the
random variable h a is given as [57]
α+ β
2(αβ) 2 α+2 β −1  p 
f GG,ha (h a ) = ha Kα− β 2 αβh a , (8)
Γ(α)Γ( β)

with Kν (.) standing for the ν-th order modified Bessel function of the second kind [58],
(8.432.2), Γ(.) denoting the gamma function [58], (8.310.1), while α and β depend on the
atmospheric channel and the link’s characteristics, respectively, as [59]:
    −1
0.49δ2
α = exp 7/6 −1
(1+0.18d +0.56δ12/5 )
2
  −5/6   −1 (9)
0.51δ2 (1+0.69δ12/5 )
β = exp 5/6 −1 ,
(1+0.9d2 +0.62d2 δ12/5 )
Electronics 2023, 12, 1336 6 of 18


where d = 0.5D 2πλ−1 z−1 , with D being the receiver’s aperture diameter and λ = c/ f
denoting the operational wavelength, while the parameter δ2 stands for the Rytov variance,
which, for wave propagation, is given as [60,61]
7 11
δ2 = 0.5Cn2 k 6 z 6 (10)

where k = 2π/λ is the wavenumber and Cn2 represents the refractive index structure
parameter, which is proportional to atmospheric turbulence strength [14,16,60,61].
The PDF of the gamma distribution model, which focuses on weak turbulent channels,
is given as [22]
ζ ζ ζ −1
f G,ha (h a ) = h a exp(−ζh a ), (11)
Γ(ζ )
where ζ is the gamma distribution parameter, which is related, in turn, to the parameters α
and β as follows [34,50]
1 1 1 −1
 
ζ= + + , (12)
α β αβ

2.4. Generalized Pointing Errors Model with NZB


The stochastic nature of generalized PE incorporating the boresight part is accurately
described by Beckmann statistical distribution model. Its PDF of the radial displacement at
the receiver θ is given as [23,24,31]:
2 !
(θcosϕ − µ x )2
Z 2π
θ θsinϕ − µy
f θ (θ ) = × exp − 2
− 2
dϕ, (13)
2πσx σy 0 2σx,n 2σy,n

with ϕ representing the divergence angle related to the increase of the beam radius at
the receiver with distance z from the transmitter’s side, while the beam width
q can be
approximated as wz ≈ ϕz for long propagation distances. Moreover, θ = θ x2 + θy2
where θ x2 and θy2 are the offsets along the horizontal and elevation axes while θ x and
θy are considered asnonzero mean Gaussian random variables, i.e., θ x ∼ N µ x , σx2


and θy ∼ N µy , σy2 , with µ x and µy denoting the mean values, whereas σx and σy
represent the standard deviations for horizontal and elevation displacements, respectively.
Considering [62], the PDF of (13) can be simplified through the modified Rayleigh model:
!
θ θ2
f θ (θ ) = 2 exp − 2 , θ ≥ 0, (14)
σmod 2σmod

where σmod denotes the joint standard deviation of σx and σy , which can be expressed
as [23]
!1
2
3µ2x σx4 + 3µ2y σy4 + σx6 + σy6 3
σmod = , (15)
2

Hence, according to [23,62], the PDF of h p is given as

ψ2 ψ2 −1
h p , 0 ≤ h p ≤ A0 g, (16)

fhp hp = 
2
A0 g ) ψ

where ψ = wz,eq /2σmod is directly related to the total generalized pointing error at the
receiver’s input. In fact, greater values of ψ correspond to weaker PE effect [24]. In the same
context, ψx = wz,eq /2σ x and ψy = wz,eq /2σy refer to the specific generalized misalignment
strengths along the horizontal and the elevation axis, respectively. It is to be mentioned that
√ 1/2
wz,eq = πerf(v)w2z /2vexp −v2 , where A0 = erf2 (v) stands for the fraction of the
Electronics 2023, 12, 1336 7 of 18

collected power at r = 0, with erf(.) denoting for the error function [63]√Equation
√ (7.1.1) and
r representing the radius of the aperture of the receiver, while v = πr/ 2wz , with wz
denoting the beamwaist on the receiver plane at propagating distance z [20,28]. Next, the
µ2y

1 1 1 µ2x,n
parameter g = exp ψ2
− 2ψx2
− 2ψy2
determines the presence of boresight
− 2σx2 ψx2
− 2σy2 ψy2
q
component. In this respect, when the boresight displacement s = µ2x + µ2y is equal to
zero, i.e., µ x = µy = 0, and σx = σy , then g = 1. Consequently, for g = 1 we have only
ZB pointing errors and thus the Beckmann model simplifies to a Rayleigh distribution
with s = 0. Therefore, in this case, Equation (14) reduces to [28], Equation (10) while
Equation (16) also reduces to [28] and Equation (11).

2.5. Joint Influence of Effects


The PDF of the total channel state, considering the joined influence of the above
presented effects, is expressed as:
Z
f h (h) = f h|ha (h| h a ) f ha (h a )dh a , (17)

where, for the sake of brevity, f ha (h a ) denotes either the gamma–gamma modeled PDF or
the gamma modeled PDF through Equation (8) and Equation (11), respectively. In this
context, f h|ha (h| h a ) is the corresponding conditional probability of random variable h given
h a , while it can be expressed as [28]

dFh p (h| h a hl )
f h|h a ( h | h a ) = dh
2 2  ψ2 −1 (18)
( A0 g ) − ψ ψ

h
= h a hl h a hl , 0 ≤ h ≤ A0 gh a hl ,

with F (.) denoting the cumulative distribution function (CDF).


