Expand this Topic clickable element to expand a topic
Skip to content
Optica Publishing Group

RSS-based visible light positioning with unknown receiver tilting angle: robust design and experimental demonstration

Open Access Open Access

Abstract

In this paper, we consider the received signal strength (RSS)-based indoor visible light positioning (VLP) with unknown receiving angle. A novel VLP scheme is proposed by joint estimation of the receiver coordinate and receiver characteristic vector. We equivalently convert the original estimation problem into the problem that maximizes the projection of the RSS vector on the column space spanned by the measurement matrix. The proposed scheme does not require the prior knowledge of receiving coefficient and receiving angle. To further reduce the computational complexity, we also propose an importance sampling (IS) method for solving the VLP problem. Simulation and experimental results demonstrate the robustness of the proposed scheme to the receiving coefficient and receiving angle. Specifically, experimental results show that the proposed VLP scheme achieves a stable positioning accuracy below 7 cm under different receiver tilting angles in a 60 cm × 60 cm × 150 cm space, and is not sensitive to height perturbation.

© 2022 Optica Publishing Group under the terms of the Optica Open Access Publishing Agreement

1. Introduction

Recently, rapid development of mobile communication networks and electronic information technologies has promoted numerous location-based services [13]. Satellite positioning systems such as global positioning system (GPS) can provide users with meter-level localization services in outdoor environments [4]. Nevertheless, in indoor environments, the positioning accuracy of GPS cannot satisfy the demands of real-time localization and navigation [5]. To realize high-accuracy positioning in indoor environments, various positioning technologies have been developed using different wireless signals, such as WiFi [6], infrared [7], ultrasonic [8], Bluetooth [9], ultra-wideband [10] and radio-frequency identification (RFID) [11]. However, these technologies still suffer low positioning accuracy or high deployment cost.

Compared with the above positioning technologies, light emitting diode (LED)-based visible light positioning (VLP) enjoys appealing advantages such as high accuracy, low implementation cost and long life time [1215]. Existing VLP schemes estimate the terminal position from different optical measurements such as image [16], time of arrival (TOA) [17], time difference of arrival (TDOA) [18], angle of arrival (AOA) [19], angle difference of arrival (ADOA) [20] and received signal strength (RSS). Among these techniques, the RSS-based VLP is the most popular since RSS values can be easily acquired using a single photodiode (PD) with no need of auxiliary devices [21].

There have been plenty of studies on RSS-based VLP [2228]. In [2224], the receiver position was estimated using the received direct current (DC) signal strength, which is susceptible to ambient light and less practical in real environments. Reference [25] experimentally demonstrated a VLP system using the output alternating current (AC) signal strength of PD. As it requires specific values of all transceiver parameters, the VLP scheme in [25] is not robust to receiving coefficient. In order to achieve the robustness to receiving coefficient, Ref. [26] proposed a VLP scheme for the case where the receiver is perpendicular to the ceiling. However, the receiver may be tilted in practice, which limits the application scenarios of the VLP scheme in [26]. References [2730] considered the RSS-based VLP in the case of tilted receiver. However, they assumed perfect knowledge of the receiver orientation angle, which is a strong assumption in practical applications.

To address the above issues, this work proposes a robust RSS-based VLP scheme with unknown receiving angle. In the proposed scheme, the receiver coordinate and receiver characteristic vector are jointly estimated from the RSS vector. Through mathematical derivations, the original estimation problem can be equivalently converted into the problem that maximizes the projection of the RSS vector on the column space spanned by the measurement matrix. The proposed scheme is robust to the receiver characteristics, as it does not require the prior knowledge of receiving coefficient and receiving angle. To further reduce the computational complexity, we also propose an importance sampling (IS) method for solving the VLP problem instead of brute-force search (BFS). Simulation and experimental results demonstrate the robustness of the proposed scheme to receiving coefficient and receiving angle. Specifically, simulation results show that the computation time of the BFS method is about 26 times that of the IS method. Experimental results show that the proposed VLP scheme achieves a stable positioning accuracy below 7 cm under different PD tilting angles in a 60 cm $\times$ 60 cm $\times$ 150 cm space, and is not sensitive to height perturbation.

