1. Introduction
Visible light communication (VLC) is an emerging technology that poses an alternative to traditional physical communication technologies such as radio frequency (RF) communication. Due to its high frequency, it presents several advantages over RF communication, such as resistance to electromagnetic interference, an unlicensed spectrum, lower energy consumption, and reduced costs. The benefits come from the usage of energy-efficient light-emitting diodes (LEDs), which are now being deployed as the primary source of lighting.
A relevant use case for VLC is underground mining (UM) communications [
1]. Visible light communication in underground mining (UM-VLC) is a promising solution due to its low cost and the latent necessity of lighting inside mines [
2]. Although VLC solves many issues, such as interference and fading phenomena, highly directional propagation harms the system’s performance. The unpredictable receiver orientation and the user’s mobility frequently lead to a link misalignment. The misalignment of the link disrupts communication, causing link failures. To address these outages, different studies have studied the application of diversity techniques such as receiver diversity [
3,
4].
The existing proposed VLC implementations in mining environments require both mobility and a variable receiver orientation, which directly impact the availability of line-of-sight (LOS) links. Consequently, characterizing the availability of these links is a crucial element in understanding the system’s behavior. Although the impact of the receiver orientation has been studied extensively in the literature [
5,
6], research on random receiver orientations is limited by assumptions that do not accurately represent the underground mining scenario studied.
In the context of relevant and related work, the probability of a power outage has been studied under the case of a random transmitter orientation [
7]. However, the receiver orientation has been considered constant and orthogonal to the ground plane. In [
8,
9], the power outage probability and the bit error rate for on–off keying (OOK) were studied under a uniform and Gaussian distribution of the receiving angle. However, the transmitter was considered orthogonal to the ground, and the distributions were not a function of the receiver’s position. In [
10], the bit error rate and the signal-to-noise ratio distribution were considered for a case with arbitrary user positioning and an arbitrary orientation, and the optimum tilt angle was obtained. However, the article derived the effect of a random receiver orientation from fitting a known distribution. Likewise, [
11] derived the bit error rate for DCO-OFDM with an arbitrary orientation by modeling the orientation and the channel using a Laplace distribution. However, the user positioning was maintained as fixed in the expression. In [
12], the probability of a blockage and the probability of coverage were obtained by studying the shadowing generated by a cylinder. However, the transmitter was orthogonal to the ground, and the receiver was unaffected by a random orientation.
A random receiver orientation has also been studied in the context of non-orthogonal multiple-access (NOMA) schemes for VLC [
13]. However, the reception angle was modeled using a normal distribution, which was also supposed to be independent of the radius. The bit error rate has been studied for single-input single-output (SISO) underground mining channels [
14]. However, the receiving angle was considered to be independent of the user’s position. Finally, the power outage probability of a hybrid RF-VLC link was studied in a random receiver orientation [
15]. However, the receiving angle was assumed to follow a normal distribution, similarly to in the NOMA study.
In this work, we investigate the effects of the receiver orientation from the perspective of an underground mining use case in SISO, multiple-input single-output (MISO), and single-input multiple-output (SIMO) scenarios. In particular, we investigate the case where the random receiver orientation’s density function is determined by its position with respect to the transmitter in a tunnel. The contributions of this paper can be summarized as follows:
We develop an analytical framework to obtain the LoS probability of a mobile receiver with a random orientation given a rectangular cell for SISO, MISO, and SIMO cases.
We obtain the closed analytical form of the LOS probability, given a uniform distribution of the receiver in the rectangular area and a uniform orientation of the user.
We verify the analytical results through extensive simulation in all cases of interest.
The structure of this article is as follows. The MIMO-VLC model for use in underground environments and the impact of LoS communication links are presented in
Section 2. The LoS probability and the angle of incidence are derived in
Section 3. The analytical expressions of the SISO LoS probability, the SIMO LoS probability, and the MISO LoS probability are derived in
Section 4,
Section 5, and
Section 6, respectively. The numerical and analytical results are presented in
Section 7. Finally, conclusions are given in
Section 8.