By following the analysis performed in [22], after substituting Equations (8) and
(18) into Equation (17) and Equations (11) and (18) into Equation (17) for gamma–gamma
modeled turbulence and for gamma modeled turbulence, respectively, and by then utilizing
Equation (3) we correspondingly obtain:

αβψ2
f GG,h (h) = A0 gh f l hvl hwl Γ(α)Γ( β)

ψ2
 (19)
3,0 αβh
× G1,3 ,
A0 gh f l hvl hwl ψ2 − 1, α − 1, β − 1

and
ζψ2
f G,h (h) = A0 gh f l hvl hwl Γ(ζ )

ψ2
 (20)
3,0 ζh
× G1,3 ,
A0 gh f l hvl hwl ψ2 − 1, ζ − 1
where G m,n
p,q (.) denotes the Meijer G-function [64], Equation (5).

2.6. SNR Expressions


The instantaneous signal to noise ratio (SNR) for the THz link under investigation is
defined as [20]
2P2 h2
γ = t2 (21)
σn
Additionally, the corresponding average SNR is defined as [20],

2Pt2 ( E[ h])2
µ= , (22)
σn2
Electronics 2023, 12, 1336 8 of 18

R∞
where E[ h] = 0 h f h (h)dh, with E[.] denoting the statistical expectation. By substituting
Equations (8) and (11) into the latter integral for the gamma–gamma and gamma distri-
bution models, respectively, and by following next the analysis performed in [65], along
with using Equation (3), the latter integral gives for both distribution models the following
solution [25,34,35]:
  −1
E[ h] = h f l hvl hwl A0 g 1 + ψ−2 (23)

Considering Equations (21)–(23) and Equations (8) and (11) for gamma–gamma and
gamma distributions respectively, along with applying standard technique of transform-
f h (h)
ing random variables f γ (γ) =  −1 q
γ [35], we obtain their
∂γ
∂h
A0 g 1 + ψ −2 h f l hvl hwl µ
corresponding PDFs of random variable γ after some algebraic manipulations as follows:

αβψ2
f GG,γ (γ) = 2Γ(α
)Γ( β)γ

1 + ψ2 (24)

3,0 αβ γσn
× G1,3 √ ,
2Pt A0 gh f l hvl hwl ψ2 , α, β

and
ψ2
f G,γ (γ) = 2Γ(ζ


1 + ψ2 (25)

2,0 ζ γσn
× G1,2 √ .
2Pt A0 gh f l hvl hwl ψ2 , ζ

3. Average Capacity Analysis


The instantaneous channel capacity C, which is a very critical metric for the perfor-
mance of any communication system, is defined as [34,66]

C = Blog2 (1 + γ), (26)

where B stands for the channel’s bandwidth.


Consequently, according to Equations (21) and (26), the average capacity can be
expressed through the following integral [36,39]:

1 2P2 h2
Z ∞  
Cav = Bln 1 + t 2 f h (h)dh, (27)
ln2 0 σn

By substituting Equation (8) for gamma–gamma modeled turbulence and Equation (8)
for gamma modeled turbulence into Equation (27), the latter correspondingly gives

αβψ2 B
CGG,av = ln2A0 gh f l hvl hwl Γ(α)Γ( β)

2Pt2 h2
 
ψ2
 (28)
R∞ 3,0 αβh
× 0 ln 1 + σn2
G1,3 2 dh,
A0 gh f l hvl hwl ψ − 1, α − 1, β − 1

and
ζψ2 B
CG,av = ln2A0 gh f l hvl hwl Γ(ζ )

2Pt2 h2
 
ψ2
 (29)
R∞ 2,0 ζh
× 0 ln 1 + σn2
G1,2 2 dh.
A0 gh f l hvl hwl ψ − 1, ζ − 1
1, 1
 
1,2
Bearing in mind that ln(1 + x ) = G2,2 x [64], Equation (11) and Equations (28)
1, 0
and (29) can be written as

αβψ2 B
CGG,av = ln2A0 gh f l hvl hwl Γ(α)Γ( β)

ψ2
 
2Pt2 h2 1, 1
 (30)
R∞ 3,0 αβh 1,2
× 0 G1,3 G2,2 dh,
A0 gh f l hvl hwl ψ2 − 1, α − 1, β − 1 σn2 1, 0
Electronics 2023, 12, 1336 9 of 18

and
ζψ2 B
CG,av = ln2A0 gh f l hvl hwl Γ(ζ )

ψ2
 
2Pt2 h2 1, 1
 (31)
R∞ 2,0 ζh 1,2
× 0 G1,2 2 G2,2 σn2
dh.
A0 gh f l hvl hwl ψ − 1, ζ − 1 1, 0
By next applying [64], Equation (21) to solve the integrals of Equations (30) and (31),
we correspondingly obtain

2α + β −3 ψ2 B
CGG,av = πln2Γ(α)Γ( β)
1− ψ2 2− ψ2 1− α 2− α 1− β 2− β (32)
!
2
1,8 32( Pt A0 gh f l hvl hwl ) 1, 1, 2 , 2 , 2 , 2 , 2 , 2
× G8,4 2 − ψ2 1− ψ2
,
(αβσn ) 1, 2 , 2 , 0

and
ζ −2 2
CG,av = √2 ψ B
πln2Γ(ζ )
1− ψ2 2− ψ2 1− ζ 2− ζ (33)
!
2
1,6 8( Pt A0 gh f l hvl hwl ) 1, 1, 2 , 2 , 2 , 2
× G6,4 − ψ2 1− ψ2
,
(ζσn )2 1, 2 , 2 , 0

where we can set CBav standing for the average capacity over channel’s bandwidth, expressed
in (b/s/Hz) [32].
It therefore becomes evident that the extracted average capacity closed-from expres-
sions (32), and (33) include the joint influence of generalized pointing errors with NZB, the
FSPL, the water vapor attenuation, and the potential emergence of adverse weather attenu-
ation, along with gamma–gamma or gamma modeled turbulence-induced scintillations,
respectively, on the average THz capacity performance.

4. Outage Capacity Analysis


A very crucial outage performance metric for the availability of any communication
system is its outage capacity, Cout , which refers to the capacity guaranteed for a percentage
rate of (100 − R%) or, equivalently, (1 − R), of the channel realizations [39,66,67]

Pr[C < Cout ] = R (34)

with Pr[.] denoting probability.