The remainder of this paper is organized as follows. The signal model and benchmark VLP schemes are presented in Section 2. Section 3. elaborates the proposed VLP scheme. Simulation and experimental results are given in Section 4. and Section 5. Finally, Section 6. concludes this paper.

2. System descriptions

In this section, we describe the considered VLP system and present the corresponding signal model.

2.1 System configuration

Figure 1(a) illustrates the considered indoor VLP system with $L$ LEDs as the transmitter and one tilted PD as the receiver. Assume that all the $L$ LEDs are at the same height $z_t$ and pointing downwards, which is the common case. Each LED acts as one anchor for RSS-based VLP, and the coordinate of LED $l$ is known as $\boldsymbol {c}_l = [x_l, y_l, z_t]^\top \in {\mathbb {R}}^3$, $l=1,\ldots, L$. The tilted PD receiver is located at the position ${\boldsymbol {c}}_r=[x_r, y_r, z_r]^\top \in {\mathbb {R}}^3$. The vertical tilting angle is denoted by $\phi$, and the horizontal tilting angle is denoted by $\beta$. As in Refs. [2224], we assume that the receiver height $z_r$ is known and the receiver moves in a horizontal plane. This assumption holds in various practical applications, e.g., where the receiver is installed on a robot or a cart. Without loss of generality, the receiver orientation vector is denoted by ${\boldsymbol {u}}_r = [\sin \beta \cos \phi, \cos \beta \cos \phi, \sin \phi ]^\top \in {\mathbb {R}}^3$ with $\| \boldsymbol {u}_r \|_2 = 1$. The distance between the LED $l$ and PD is calculated as $d_{r,l} = \|\boldsymbol {c}_l - \boldsymbol {c}_r\|_2 = \sqrt {(x_l - x_r)^2 + (y_l - y_r)^2 + z^2}$ with $z = z_t - z_r$.

 figure: Fig. 1.

Fig. 1. Illustration of the considered indoor VLP system. (a) System configuration. (b) Signal flow.

Download Full Size | PDF

Remark 1: Note that even in real scenarios where the receiver is relatively fixed on a terminal (e.g., a robot or an auto-cart), there will be height and angle perturbations due to the terminal movement and certain environmental factors (e.g., uneven ground, inaccurate installation and measurement error). Typically, the height perturbation is small while the angle perturbation can be large. Therefore, we consider the case where the receiver height is relatively fixed but the receiver angle can be updated, similarly as in Refs. [2730].

Figure 1(b) shows the signal flow of the considered VLP system. At the transmitter, each LED transmits its own beacon signal containing both DC and AC components to the PD in different time slots to eliminate the mutual interference. In the $l$-th time slot, the PD detects the optical signal from the LED $l$ and converts it into an electrical signal. After removing the DC component via a DC-blocking circuit, the output AC signal strength of PD is given by

$$p_{r,l} = B H_{r,l} p_l + n_{r,l},$$
where $p_l$ denotes the emitted AC signal strength from LED $l$; $H_{r,l}$ denotes the optical channel gain from LED $l$ to the PD; $B$ is the detection gain of PD; and $n_{r,l}$ denotes the additive Gaussian noise with zero mean and variance $\sigma _n^2$. According to the Lambertian model, the optical channel gain $H_{r,l}$ is given by
$$H_{r,l} = \frac{A_r (m_l + 1)}{2 \pi \|\boldsymbol{c}_l - \boldsymbol{c}_r\|_2^2} \cos^{m_l}{(\theta_l)} \cos{(\varphi_l)} {\rm rect}\Big(\frac{\varphi_{l}}{\Phi}\Big),$$
where $m_l$ denotes the Lambertian order of LED $l$; $\theta _l$ is the irradiance angle of LED $l$; $\varphi _l$ denotes the incidence angle; $A_r$ is the detector area of the PD; $\Phi$ is the field-of-view (FOV) of the PD; and ${\rm rect}(\cdot )$ is the rectangular function. We assume a sufficiently large FOV of PD (The FOV of common PDs can be larger than $70^\circ$) such that $\varphi _{l} \leq \Phi, \forall l$. Since all LEDs are pointing downwards, we have $\cos (\theta _l) = \frac {z}{\|\boldsymbol {c}_l - \boldsymbol {c}_r\|_2}$. Considering the tilted PD receiver with orientation vector ${\boldsymbol {u}}_r$, it yields $\cos (\varphi _l) = \frac {(\boldsymbol {c}_l - \boldsymbol {c}_r)^\top \boldsymbol {u}_r}{\|\boldsymbol {c}_l - \boldsymbol {c}_r\|_2}$. Combining the above relationships with Eq. (2), Eq. (1) can be further recast as
$$p_{r,l} = \frac{B A_r (m_l + 1) p_l}{2 \pi} \cdot \frac{z^{m_l} (\boldsymbol{c}_l - \boldsymbol{c}_r)^\top \boldsymbol{u}_r}{\|\boldsymbol{c}_l - \boldsymbol{c}_r\|_2^{m_l + 3}} + n_{r,l}.$$
For convenience of the robust VLP design, we rewrite Eq. (3) as
$$p_{r,l} = \frac{g_l z^{k_l} (\boldsymbol{c}_l - \boldsymbol{c}_r)^\top \boldsymbol{\mu}_r}{\|\boldsymbol{c}_l - \boldsymbol{c}_r\|_2^{k_l}} + n_{r,l},$$
where $k_l = m_l + 3$ represents the attenuation exponent of the LED $l$, $g_l = (m_l + 1) p_l$ denotes the transmitter characteristic scalar, and $\boldsymbol {\mu }_r = \frac {B A_r \boldsymbol {u}_r}{2 \pi z^3}$ denotes the receiver characteristic vector.