2. MIMO Channel Model
The underground mining MIMO channel was modeled to account for multiple phenomena, such as reflections caused by irregular walls, scattering effects caused by dust particles, random shadowing caused by machinery entering the mine, and the random tilting of both the receiver and transmitter [
16,
17]. The tilt of the receiver and transmitter is defined by their corresponding azimuth angle
and elevation angle
. The geometry of the system is depicted in
Figure 1. The orientation of the receiver and the transmitter in space is represented by their vectors:
where
and
correspond to the azimuthal and elevation angles of the receiver, and
and
correspond to the azimuthal and elevation angles of the transmitter.
The channel model is characterized by a matrix of
, where
M represents multiple transmitters and
N represents multiple receivers. Each transmitter–receiver pair employs ray tracing to generate a channel that linearly sums the LoS component, the reflective components (NLoS), and the scattering components. The channel matrix is defined as follows:
where
is the LOS component,
is the scattering component, and
is the reflective component.
is defined by the equation [
2]
where
represents the photodetector active area,
m denotes the Lambertian order,
is the Euclidean distance between the receiver and the transmitter,
is the field of view, and
is the gain produced by the concentrator, defined by
where
is the gain and
is the angle between the receiver and the path between the receiver and the transmitter.
is the dot product between
and
, and
is the Heaviside function. Finally,
is the shadowing probability between the photodetector and the receiver. The shadowing probability is defined using a Poisson point process, where the expected probability of not encountering an obstacle during a period, t, is
with
being the intensity parameter,
t the expected time, and
the expected probability of an obstacle causing shadowing [
18].
where
X and
Y are the horizontal and vertical dimensions of the tunnel,
w is the width of the object,
h is the height of the object,
is the probability associated with the center of the object, and
is the probability associated with the dimensions of the object. However, due to the cost of computing this equation multiple times over the simulation duration, a closed analytical form is obtained by approximating the probability density of the object dimensions using
A closed analytical form of the probability can be found in
Appendix A.
The non-line-of-sight (NLoS) component is the result of multiple reflections that can occur between the transmitter and the receiver. Hence, considering the NLoS component results in the recursive infinite summation of all the path combinations produced by the discrete reflectors. However, to simplify the calculation, assuming that the distance between the walls is considerable, we can limit the sum to one iteration. This relationship can be defined using the following equation:
where
corresponds to the area of the reflector and
is the reflection coefficient of the wall. Similarly to the transmitter and receiver case, the reflector is characterized by a vector,
, which is associated with a random azimuth angle,
, and an elevation,
, and the distance between the reflector and the transmitter or receiver is defined by
and
, respectively. Finally, the scattering component is defined by
where
is the total distance covered by the rays from the transmitter to the receiver,
is the angle between the scattering and the transmitter,
is the angle between the scattering and the receiver, and
is the expected value of the scattering given the parameters
where
,
g, and
f are atmospheric constants,
is the reflection coefficient of the particles,
is
,
N is the number of particles, and
is the angle of the scattering. The equations for
and
are [
19]
where
, with
being the radius of the particles,
m the refractive coefficient,
the depolarization factor of air,
the molecular number density of air, and
the refractive index of air.
and
are the Bessel function and the first kind of Hankel function, respectively.
An important metric for assessing the effect of the receiving angle is the power outage probability. However, as will be shown in the following sections, the power outage probability cannot be obtained analytically, given the interdependence of the receiving angle with the mobility of the user and the orientation of the receiver. However, it is possible to obtain the LOS probability, which correlates strongly with the system’s average power. Given a variable angle of inclination for the receiver from
to
,
Figure 2 shows the average power compared to the LoS probability.