Considering Equations (21) and (26), it is deduced that, as is the case with the instanta-
neous electrical SNR, the instantaneous capacity is a random variable too. Therefore, the
probability defined by Equation (34) is evaluated either for gamma–gamma or for gamma
modeled turbulence as [34,39]
Z Cout
R= f C (C )dC (35)
0
with f C (C ) representing the PDF of the random variable C either under gamma–gamma
modeled turbulent or gamma turbulent channels, respectively.
Considering Equations (21)–(23) and (26) and Equations (24) and (25) for gamma–
gamma and gamma distributions respectively, along with utilizing standard technique of
f γ (γ)
transforming random variables f C (C ) = C [35], we obtain their corre-
γ = 2B − 1 ∂C
∂γ
sponding PDFs of random variable C (after some algebraic manipulations) as follows:
  −1
C C
ψ2 2 B 2 B −1 ln2
f GG,C (C ) = 2Γ(α)Γ( β) B
q ! (36)
1 + ψ2
C
3,0 √ αβσn 2 −1
B
× G1,3
2Pt A0 gh f l hvl hwl ψ2 , α, β
Electronics 2023, 12, 1336 10 of 18

and   −1
C C
ψ2 2 B 2 B −1 ln2
f G,C (C ) = 2Γ(ζ ) Bq ! (37)
C
1 + ψ2
2,0 √ ζσn 2 −1
B
× G1,2
2Pt A0 gh f l hvl hwl ψ2 , ζ

Next, by substituting Equations (36) and (37) into Equation (35), we obtain the fol-
lowing expressions for the latter under gamma–gamma turbulent and gamma turbulent
channels, respectively:

ψ2 ln2
RGG = 2Γ(α)Γ( β) B
(38)
q !
ψ2 + 1
 −1 C
√ αβσn 2 −1
C
 C
R Cout 3,0 B
× 2 B 2 −1
B G1,3 dC,
0 2Pt A0 gh f l hvl hwl ψ2 , α, β

and
ψ2 ln2
RG = 2Γ(ζ ) B
(39)
q !
  −1 C
ψ2 + 1
√ ζσn 2 −1
R Cout C C 2,0 B
× 2 B 2 −1
B G1,2 dC.
0 2Pt A0 gh f l hvl hwl ψ2 , ζ

q In order to solve the latter integrals of Equations (38) and (39), we initially set u =
C
C 2qB ln2
2 − 1, and, thus, du =
B
C
dC. Next, by utilizing [64] and Equation (26) and after
2B 2 B −1
some algebraic manipulations, we correspondingly obtain
 q 
Cout
ψ2 αβσn 2 B − 1 1, ψ2 + 1 
RGG = G3,1  √ (40)
Γ(α)Γ( β) 2,4 2Pt A0 gh f l hvl hwl ψ2 , α, β, 0

and  q 
Cout
ψ2 2,1  ζσn 2 B − 1 1, ψ2 + 1 
RG = G √ (41)
Γ(ζ ) 2,3 2Pt A0 gh f l hvl hwl ψ2 , ζ, 0

where CBout stands for the outage capacity over channel’s bandwidth, expressed in (b/s/Hz),
similar to what is shown above [33].
It is thus notable that the derived outage capacity closed-form expressions (40) and (41)
incorporate the combined impact of generalized pointing errors with NZB, the FSPL, the
water vapor attenuation, and the potential emergence of adverse weather attenuation, along
with gamma–gamma or gamma modeled turbulence-induced scintillations, respectively,
on the THz outage capacity availability for every value of the probability R.

5. Analytical Results
In this section, by mainly utilizing the above derived mathematical expressions (32),
(33), (40), and (41), proper analytical results are presented, which are very insightful in
terms of average and outage capacity performance and availability metrics for the exam-
ined THz link. Specifically, the typical LOS THz link under investigation is z = 150 m in
length, operating with a carrier frequency equal to f = 0.3 THz. Regarding the transceiver
antenna characteristics, both transmitter and receiver have aperture radius r = 0.15 m
along with gain G = Gt = Gr = 55 dBi [20,50,68]. The Cn 2 parameter is assumed to be
Cn 2 = 5 × 10−10 m−2/3 for weak or Cn 2 = 2.3 × 10−9 m−2/3 for strong turbulence, as is the
case with the experimental results conducted in [14]. Unless otherwise stated, gamma distri-
bution is utilized for the weak turbulent channel case by means of Equations (32) and (40),
while the gamma–gamma model is utilized for the strong turbulent channel case via
Equations (33) and (41) for the average capacity and the outage capacity estimation, respec-
tively. In addition, varying practical scenarios for stochastic generalized pointing errors
Electronics 2023, 12, 1336 11 of 18

have been considered, i.e., w/r = 9 with µ x /r, µy /r,  σx /r, σy /r = (0, 0, 7, 7) for strong


ZB pointing errors as well as µ x /r, µy /r, σx /r, σy /r = (3, 1, 6, 5) or (3, 1, 7, 7) for weak to
strong NZB PE, respectively [62]. Moreover, according to realistic channel and environmen-
tal circumstances, it is assumed that the surface temperature is 20 ◦ C and the air pressure is
1 atm, while ρ = 7.5 g/m3 or ρ = 10 g/m3 for moderate to strong molecular attenuation
due to water vapor along the THz link, respectively [2,20,56]. Regarding the impact of
weather effects, apart from the clear sky scenario along the THz channel, two other weather
cases are investigated. In fact, the influence of fog attenuation is aw = 0.6 dB/km, whereas
the influence of rain attenuation is aw = 3 dB/km at a precipitation rate of 2 mm/h [7,46],
(Figure 1). Considering expression (6) in each weather scenario under investigation at-
tenuation due to vapor dominates, while attenuation due to rain droplets is expected to
increase to a greater extent the total attenuation than attenuation due to fog droplets. Note
also that σn = 10−7 A/Hz in all the above-mentioned link scenarios [20,50]. Finally, for the
outage capacity estimation scenarios under investigation we have set R = 0.1 which is very
common for any practical wireless communication link [33–35,66,67,69].