Note that parameters $\{ g_l \}_{l=1}^{L}$ and $\{ k_l \}_{l=1}^{L}$ are prior knowledge in a VLP system, because the LEDs are generally deployed in advance and fixed in practical applications. In addition, the values of $\{ g_l \}_{l=1}^{L}$ and $\{ k_l \}_{l=1}^{L}$ may be different for each LED even when all LEDs are driven under the same DC bias and AC signal. As for the receiver characteristic vector $\boldsymbol {\mu }_r$, it is usually unknown since the PD device or its operating states (e.g., orientation angle and detection gain) can be updated in a VLP system, where perfect knowledge of the receiver parameters is typically infeasible.

2.2 Benchmark schemes

In this work, we consider two benchmark VLP schemes, namely the typical least square (LS) scheme and the VLP scheme in Ref. [26] based on the direction of RSS vector (named “RSSD” scheme in this work).

We first present the details of the LS scheme which assumes perfect knowledge on parameters $\{ g_l \}_{l=1}^{L}$, $\{ k_l \}_{l=1}^{L}$ and $\boldsymbol {\mu }_r$ with $\phi =0^\circ$. Under this assumption, we have $\boldsymbol {u}_r=[0,0,1]^\top$ and $(\boldsymbol {c}_l - \boldsymbol {c}_r)^\top \boldsymbol {\mu }_r=\frac {BA_r}{2\pi z^2}$. Ignoring the noise in Eq. (4), the distance $d_{r,l}$ can be approximated as

$$d_{r,l} \approx z \left(\frac{BA_r g_l}{2\pi z^2 p_{r,l}} \right)^{\frac{1}{k_l}}, \quad l = 1,\ldots, L.$$
Substituting relationship $d_{r,l} = \sqrt {(x_l - x_r)^2 + (y_l - y_r)^2 + z^2}$ into Eq. (5), we obtain the following equations
$$\begin{aligned} (x_1 - x_r)^2 + (y_1 - y_r)^2 + z^2 &\approx z^2 \left(\frac{BA_r g_1}{2\pi z^2 p_{r,1}} \right)^{\frac{2}{k_1}},\\ (x_2 - x_r)^2 + (y_2 - y_r)^2 + z^2 &\approx z^2 \left(\frac{BA_r g_2}{2\pi z^2 p_{r,2}} \right)^{\frac{2}{k_2}},\\ &\vdots\\ (x_L - x_r)^2 + (y_L - y_r)^2 + z^2 &\approx z^2 \left(\frac{BA_r g_L}{2\pi z^2 p_{r,L}} \right)^{\frac{2}{k_L}}. \end{aligned}$$
Define $v_{r,l} \triangleq z \left (\frac {BA_r g_l}{2\pi z^2 p_{r,l}} \right )^{\frac {1}{k_l}}$. Subtracting the first equation in Eq. (6) from the last $L-1$ ones, it yields
$$\boldsymbol{A} \boldsymbol{c}_{rh} = \boldsymbol{b},$$
where
$$\begin{aligned} &\boldsymbol{A} = \begin{bmatrix} 2(x_1 - x_2), 2(y_1 - y_2) \\ \cdots \\ 2(x_1 - x_L), 2(y_1 - y_L) \end{bmatrix}, \boldsymbol{c}_{rh} = \begin{bmatrix} x_r \\ y_r \end{bmatrix},\\ &\boldsymbol{b} = \begin{bmatrix} v_{r,2}^2 - v_{r,1}^2 + x_1^2 + y_1^2 - x_2^2 - y_2^2 \\ v_{r,3}^2 - v_{r,1}^2 + x_1^2 + y_1^2 - x_3^2 - y_3^2 \\ \cdots \\ v_{r,L}^2 - v_{r,1}^2 + x_1^2 + y_1^2 - x_L^2 - y_L^2 \end{bmatrix}. \end{aligned}$$
The LS solution to Eq. (7) is given by $\boldsymbol {c}_{rh,LS} = (\boldsymbol {A}^\top \boldsymbol {A})^{-1} \boldsymbol {A}^\top \boldsymbol {b}$ [21]. From Eq. (8), the LS scheme requires the knowledge of all transceiver parameters (e.g., the transceiver coefficients and receiving angle), which is not robust in practical applications.