The average power and the LoS probability are obtained by simulating the channel at each point inside the tunnel by varying
between 0 and
. The systems and scenario parameters used for the simulation are presented in
Table 1 and
Table 2. The average power is obtained by computing the mean power of all the resulting DC channel values, and the LoS probability is obtained by taking the number of available LoS links and dividing them by the total number of links. The presented average power is further divided into a case with an LoS component and a case without an LoS component. The average for the case without an LoS component is computed by taking the mean of the DC power from all the corresponding phenomena within the channel bar the LoS component, even if an LoS component is available. Since we consider all the computed channel responses, the average power is equally conditioned by the distance between the receiver and the transmitter for all the varying inclination values.
The figure shows that the LoS probability and the average power in the tunnel are directly correlated when the LoS component is taken into account. Indeed, the correlation between the Los probability and the average power is . Based on this, we can conclude that obtaining a metric for the LoS probability given the present scenario would help predict the performance according to other metrics such as the power and optimize for the optimal angle.
3. Line-of-Sight Probability
The LoS probability is the probability that the receiver and the transmitter send data to each other using the LoS channel. To satisfy this condition, the gain from the concentrator must be greater than 0. The condition is satisfied as long as the reception angle is less than the total field of view. So, the probability is
where
can be defined by the inverse cosine:
and
is the Euclidean norm, and
is the vector from the transmitter to the receiver, which is
where
r is the radius of the sphere around the receiver that parameterizes the position of the receiver,
L is the distance from the transmitter to the receiver in the X–Y plane, and
h is the height difference between the transmitter and the receiver.
Unlike in the orthogonal case (
), the tilt produced by
and
involves a non-circular area of possible transmitter positions, which depends on the angle
. Since the azimuth angle changes with the rotations of the receiver, the probability of reception given a certain coordinate of the center of the receiver is a function of the radius
L. The density probability function of the receiving angle is as follows:
As using the probability density function has an important difference to the orthogonal case, the analysis of the system changes. Since
L is not deterministic, the value of the function changes with respect to the probability given by the mobility of the user, and the power will depend on both
and
L. To illustrate the difference between both systems,
Figure 3 shows the set of transmitters with their LoS probability for the orthogonal case (
) and the non-orthogonal case (
,
). The receiver is positioned at
, which is represented by the red dot in both figures. The geometry of both cases shows the reason behind the complexity of obtaining the LoS probability. In the orthogonal case, the probability can be calculated by taking the indicator function over the area of potential transmitters. However, in the studied case, each coordinate will have an LoS probability that can be understood as the integration over the realization of rotating the receiver around itself. In that case, points closer to the origin will still have a probability of one, as they will always have the link in each rotation, but points far from the receiver will be visited with less frequency.
4. SISO LoS Probability
The LoS probability can be obtained using Equation (
14) by combining
and applying cosine to both sides. Since
belongs to the interval
, the cosine is monotonically decreasing. The consistent decline enables us to express Equation (
14) by utilizing
where
due to the parity of the cosine function.
The left argument of the probability is obtained using the definitions of
and
. The resulting expressions for the norm and the dot product are
where the approximation of the norm holds as long as
r is significantly smaller than both
h and
L. Using both expressions, we can rewrite the probability as
The above equation has three random variables, which dictate the position of the receiver, determined by
and
L, and the orientation of the receiver, which is determined by
. Since the receiver’s orientation is assumed to be independent of its position, we can obtain the expected LOS probability using the following equation:
where
is the probability density function of the receiver’s orientation, assumed to be uniform within the interval
, and
is the probability density function of the receiver’s position, assumed to be uniform too. Due to the use of the radius and the angle, and given the rectangular boundaries of the system, the spatial density function has to be rewritten to
where
n is an index from 1 to 8 that represents the different triangular partitions of the corresponding rectangle. The resulting integral is the equation
where
and
.
The inverse cosine can be approximated by using an odd number of linear approximations that are symmetric to the center. Each linear approximation will be associated with a slope,
, and an offset,
. Linear approximations represent an interval of the function. These intervals can be found by solving the following expression:
The orientation of the inequality depends on the sign of
. If the sign is positive, then the thresholds
represent either the lower bound from
or the upper bounds from
, where
represents the minimum of the function if it exists. If the sign is negative, then the thresholds represent the upper bound. Because of its quadratic form, the equation can have zero to two feasible solutions over zero that follow the following solution:
The intervals are generated depending on the number of solutions. If has zero solutions and has a solution, then is valid in the interval . If has zero solutions, then is skipped. If has one solution, then is valid in the interval . If it has two positive solutions, then is valid in the interval and the interval .