Figure 1. Average capacity (Cav /B) versus transmitted power for strong turbulence, moderate water
vapor concentration, and strong PE.

Figure 1 illustrates the average capacity evolution for the investigated THz link versus

the transmitted power through a strong gamma–gamma modeled turbulent channel with
either clear, rain, or fog weather conditions and moderate air humidity along with the
presence of FSPL and strong ZB or NZB PE. Thus, for the same transmitted power, channel
attenuation and zero boresight pointing errors, the average capacity for the THz link is
significantly degraded under rain compared to fog and, especially, to a clear weather
scenario. Indeed, slightly decreased average capacity values for the THz link are depicted
by comparing clear to fog weather scenarios, whereas significantly decreased average
capacity values are illustrated by comparing fog to rain weather scenarios, much less so
clear to rain weather scenarios. This average capacity performance comparison between
the three different weather scenarios (clear sky, fog, and rain) highlight that rain is the‐
most detrimental weather effect for the propagation of THz signals, which is consistent
Electronics 2023, 12, 1336 12 of 18

with the experiments in [11,12,14,16–18]. It is also notable that, by focusing on the rain
weather scenario, even lower average performance values are depicted as considering NZB
pointing errors (dashed line), and, therefore, the need for also considering the boresight
component on the average capacity performance estimation of the THz link is revealed.
Furthermore, for every illustrated channel scenario, increased transmitted power levels
lead to higher average capacity values, as it was expected.
Figure 2 is devoted to the clear sky scenario and visualizes the average capacity
evolution for the investigated THz link over the same transmitted power, FSPL, and the
same strong amount of NZB pointing mismatch along with weak to strong turbulent
channel conditions as well as moderate to strong air humidity. Note that, regarding
Figure 2, weak turbulence is pictured with dashed lines and strong turbulence is pictured
with solid lines, while moderate water vapor concentration is pictured with square markers
and strong water vapor concentration is pictured with triangle markers. Under these
circumstances, it is observed that the case of weak turbulence along with moderate water
vapor concentration (dashed line with square markers) achieves the best average capacity
performance. By maintaining the moderate water vapor concentration value, we observe
that the increase of atmospheric turbulence from weak to strong slightly but not negligibly
decreases the average capacity of the link. On the other hand, by maintaining the weak
turbulence value, we observe that the increase of water vapor concentration from moderate
to strong significantly decreases the average capacity of the link. In other words, it is
shown that the influence of the water vapor attenuation is more significant than the
detrimental impact of turbulence on the average capacity of THz links; however, the
atmospheric turbulence effect should not be neglected, since it further degrades the air
humidity-induced average capacity degradation. It is remarkable that the latter behavior
is qualitatively in a good agreement with experimental results in [11,14] that highlight
that, for THz signal transmissions, molecular attenuation is far more destructive than
atmospheric turbulence-induced scintillations, which should not be omitted in their own
right. Moreover, as it has been mentioned above, for weak turbulence (dashed lines)
we have selected gamma distribution, whereas for strong turbulence (solid lines), we
have selected the gamma–gamma distribution. In fact, gamma–gamma distribution could
have very well been utilized for weak turbulence as well. A representative example of
the latter is that, for the most average capacity performance effective case, when the
transmitted power is equal to 5 dBm, it is depicted an average capacity value equal to
9.000135124 b/s/Hz through gamma distribution, while its corresponding value would be
equal to 9.003206569 b/s/Hz through gamma–gamma distribution. Consequently, gamma
distribution is accurate enough for weak turbulence, which is also consistent with the
results in [50] and has been selected due to its simplicity compared to the more complex
gamma–gamma distribution. Finally, once again, even in the worst depicted case (strong
water vapor concentration along with strong turbulence) the average capacity increases by
increasing the transmitted power.
Figure 3 presents the outage capacity of the investigated THz link versus the transmit-
ted power through a strong gamma–gamma modeled turbulent channel with either rain or
fog adverse weather effects, moderate or strong water vapor concentration, and FSPL and
weak NZB PE. As is the case with the average capacity metric (where higher values lead
to enhanced performance results for the THz link), higher outage capacity metric values
lead to improved availability results for the THz link. It is therefore mainly shown that
more important outage performance degradations are observed for rain weather conditions
along with stronger water vapor concentrations, especially for lower transmission power
levels. Once again (but in terms of outage capacity availability this time), it is validated
that rain is a more destructive effect than fog for THz signal transmissions. Additionally,
it is highlighted that, even in rain conditions, the increase in air humidity along the THz
channel from moderate water vapor concentration (dashed line) to strong water vapor
concentration (solid line) strong air humidity (solid lines) remains a major limiting factor
in the THz performance in terms of outage capacity metric.
9.003206569 b/s/
Hz

Electronics 2023, 12, 1336 13 of 18

Figure 2. Average capacity (Cav /B) versus transmitted power under clear weather, weak to strong
turbulence, moderate to strong water vapor concentration, and strong NZB PE.

Figure 3. Outage capacity (Cout /B) versus transmitted power, strong turbulence, moderate to strong
water vapor concentration, weak NZB PE, and R = 0.1.