The RSSD scheme in Ref. [26] jointly estimates the receiver position and receiver coefficient, which is equivalently transformed into the position estimation problem maximizing the projection of the RSS vector $\boldsymbol {p}_r$ on the signal feature vector $\boldsymbol {s}^{fe}=[s_1^{fe},\ldots,s_L^{fe}]^\top$ with $s_l^{fe}=g_l \frac {z^{k_l}}{\|{\boldsymbol {c}}_l - {\boldsymbol {c}}_r \|^{k_l}}$. Consequently, it realizes VLP based on the direction of RSS vector and only requires the knowledge of LED parameters, which is robust to the receiver coefficient. However, it only considers the case of $\phi = 0^\circ$, which limits the application scenarios. In summary, the LS scheme and RSSD scheme are proposed based on strong assumption of perfect knowledge of receiver tilting angle. In real scenarios, the receiver tilting angle can be updated, e.g., in the case of receiver perturbation. As will be shown in Sections 4. and 5, imperfect knowledge of receiver tilting angle has a great impact on the positioning accuracy of the LS scheme and RSSD scheme. To address the mentioned issue, this work considers the VLP problem of estimating the receiver horizontal coordinate $(x_r,y_r)$ from the RSS vector ${\boldsymbol {p}}_r = [p_{r,1}, \ldots, p_{r,L}]^\top \in {\mathbb {R}}^L$ with unknown receiver characteristic vector $\boldsymbol {\mu }_r$.

3. Proposed VLP scheme

In this section, we propose a robust VLP scheme to estimate the receiver position with unknown receiving angle.

3.1 Estimation of receiver position

We consider joint maximum-likelihood (ML) estimation of the receiver horizontal coordinate $(x_r, y_r)$ and receiver characteristic vector $\boldsymbol {\mu }_r$ based on RSS vector ${\boldsymbol {p}}_r$. Mathematically, the joint ML estimation problem can be formulated as

$$\begin{aligned} (\hat{x}_r, \hat{y}_r, \hat{\boldsymbol{\mu}}_r) & = \arg\max_{x_r, y_r, \boldsymbol{\mu}_r} \prod_{l=1}^{L} f (p_{r,l}|x_r, y_r, \boldsymbol{\mu}_r)\\ & = \arg\max_{x_r, y_r, \boldsymbol{\mu}_r} \sum_{l=1}^{L} \ln f (p_{r,l}|x_r, y_r, \boldsymbol{\mu}_r), \end{aligned}$$
where $f(p_{r,l}|x_r, y_r, \boldsymbol {\mu }_r)$ denotes the conditional probability density function (PDF) of $p_{r,l}$ given $(x_r, y_r)$ and $\boldsymbol {\mu }_r$.