Using the solutions, we can form a succession of solutions such that the nth element of the solution is associated with some
constant that corresponds to the approximation from
to
, the values of which depend on the solutions for the previous equation. Using this sequence, we can rewrite the integral as the sum of the intervals that are associated with each
over the sum of the different triangles that compose the probability distribution. The resulting equation is
The integral can be divided into the
and
cases, assuming that
is equivalent to having
and
. Furthermore, the
case corresponds to the density functions of the tangent denominator, while the
case corresponds to the density functions of the cotangent denominator. For the
case, the integral is
with
. The function
obeys
For the purpose of notation, it will be called
.
The integral solution consists of the following equations:
where
integrates an interval completely inside the rectangular boundary,
integrates an interval where the upper bound is completely outside of the rectangular boundary for every angle, and
integrates an interval where
defines whether the upper boundary is inside or outside of the boundary. The subterm
is
The subterm
, with
, is given in
Appendix B.
The integral of the even case follows the following equations:
where the subterm
is
and the subterm
is given in
Appendix B.
5. SIMO LoS Probability
The single-input multiple-output case includes
receivers with an angular offset,
, such as
. Let us call
the sum of the probabilistic
and the offset. The dot product between
and
will be
Obtaining the probability of
receivers communicating during the same period involves intersecting the intervals produced by each
. The intervals produced by each
are in the form of
where
is the argument of the inverse, or
Since
and the arguments of the intervals mentioned previously can exist outside of the interval, to solve this, we introduce the following transformation:
To obtain the points where the modulus cycles back to 0 or
, we need to find the solutions to the equation
with
being the solution to the equation. The equation has zero to two solutions. The possible cases are the following:
If the equation has zero solutions, then the argument of the modulus either belongs to or is always below the interval. If , then it is the former; if , then it is the latter.
If the equation has one solution, then it is either from the upper bound or the lower bound. If , then it is the former; if , then it is the latter.
If the equation has two solutions, then one of the bounds oscillates around . If , then it goes below from to ; if , then it goes above from to .
Without the loss of generality, the solution is always a pair,
and
, where
. If there are one or fewer solutions, then
goes to
. If there are zero solutions, then
goes to
too. Then, the intervals can be generalized using
where
.
Similarly to in the SISO case, we can form a succession of distances to the center such that the nth element of the sequence corresponds to an interval,
to
, such that both
and
are solutions for some pair of receivers. Let us assume that the first
M receivers have non-divergent interval solutions, where
corresponds to those which are not subject to a modulus and
to those which are. Finally, the last
N receivers have divergent interval solutions, which means that their lower bounds have a modulus applied to them. If
, then the resulting intervals are
Since
is 0,
. For the union to be between two non-intersecting intervals, the condition
must be satisfied, which is always true given the fact that the
associated with the minimum in
is always bigger than
associated with the maximum in
. If
, then
Using the succession of intervals
, we can compute the integral by using the sum
with
being the number of receivers that have disjoint intervals of
. The inequality
has zero to two solutions and can be trivially solved within each interval,
n. The solution to the integral of each
is equivalent to the integral of
multiplied by
.