Figure 4 illustrates the outage capacity evolution for the investigated THz link over
the same wide range of transmitted, identical rain or fog adverse weather effects, identical
moderate or strong water vapor concentration along with the presence of the same FSPL,

‐ ‐
Electronics 2023, 12, 1336 14 of 18

and the same amount of weak NZB pointing mismatch (but over a weak gamma modeled
turbulent channel). The performance comparison between the two last figures therefore
reveals the impact of atmospheric turbulence-induced scintillations on the outage capacity
availability for the THz link. Indeed, the detrimental impact of both rain and strong water
vapor concentration along the propagation path is once again observed, but all of the
illustrated results of Figure 4 slightly but not negligibly outperform their corresponding
results from Figure 3 in terms of outage capacity for the THz link. It is therefore validated
that, contrary to FSO links, the outage performance for THz links (and more specifically the
outage performance in terms of outage capacity metric for THz links) is not significantly ‐
degraded to as great an extent as atmospheric turbulence along the propagation path ‐
is getting stronger. The latter does not alter the fact that the influence of atmospheric
turbulence should be also considered for the design of THz links. It is important to recall‐
here that this behavior, in terms of outage capacity metric through turbulent channels, is
qualitatively consistent with the experimental findings in [24], concerning both FSO and
THz signal transmissions in terms of other critical performance metrics as well as with the
very recently published findings in [70], where the outage performance of such THz links
is assessed in terms of the average bit error rate (ABER) metric.

Figure 4. Outage capacity (Cout /B) versus transmitted power, weak turbulence, moderate to strong
water vapor concentration, weak NZB PE and R = 0.1.

6. Discussion
In this work, the capacity performance and availability for a typical LOS THz wireless‐
communication link has been evaluated in terms of average capacity and outage capacity

critical metrics, including, for the first time, the joint influence of atmospheric turbulence,

generalized stochastic pointing errors where the boresight has been taken into account,
attenuation due the unavoidable emergence of FSPL, and molecular attenuation due to
water vapor along with attenuation due to adverse weather effects such as rain or fog.
In this context, a key novelty of this work is that, for the first time, novel closed form
average capacity and outage capacity mathematical expressions have been derived for both
gamma and gamma–gamma modeled turbulent channels along with incorporating all of
‐ ‐


Electronics 2023, 12, 1336 15 of 18

the above-mentioned major effects. Their analytical results (which are further verified by
simulations) demonstrate the individual influence of each of these crucial effects, let alone
their joint critical impact on THz performance and availability. Additionally, the obtained
results and findings have been proven to be reasonable and qualitatively consistent with
previous experimental works that evaluate other significant performance and availability
metrics. In short, our findings demonstrate that, although both average capacity and outage
capacity metrics obtain acceptable values for the establishment of LOS THz links under
realistic link parameter values, they are significantly degraded by the joint impact of all
the above impairments, especially for lower transmission power levels. Specifically, they
are degraded to a greater extent by molecular attenuation due to water vapor along with
rain and strong stochastic pointing errors especially when including boresight component,
whereas to a smaller extent they are also degraded by turbulence-induced scintillations
along with fog or clear weather conditions. Consequently, our findings can be used for
the design of the emerging high speed LOS THz links, taking into consideration the prime
factors that affect their capacity performance.

Author Contributions: Conceptualization, G.K.V., K.A., A.K., H.E.N.; Methodology, G.K.V., K.A,
A.K., H.E.N.; Software, G.K.V., K.A.; Validation, G.K.V., H.E.N., K.A.; Investigation, G.K.V., H.E.N.,
K.A.; Resources, G.K.V., K.A., H.E.N.; Writing—original draft preparation, G.K.V.; Writing—review
and editing, G.K.V., H.E.N.; Supervision, G.K.V., K.A, H.E.N.; Funding acquisition, G.K.V., K.A,
H.E.N. All authors have read and agreed to the published version of the manuscript.
Funding: This research was funded by Ajman University, grant number 2022-IRG-ENIT-19.
Acknowledgments: The authors acknowledge funding from Ajman University under grant agree-
ment 2022-IRG-ENIT-19.
Conflicts of Interest: The authors declare no conflict of interest.

References
1. Kokkoniemi, J.; Boulogeorgos, A.A.A.; Aminu, M.; Lehtomäki, J.; Alexiou, A.; Juntti, M. Impact of beam misalignment on THz
wireless systems. Nano Commun. Netw. 2020, 24, 100302. [CrossRef]
2. Boulogeorgos, A.A.A.; Papasotiriou, E.N.; Alexiou, A. Analytical performance assessment of THz wireless systems. IEEE Access
2019, 7, 11436–11453. [CrossRef]
3. Kürner, T.; Priebe, S. Towards THz communications-status in research, standardization and regulation. J. Infrared Millim. Terahertz
Waves 2014, 35, 53–62. [CrossRef]
4. Akyildiz, I.F.; Jornet, J.M.; Han, C. Terahertz band: Next frontier for wireless communications. Phys. Commun. 2014, 12, 16–32.
[CrossRef]
5. Seeds, A.J.; Shams, H.; Fice, M.J.; Renaud, C.C. Terahertz photonics for wireless communications. J. Light. Technol. 2015, 33,
579–587. [CrossRef]
6. Federici, J.F.; Ma, J.; Moeller, L. Review of weather impact on outdoor terahertz wireless communication links. Nano Commun.
Netw. 2016, 10, 13–26. [CrossRef]
7. Shawon, M.E.; Chowdhury, M.Z.; Hossen, M.B.; Ahmed, M.F.; Jang, Y.M. Rain Attenuation Characterization for 6G Terahertz
Wireless Communication. In Proceedings of the 2021 International Conference on Artificial Intelligence in Information and
Communication (ICAIIC), Jeju Island, Republic of Korea, 20–23 April 2021; pp. 3535–3538.
8. Chowdhury, M.Z.; Shahjalal, M.; Ahmed, S.; Jang, Y.M. 6G wireless communication systems: Applications, requirements,
technologies, challenges, and research directions. IEEE Open J. Commun. Soc. 2020, 1, 957–975. [CrossRef]
9. Federici, J.; Moeller, L. Review of terahertz and subterahertz wireless communications. J. Appl. Phys. 2010, 107, 6. [CrossRef]
10. Ghassemlooy, Z.; Arnon, S.; Uysal, M.; Xu, Z.; Cheng, J. Emerging optical wireless communications-advances and challenges.
IEEE J. Sel. Areas Commun. 2015, 33, 1738–1749. [CrossRef]
11. Su, K.; Moeller, L.; Barat, R.B.; Federici, J.F. Experimental comparison of terahertz and infrared data signal attenuation in dust
clouds. JOSA A 2012, 29, 2360–2366. [CrossRef] [PubMed]
12. Su, K.; Moeller, L.; Barat, R.B.; Federici, J.F. Experimental comparison of performance degradation from terahertz and infrared
wireless links in fog. JOSA A 2012, 29, 179–184. [CrossRef]
13. Yang, Y.; Mandehgar, M.; Grischkowsky, D.R. Broadband THz signals propagate through dense fog. IEEE Photonics Technol. Lett.
2014, 27, 383–386. [CrossRef]
14. Ma, J.; Moeller, L.; Federici, J.F. Experimental comparison of terahertz and infrared signaling in controlled atmospheric turbulence.
J. Infrared Millim. Terahertz Waves 2015, 36, 130–143. [CrossRef]
Electronics 2023, 12, 1336 16 of 18