Under the assumption of Gaussian noise in Eq. (1), the conditional probability density function $f(p_{r,l}|x_r, y_r, \boldsymbol {\mu }_r)$ is given by

$$f(p_{r,l}|x_r, y_r, \boldsymbol{\mu}_r) \propto \exp \left\{-\frac{(p_{r, l} - \boldsymbol{s}_l \boldsymbol{\mu}_r)^2}{2 \sigma_n^2} \right\},$$
where ${\boldsymbol {s}}_l = g_l \frac {z^{k_l} (\boldsymbol {c}_l - \boldsymbol {c}_r)}{\|\boldsymbol {c}_l - \boldsymbol {c}_r\|_2^{k_l}} \in {\mathbb {R}}^3$. Define the measurement matrix ${\boldsymbol {S}} = [\boldsymbol {s}_1,\ldots, \boldsymbol {s}_L]^\top \in {\mathbb {R}} ^{L\times 3}$. Using Eq. (10), problem (9) is equivalent to
$$(\hat{x}_r, \hat{y}_r, \hat{\boldsymbol{\mu}}_r) = \arg\min_{x_r, y_r, \boldsymbol{\mu}_r} \| \boldsymbol{p}_r - \boldsymbol{S} \boldsymbol{\mu}_r\|_2^2.$$
Note that problem (11) can be equivalently decomposed into the following two-level problem:
$$\begin{aligned} (\hat{x}_r, \hat{y}_r, \hat{\boldsymbol{\mu}}_r) &= \arg\min_{x_r, y_r, \boldsymbol{\mu}_r} \| \boldsymbol{p}_r - \boldsymbol{S} \boldsymbol{\mu}_r\|_2^2\\ &= \arg\min_{x_r, y_r}\min_{\boldsymbol{\mu}_r} \| \boldsymbol{p}_r - \boldsymbol{S} \boldsymbol{\mu}_r\|_2^2. \end{aligned}$$
At the lower level, we have the following subproblem from problem (12) with fixed $(x_r, y_r)$:
$$\boldsymbol{\mu}_r^* = \arg\min_{\boldsymbol{\mu}_r} \| \boldsymbol{p}_r - \boldsymbol{S} \boldsymbol{\mu}_r\|_2^2.$$
The solution to problem (13) is readily obtained as $\boldsymbol {\mu }_r^* = (\boldsymbol {S}^\top \boldsymbol {S})^{-1} \boldsymbol {S}^\top \boldsymbol {p}_r$ [31]. At the higher level, we have the following master problem which updates $(x_r, y_r)$:
$$\begin{aligned} (\hat{x}_r, \hat{y}_r) &= \arg\min_{x_r, y_r} \| \boldsymbol{p}_r - \boldsymbol{S} \boldsymbol{\mu}_r^*\|_2^2\\ &= \arg\min_{x_r, y_r} \| \boldsymbol{p}_r - \boldsymbol{S} (\boldsymbol{S}^\top \boldsymbol{S})^{{-}1} \boldsymbol{S}^\top \boldsymbol{p}_r \|_2^2\\ &= \arg\max_{x_r, y_r} \| \boldsymbol{S} (\boldsymbol{S}^\top \boldsymbol{S})^{{-}1} \boldsymbol{S}^\top \boldsymbol{p}_r \|_2^2. \end{aligned}$$
Consequently, the optimal solution obtained in Eqs. (13) and14 can reach the global minimum of $\| \boldsymbol {p}_r - \boldsymbol {S} \boldsymbol {\mu }_r\|_2^2$ in Eq. (11) [32]. We can see that the objective of the proposed VLP scheme in Eq. (14) is to maximize the projection of the RSS vector $\boldsymbol {p}_r$ on the column space spanned by $\boldsymbol {S}$ with projection matrix $\boldsymbol {S} (\boldsymbol {S}^\top \boldsymbol {S})^{-1} \boldsymbol {S}^\top$. In other words, the column space spanned by the measurement matrix $\boldsymbol {S}$ is adopted as the key feature for estimating the receiver horizontal coordinate $(x_r, y_r)$. In contrast, the RSSD scheme estimates $(x_r, y_r)$ by maximizing the projection of the RSS vector $\boldsymbol {p}_r$ on the signal feature vector $\boldsymbol {s}^{fe}=[s_1^{fe},\ldots,s_L^{fe}]^\top$ with $s_l^{fe}=g_l \frac {z^{k_l}}{\|{\boldsymbol {c}}_l - {\boldsymbol {c}}_r \|^{k_l}}$. Compared with the RSSD scheme, the proposed scheme is more robust to the variations of receiver characteristics since matrix $\boldsymbol {S}$ is a more generalized feature than vector $\boldsymbol {s}^{fe}$. Note that the proposed VLP scheme can work with only the ratios of $\{g_l\}_{l=1}^L$ instead of the exact values of $\{g_l \boldsymbol {\mu }_r\}_{l=1}^L$. Consequently, it is robust to both the receiving coefficient and receiving angle, making the proposed scheme more practical in real VLP scenarios, e.g., where the receiver detection gain or orientation angle varies.