6. MISO LOS Probability
For the MISO case, we assume that the
transmitter is at a distance,
, from the first transmitter, with an angle of
. The general formula for the dot product between
and
is
To simplify the expression, we use
. The case with
has a similar solution but was not evaluated. If
, then the norm and the dot product are
Since both the denominator and the numerator depend on
, the solutions for the intervals, inverse cosine approximations, and modulus depend on
. To address the dependency on
, we use small
steps to integrate and solve the approximations and modulus equations. The rate of change dependent on
allows for nearly constant solutions in a certain
interval. When we integrate small variances of
, the sectors being integrated have their own probability density, which is
where
corresponds to the lower bound of the interval and
to the upper bound of the interval.
corresponds to either
or
, while
k refers to either the odd or even case. Using these intervals, we can rewrite the integral as
The corresponding intervals generated for
are
The interval shows that the modulus equations depend on
. In the case of the threshold equations, they can be solved in a straightforward manner using the solutions given by the solutions for the
case. Using these solutions, we can derive zero to four solutions for the threshold using
Since the solutions are centered around
, we can use a similar method to generate the sequence for the case with
, using the reverse methodology for the case where
. The new modulus breakpoints are
Unlike in the SIMO case, the arrangement of the intervals does not follow a straightforward set of equations, as they are dependent on the corresponding phase which varies with both the distance L and the angle , which makes it possible, for example, that corresponds to a solution in the upper bound and corresponds to a solution in the lower bound.
Since
is injective from
to
, the function does not behave as desired between
and
. A solution to this is to patch the function using the following equation:
Since we are assuming that all transmitters are on the same axis, we can assume that
and
. Using this condition, we can divide the interval formation for a single transmitter in four cases depending on
. It is also relevant to note that the breaking points, given the new equation, have up to four solutions. We use
for the cases where
L is below the patching and
for the cases where
L is above the patching. The solutions for the cases above and below the patching are in
Appendix D.
Similarly to in the SIMO case, the resulting intervals have to be intersected to obtain the integrating intervals. Unlike in the SIMO case, the solution for the intersection does not have a closed, exact analytical solution. The equation can be solved using piecewise approximations or numerical methods. In this paper, we use the Newton–Raphson method to obtain both the minimum and maximum for each interval. A pair has a maximum of two real solutions, as the inverse of the tangent monotonically increases and the inverse cosine is either monotonically increasing on the positive side or has a single minimum. Thus, we can use both sides of the interval as starting points to obtain either zero, one, or two solutions. Given an interval from
to
of a set,
p, such that every approximation of the inverse cosine is constant and no equation,
, crosses
or
, the resulting intervals are
with
M being the intervals that are continuous between
and
and
N the intervals that are not continuous. Unlike in the SIMO case,
M can be zero under certain conditions. In this case, we define
and
as
and
, respectively.
Using the succession of intervals
corresponding to the intersections generated by the modulus equation, we obtain the expected probability as
where
M corresponds to the number of divisions. The solution of the integral for each bound is
where
corresponds to the terms associated with the coefficient
and
to those associated with the constant
, and
is the integral of the inverse tangent. The integral of
can be approximated using first-order approximations of the square root, which results in
where the resulting expression is in
Appendix C.
The expression for the integral of
is
On the other hand, the inverse tangent integral does not have a simple primitive when integrating by
. Exploiting the fact that
uses small steps, we linearize the inverse tangent around
. The resulting value is
8. Conclusions
In this paper, we obtained a generalized analytical expression for the LOS probability for SISO, SIMO, and MISO cases when the probability density function of the reception angle was a function of the receiver’s geometry and the user’s mobility. The expressions showed a close alignment between the simulation and analytical results, with slight errors produced by the approximations necessary to obtain an analytical integration form. The error of the analytical expression was , , and for the SISO, SIMO, and MISO cases, respectively. We studied the effect of the receiver’s angular elevation with a constant FOV, which showed that an optimal elevation existed for each FOV. Furthermore, the optimality of this angular elevation was not preserved from the SISO case to the SIMO or MISO cases. The effect of the height on the LoS probability was also obtained, showing that higher height differences increased the LoS probability. Finally, the resulting analytical expressions are a representative approach to the application of the underground mining proposal, showing predictive power to optimize the channel with respect to the geometrical elements inside it.
The SIMO and MISO analytical expressions can be expanded to an MIMO case, which can be performed in future work. Furthermore, future work will involve a comparison between different communication schemes and prove the improvements in an experimental setup.