15. Koch, M. Terahertz communications: A 2020 vision. In Terahertz Frequency Detection and Identification of Materials and Objects; Miles,
R.E., Zhang, X.C., Eisele, H., Krotkus, A., Eds.; Springer: Dordrecht, The Netherlands, 2007; pp. 325–338.
16. Ma, J.; Vorrius, F.; Lamb, L.; Moeller, L.; Federici, J.F. Experimental comparison of terahertz and infrared signaling in laboratory-
controlled rain. J. Infrared Millim. Terahertz Waves 2015, 36, 856–865. [CrossRef]
17. Ma, J.; Vorrius, F.; Lamb, L.; Moeller, L.; Federici, J.F. Comparison of experimental and theoretical determined terahertz attenuation
in controlled rain. J. Infrared Millim. Terahertz Waves 2015, 36, 1195–1202. [CrossRef]
18. Ishii, S.; Sayama, S.; Kamei, T. Measurement of rain attenuation in terahertz wave range. Wirel. Eng. Technol. 2011, 2, 119–124.
[CrossRef]
19. Chaccour, C.; Soorki, M.N.; Saad, W.; Bennis, M.; Popovski, P.; Debbah, M. Seven defining features of terahertz (THz) wireless
systems: A fellowship of communication and sensing. IEEE Commun. Surv. Tutor. 2022, 24, 967–993. [CrossRef]
20. Taherkhani, M.; Kashani, Z.G.; Sadeghzadeh, R.A. On the performance of THz wireless LOS links through random turbulence
channels. Nano Commun. Netw. 2020, 23, 100282. [CrossRef]
21. Taherkhani, M.; Sadeghzadeh, R.A.; Kashani, Z.G. Attenuation analysis of THz/IR waves under different turbulence conditions
using gamma-gamma model. In Proceedings of the Electrical Engineering (ICEE), Iranian Conference on, Sadjad University of
Technology, Mashhad, Iran, 8–10 May 2018; pp. 424–428.
22. Varotsos, G.K.; Nistazakis, H.E.; Volos, C.K.; Tombras, G.S. FSO links with diversity pointing errors and temporal broadening of
the pulses over weak to strong atmospheric turbulence channels. Optik 2016, 127, 3402–3409. [CrossRef]
23. Varotsos, G.K.; Nistazakis, H.E.; Petkovic, M.I.; Djordjevic, G.T.; Tombras, G.S. SIMO optical wireless links with nonzero boresight
pointing errors over M modeled turbulence channels. Opt. Commun. 2017, 403, 391–400. [CrossRef]
24. Varotsos, G.K.; Nistazakis, H.E.; Tombras, G.S. OFDM RoFSO Links with Relays Over Turbulence Channels and Nonzero
Boresight Pointing Errors. J. Commun. 2017, 12, 644. [CrossRef]
25. Varotsos, G.K.; Nistazakis, H.E.; Stassinakis, A.N.; Volos, C.K.; Christofilakis, V.; Tombras, G.S. Mixed Topology of DF Relayed
Terrestrial Optical Wireless Links with Generalized Pointing Errors over Turbulence Channels. Technologies 2018, 6, 121. [CrossRef]
26. Chatzidiamantis, N.D.; Sandalidis, H.G.; Karagiannidis, G.K.; Matthaiou, M. Inverse Gaussian modeling of turbulence-induced
fading in free-space optical systems. J. Light. Technol. 2011, 29, 1590–1596. [CrossRef]
27. Ghassemlooy, Z.; Popoola, W.O. Terrestrial free-space optical communications. In Mobile and Wireless Communications: Network
Layer and Circuit Level Design; InTech: London, UK, 2010; pp. 355–392.
28. Farid, A.A.; Hranilovic, S. Outage capacity optimization for free space optical links with pointing errors. IEEE/OSA J. Light.
Technol. 2007, 25, 1702–1710. [CrossRef]
29. Tsiftsis, T.A.; Sandalidis, H.G.; Karagiannidis, G.K.; Uysal, M. Optical wireless links with spatial diversity over strong atmospheric
turbulence channels. IEEE Trans. Wirel. Commun. 2009, 8, 951–957. [CrossRef]
30. Gappmair, W.; Hranilovic, S.; Leitgeb, E. Performance of PPM on terrestrial FSO links with turbulence and pointing errors. IEEE
Commun. Lett. 2010, 14, 468–470. [CrossRef]
31. Yang, F.; Cheng, J.; Tsiftsis, T.A. Free-space optical communication with nonzero boresight pointing errors. IEEE Trans. Commun.
2014, 62, 713–725. [CrossRef]
32. Nistazakis, H.E.; Tsiftsis, T.A.; Tombras, G.S. Performance analysis of free-space optical communication systems over atmospheric
turbulence channels. IET Commun. 2009, 3, 1402–1409. [CrossRef]
33. Nistazakis, H.E.; Tsigopoulos, A.D.; Hanias, M.P.; Psychogios, C.; Marinos, D.; Aidinis, C.; Tombras, G.S. Estimation of Outage
Capacity for Free Space Optical Links over IK and K Turbulent Channels. Radioengineering 2011, 20, 493–498.
34. Stassinakis, A.N.; Nistazakis, H.E.; Varotsos, G.K.; Tombras, G.S.; Tsigopoulos, A.D.; Christofilakis, V. Outage capacity estimation
of FSO links with pointing errors over gamma turbulence channels. In Proceedings of the 2016 5th International Conference on
Modern Circuits and Systems Technologies (MOCAST), Thessaloniki, Greece, 12–14 May 2016; pp. 1–4.
35. Djordjevic, G.T.; Petkovic, M.I.; Spasic, M.; Antic, D.S. Outage capacity of FSO link with pointing errors and link blockage. Opt.
Express 2016, 24, 219–230. [CrossRef]
36. Petkovic, M.I.; Djordjevic, G.T. Effects of pointing errors on average capacity of FSO links over gamma-gamma turbulence channel.
In Proceedings of the 2013 11th International Conference on Telecommunications in Modern Satellite, Cable and Broadcasting
Services (TELSIKS), Nis, Serbia, 16–19 October 2013; Volume 2, pp. 481–484.
37. Ansari, I.S.; Yilmaz, F.; Alouini, M.S. Performance analysis of FSO links over unified Gamma-Gamma turbulence channels. In
Proceedings of the 2015 IEEE 81st Vehicular Technology Conference (VTC Spring), Glasgow, Scotland, 11–14 May 2015; pp. 1–5.
38. Garrido-Balsells, J.M.; Jurado-Navas, A.; Paris, J.F.; Castillo-Vazquez, M.; Puerta-Notario, A. On the capacity of M-distributed
atmospheric optical channels. Opt. Lett. 2013, 38, 3984–3987. [CrossRef] [PubMed]
39. Nistazakis, H.E.; Tombras, G.S.; Tsigopoulos, A.D.; Karagianni, E.A.; Fafalios, M.E. Capacity estimation of optical wireless
communication systems over moderate to strong turbulence channels. J. Commun. Netw. 2009, 11, 384–389. [CrossRef]
40. Peppas, K.P.; Stassinakis, A.N.; Nistazakis, H.E.; Tombras, G.S. Capacity analysis of dual amplify-and-forward relayed free-space
optical communication systems over turbulence channels with pointing errors. J. Opt. Commun. Netw. 2013, 5, 1032–1042.
[CrossRef]
41. Moon, E.B.; Jeon, T.I.; Grischkowsky, D.R. Long-path THz-TDS atmospheric measurements between buildings. IEEE Trans.
Terahertz Sci. Technol. 2015, 5, 742–750. [CrossRef]
Electronics 2023, 12, 1336 17 of 18