It is seen from Eq. (14) that the proposed scheme requires that the measurement matrix $\boldsymbol {S}$ is full column rank. Regarding this requirement, we have the following results.

Theorem 1 With all LEDs located at the same height, matrix $\boldsymbol {S}$ is full column rank if and only if the $L$ LEDs are non-collinear.

Proof: Please see Appendix. $\hfill\blacksquare$

Theorem 1 implies that the $L$ LEDs should be non-collinear to ensure the proposed VLP scheme is feasible. Moreover, it is noteworthy that when matrix $\boldsymbol {S}$ is full rank with $L=3$, $\boldsymbol {S} \in \mathbb {R}^{3\times 3}$ has a unique inverse matrix $\boldsymbol {S}^{-1}$. Then, the objective function in problem (14) can be rewritten as

$$\| \boldsymbol{S} (\boldsymbol{S}^\top \boldsymbol{S})^{{-}1} \boldsymbol{S}^\top \boldsymbol{p}_r \|_2^2 = \| \boldsymbol{S} (\boldsymbol{S}^{{-}1} \boldsymbol{S}^{-\top}) \boldsymbol{S}^\top \boldsymbol{p}_r \|_2^2 = \| \boldsymbol{p}_r\|_2^2,$$
which is a constant irrelevant to $(x_r, y_r)$, and results in a positioning failure. In summary, the following two conditions should be satisfied to guarantee the feasibility of the proposed VLP scheme: i) $L \ge 4$; ii) The $L$ LEDs are non-collinear. The above two conditions can be easily satisfied in practical applications.

Remark 2: Note that the positioning accuracy increases with the number of LEDs, since signals from additional LEDs can provide more information about the receiver position. To achieve a good complexity-performance tradeoff, it is essential to properly choose the number of LEDs in real VLP scenarios. Based on the results in Section 4.2, we can choose 5 LEDs to realize a relatively high-accuracy positioning.

3.2 IS method for solving the VLP problem

Problem (14) is non-convex in $(x_r, y_r)$ and it is difficult to obtain its optimal solution. One straightforward solution is to adopt the BFS method in the 2-dimensional positioning area. Specifically, we divide the positioning area into multiple grid points, and choose the grid point yielding the largest value of $||\boldsymbol {S} (\boldsymbol {S}^\top \boldsymbol {S})^{-1} \boldsymbol {S}^\top \boldsymbol {p}_r||_2^2$ as the estimated receiver position. However, the computational complexity of BFS method can be very high under large number of divided grid points, which may lead to high latency for VLP in real scenarios. To reduce the computational complexity, we propose a Monte Carlo method for solving problem (14) based on the Pincus’ theorem [33] and IS method [3436].

Pincus’ theorem: For the optimization problem $\max _{\boldsymbol {c} \in \mathbb {R}^n} F(\boldsymbol {c})$ with $\boldsymbol {c}= [c_1, \ldots, c_n]^\top$, its optimal solution is given by

$$\hat{c}_i = \lim_{\lambda \rightarrow \infty} \frac{\int \cdots \int_S c_i \exp[\lambda F(\boldsymbol{c})] {\text{d}}\boldsymbol{c}}{\int \cdots \int_S \exp[\lambda F(\boldsymbol{c})] {\text{d}}\boldsymbol{c}}, \quad i=1, \ldots, n.$$
Define
$$f (\boldsymbol{c}) = \frac{\exp[\lambda F(\boldsymbol{c})]}{\int \cdots \int_S \exp[\lambda F(\boldsymbol{c})] {\text{d}}\boldsymbol{c}},$$
which can be considered as a PDF of $\boldsymbol {c}$. Then, the solution ${\hat {\boldsymbol {c}}}$ in Eq. (16) is the mean of ${\boldsymbol {c}}$ under the PDF $f(\boldsymbol {c})$ with $\lambda \rightarrow \infty$, which is denoted by $\mathbb {E}_{f} \{ \boldsymbol {c} \}$.

Particularly, in the VLP problem (14), we have $\boldsymbol {c}_{rh} = [x_r, y_r]^\top$ and

$$\begin{aligned} F (\boldsymbol{c}_{rh}) &= \| \boldsymbol{S} (\boldsymbol{S}^\top \boldsymbol{S})^{{-}1} \boldsymbol{S}^\top \boldsymbol{p}_r \|_2^2\\ &= \boldsymbol{p}_r^\top \boldsymbol{S} (\boldsymbol{S}^\top \boldsymbol{S})^{{-}1} \boldsymbol{S}^\top \boldsymbol{p}_r. \end{aligned}$$
According to the Pincus’ theorem, the solution to problem (13) is the mean $\mathbb {E}_{f} \{\boldsymbol {c}_{rh}\}$ under the PDF $f(\boldsymbol {c}_{rh})$. With $F(\boldsymbol {c}_{rh})$ in Eq. (18), the closed-form expression for $\mathbb {E}_{f} \{ \boldsymbol {c}_{rh}\}$ is not available. A straightforward Monte Carlo approximation is $\mathbb {E}_{f} \{ \boldsymbol {c}_{rh} \} \approx \frac {1}{T} \sum _{t=1}^T \boldsymbol {c}_{rh}^{(t)}$ with $T$ independent and identically distributed (i.i.d.) samples $\{\boldsymbol {c}_{rh}^{(1)},\ldots, \boldsymbol {c}_{rh}^{(T)}\} \sim f(\boldsymbol {c}_{rh})$. However, sampling directly from $f(\boldsymbol {c}_{rh})$ is difficult. Instead, we adopt IS method to obtain an approximation of $\mathbb {E}_{f} \{\boldsymbol {c}_{rh}\}$. Consider another PDF $q(\boldsymbol {c}_{rh})$ whose samples can be easily generated. According to the concept of IS, we adopt $T$ i.i.d. samples $\{\boldsymbol {c}_{rh}^{(1)},\ldots, \boldsymbol {c}_{rh}^{(T)}\} \sim q(\boldsymbol {c}_{rh})$ to numerically approximate $\mathbb {E}_{f} \{\boldsymbol {c}_{rh}\}$ as
$$\mathbb{E}_{f} \{ \boldsymbol{c}_{rh} \} \approx \frac{\sum_{t=1}^T w(\boldsymbol{c}_{rh}^{(t)}) \boldsymbol{c}_{rh}^{(t)}}{\sum_{t=1}^T w(\boldsymbol{c}_{rh}^{(t)})},$$
where $w(\boldsymbol {c}_{rh}) \propto \frac {f(\boldsymbol {c}_{rh})}{q(\boldsymbol {c}_{rh})}$ is the importance weight. One advantage of Eq. (19) is that only an unnormalized version of $f(\boldsymbol {c}_{rh})$ is required for calculating $w(\boldsymbol {c}_{rh})$.

Derivation of Sampling Distribution: The major challenge of implementing IS for calculating $\mathbb {E}_{f} \{\boldsymbol {c}_{rh}\}$ is to find a proper PDF $q(\boldsymbol {c}_{rh})$ which admits an efficient sampling. We consider the following simplifications of $F (\boldsymbol {c}_{rh})$ in (18):