42. Kim, G.R.; Jeon, T.I.; Grischkowsky, D. 910-m propagation of THz ps pulses through the atmosphere. Opt. Express 2017, 25,
25422–25434. [CrossRef]
43. Castro, C.; Elschner, R.; Merkle, T.; Schubert, C.; Freund, R. Long-range high-speed THz-wireless transmission in the 300 GHz
band. In Proceedings of the 2020 Third International Workshop on Mobile Terahertz Systems (IWMTS), Essen, Germany, 1–2 July
2020; pp. 1–4.
44. Kokkoniemi, J.; Lehtomäki, J.; Juntti, M. Simplified molecular absorption loss model for 275–400 gigahertz frequency band.
In Proceedings of the 12th European Conference on Antennas and Propagation (EuCAP 2018), London, UK, 9–13 April 2018;
pp. 1–5.
45. Boulogeorgos, A.A.A.; Papasotiriou, E.N.; Kokkoniemi, J.; Lehtomäki, J.; Alexiou, A.; Juntti, M. Performance evaluation of THz
wireless systems operating in 275–400 GHz band. In Proceedings of the 2018 IEEE 87th Vehicular Technology Conference (VTC
Spring), Porto, Portugal, 3–6 June 2018; pp. 1–5.
46. Fiorino, S.T.; Bartell, R.J.; Krizo, M.J.; Marek, S.L.; Bohn, M.J.; Randall, R.M.; Cusumano, S.J. A computational tool for evaluating
THz imaging performance in brownout or whiteout conditions at land sites throughout the world. In Atmospheric Propagation VI;
SPIE: Bellingham, WA, USA, 2009; Volume 7324, pp. 308–319.
47. Kokkonen, M.; Juttula, H.; Mäkynen, A.; Myllymäki, S.; Jantunen, H. The effect of drop shape, sensing volume and raindrop size
statistics to the scattered field on 300 GHz. IEEE Access 2021, 9, 101381–101389. [CrossRef]
48. Bao, L.; Zhao, H.; Zheng, G.; Ren, X. Scintillation of THz transmission by atmospheric turbulence near the ground. In Proceedings
of the 2012 IEEE Fifth International Conference on Advanced Computational Intelligence (ICACI), Nanjing, China, 18–20 October
2012; pp. 932–936.
49. Ma, J.; Moeller, L.; Federici, J. Terahertz performance in atmospheric turbulence. In Proceedings of the 2014 39th International
Conference on Infrared, Millimeter, and Terahertz waves (IRMMW-THz), The University of Arizona, Tucson, AZ, USA, 14–19
September 2014; pp. 1–4.
50. Varotsos, G.K.; Chatzikontis, E.V.; Kapotis, E.; Nistazakis, H.E.; Aidinis, K.; Christofilakis, V. THz Links Performance Study for
Gamma Turbulence Links with Path Loss and Pointing Errors. In Proceedings of the 2022 11th International Conference on
Modern Circuits and Systems Technologies (MOCAST), Bremen, Germany, 8–10 June 2022; pp. 1–4.
51. Priebe, S.; Jacob, M.; Kürner, T. The impact of antenna directivities on THz indoor channel characteristics. In Proceedings of the
2012 6th European Conference on Antennas and Propagation (EUCAP), Prague, Czech Republic, 26–30 March 2012; pp. 478–482.
52. Ekti, A.R.; Boyaci, A.; Alparslan, A.; Ünal, İ.; Yarkan, S.; Görçin, A.; Arslan, H.; Uysal, M. Statistical modeling of propagation
channels for terahertz band. In Proceedings of the 2017 IEEE Conference on Standards for Communications and Networking
(CSCN), Helsinki, Finland, 18–20 September 2017; pp. 275–280.
53. Han, C.; Akyildiz, I.F. Three-dimensional end-to-end modeling and analysis for graphene-enabled terahertz band communications.
IEEE Trans. Veh. Technol. 2016, 66, 5626–5634. [CrossRef]
54. Papasotiriou, E.N.; Boulogeorgos, A.A.A.; Alexiou, A. Performance analysis of THz wireless systems in the presence of antenna
misalignment and phase noise. IEEE Commun. Lett. 2020, 24, 1211–1215. [CrossRef]
55. Siles, G.A.; Riera, J.M.; Garcia-del-Pino, P. Atmospheric attenuation in wireless communication systems at millimeter and THz
frequencies [wireless corner]. IEEE Antennas Propag. Mag. 2015, 57, 48–61. [CrossRef]
56. Ippolito, L.J., Jr. Attenuation by Atmospheric Gases. In Radiowave Propagation in Satellite Communications; Springer Science &
Business Media: New York, NY, USA, 1986; pp. 25–37.
57. Al-Habash, A.; Andrews, L.C.; Phillips, R.L. Mathematical model for the irradiance probability density function of a laser beam
propagating through turbulent media. Opt. Eng. 2001, 40, 1554–1562. [CrossRef]
58. Gradshteyn, I.S.; Ryzhik, I.M. Table of Integrals, Series, and Products, 6th ed.; Academic Press, Inc.: New York, NY, USA, 2000.
59. Varotsos, G.K.; Stassinakis, A.N.; Nistazakis, H.E.; Tsigopoulos, A.D.; Peppas, K.P.; Aidinis, C.J.; Tombras, G.S. Probability of fade
estimation for FSO links with time dispersion and turbulence modeled with the gamma–gamma or the IK distribution. Optik
2014, 125, 7191–7197. [CrossRef]
60. Varotsos, G.K.; Nistazakis, H.E.; Stassinakis, A.N.; Tombras, G.S.; Christofilakis, V.; Volos, C.K. Outage performance of mixed,
parallel and serial DF relayed FSO links over weak turbulence channels with nonzero boresight pointing errors. In Proceedings of
the 2018 7th International Conference on Modern Circuits and Systems Technologies (MOCAST), Thessaloniki, Greece, 7–9 May
2018; pp. 1–4.
61. Nistazakis, H.E.; Tombras, G.S. On the use of wavelength and time diversity in optical wireless communication systems over
gamma–gamma turbulence channels. Opt. Laser Technol. 2012, 44, 2088–2094. [CrossRef]
62. Boluda-Ruiz, R.; García-Zambrana, A.; Castillo-Vázquez, C.; Castillo-Vázquez, B. Novel approximation of misalignment fading
modeled by Beckmann distribution on free-space optical links. Opt. Express 2016, 24, 22635–22649. [CrossRef]
63. Abramovitz, M.; Stegun, I.A. Handbook of Mathematical Functions with Formulas, Graphs, and Mathematical Tables, 9th ed.; Dover:
New York, NY, USA, 1972.
64. Adamchik, V.S.; Marichev, O.I. The algorithm for calculating integrals of hypergeometric type functions and its realization in
REDUCE system. In Proceedings of the International Symposium on Symbolic and Algebraic Computation, Tokyo, Japan, 20–24
August 1990; pp. 212–224.
65. Ansari, I.S.; Yilmaz, F.; Alouini, M.S. Performance analysis of free-space optical links over Málaga (M) turbulence channels with
pointing errors. IEEE Trans. Wirel. Commun. 2016, 15, 91–102. [CrossRef]
Electronics 2023, 12, 1336 18 of 18

66. Varotsos, G.K.; Aidinis, K.; Nistazakis, H.E. On the Outage Capacity of Transdermal Optical Wireless Links with Stochastic Spatial
Jitter and Skin-Induced Attenuation. Photonics 2021, 8, 553. [CrossRef]
67. Furrer, S.; Coronel, P.; Dahlhaus, D. Simple ergodic and outage capacity expressions for correlated diversity Ricean fading
channels. IEEE Trans. Wirel. Commun. 2006, 5, 1606–1609. [CrossRef]
68. Wang, H.; Dong, X.; Yi, M.; Xue, F.; Liu, Y.; Liu, G. Terahertz high-gain offset reflector antennas using SiC and CFRP material.
IEEE Trans. Antennas Propag. 2017, 65, 4443–4451. [CrossRef]
69. Anastasov, J.A.; Zdravkovic, N.M.; Djordjevic, G.T. Outage capacity evaluation of extended generalized-K fading channel in the
presence of random blockage. J. Frankl. Inst. 2015, 352, 4610–4623. [CrossRef]
70. Varotsos, G.K.; Aidinis, K.; Nistazakis, H.E. Average BER Performance Estimation of Relayed THz Links with Losses, Molecular
Attenuation, Adverse Weather Conditions, Turbulence and Generalized Pointing Errors. Photonics 2022, 9, 671. [CrossRef]

Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual
author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to
people or property resulting from any ideas, methods, instructions or products referred to in the content.

You might also like