$$\boldsymbol{p}_r^\top \boldsymbol{S} (\boldsymbol{S}^\top \boldsymbol{S})^{{-}1} \boldsymbol{S}^\top \boldsymbol{p}_r \stackrel{(a)}{\rightarrow} \boldsymbol{p}_r^\top \boldsymbol{S} \boldsymbol{S}^\top \boldsymbol{p}_r \stackrel{(b)}{\rightarrow} \boldsymbol{p}_r^\top \widetilde{\boldsymbol{S}} \widetilde{\boldsymbol{S}}^\top \boldsymbol{p}_r,$$
where $(\boldsymbol {S}^\top \boldsymbol {S})^{-1}$ is replaced with the identity matrix ${\boldsymbol {I}}_{3 \times 3}$ in step $(a)$, $\boldsymbol {S}$ is simplified as $\widetilde {\boldsymbol {S}}$ in step $(b)$, and $\widetilde {\boldsymbol {S}}$ is given by
$$\begin{aligned} \widetilde{\boldsymbol{S}} = \begin{bmatrix} x_1 - x_r & y_1 - y_r & z \\ \vdots & \vdots & \vdots \\ x_L - x_r & y_L - y_r & z \end{bmatrix} \triangleq [\widetilde{\boldsymbol{s}}_{x_r}, \widetilde{\boldsymbol{s}}_{y_r}, \widetilde{\boldsymbol{s}}_z]. \end{aligned}$$
After the two simplification steps in Eq. (20), $\boldsymbol {p}_r^\top \widetilde {\boldsymbol {S}} \widetilde {\boldsymbol {S}}^\top \boldsymbol {p}_r$ is separable in $x_r$ and $y_r$. Then, we seek to generate samples of $\boldsymbol {c}_{rh}$ from the distribution
$$\begin{aligned}q(\boldsymbol{c}_{rh}) &\propto \exp[\lambda_1 \boldsymbol{p}_r^\top \widetilde{\boldsymbol{S}} \widetilde{\boldsymbol{S}}^\top \boldsymbol{p}_r]\\ &\propto \exp[ \lambda_1 |\widetilde{\boldsymbol{s}}_{x_r}^\top \boldsymbol{p}_r|^2] \cdot \exp[ \lambda_1 |\widetilde{\boldsymbol{s}}_{y_r}^\top \boldsymbol{p}_r|^2]\\ &= \widetilde{q}_x(x_r) \cdot \widetilde{q}_y(y_r), \end{aligned}$$
where $\widetilde {q}_x(x_r) = \exp [ \lambda _1 |\widetilde {\boldsymbol {s}}_{x_r}^\top \boldsymbol {p}_r|^2]$ and $\widetilde {q}_y(y_r) = \exp [ \lambda _1 |\widetilde {\boldsymbol {s}}_{y_r}^\top \boldsymbol {p}_r|^2]$. The joint PDF $q(\boldsymbol {c}_{rh})$ can be written as the product of two marginal PDFs. As a result, generating samples of $\boldsymbol {c}_{rh}$ reduces to generating independent samples of $x_r$ and $y_r$, which can be performed in parallel to further improve the computational efficiency in practical applications. Specifically, samples of $x_r$ are generated from the distribution $q_x(x_r) \propto \widetilde {q}_x (x_r)$, and samples of $y_r$ are generated from the distribution $q_y(y_r) \propto \widetilde {q}_y (y_r)$. We can easily obtain samples $\{x_r^{(t)} \} \sim q_x (x_r)$ and samples $\{y_r^{(t)}\} \sim q_y (y_r)$ by the inversion of cumulative distribution function (CDF) [37]. According to (19), we can use $T$ i.i.d. samples $\{\boldsymbol {c}_{rh}^{(1)}, \ldots, \boldsymbol {c}_{rh}^{(T)} \} \sim q(\boldsymbol {c}_{rh}) \propto \widetilde {q}_x (x_r) \cdot \widetilde {q}_y (y_r)$ to approximately calculate $\mathbb {E}_{f} \{ \boldsymbol {c}_{rh} \}$ as
$$\hat{x}_r \approx \frac{\sum_{t=1}^T w (\boldsymbol{c}_{rh}^{(t)}) x_r^{(t)}}{\sum_{t=1}^T w (\boldsymbol{c}_{rh}^{(t)})}, \quad \hat{y}_r \approx \frac{\sum_{t=1}^T w (\boldsymbol{c}_{rh}^{(t)}) y_r^{(t)}}{\sum_{t=1}^T w (\boldsymbol{c}_{rh}^{(t)})},$$
where $w (\boldsymbol {c}_{rh}^{(t)})$ is the importance weight given by
$$w (\boldsymbol{c}_{rh}^{(t)}) = \frac{\exp[ \lambda {F} (\boldsymbol{c}_{rh}^{(t)})]}{\widetilde{q}_x (x_r^{(t)}) \cdot \widetilde{q}_y (y_r^{(t)})} \propto \frac{f (\boldsymbol{c}_{rh}^{(t)})}{q(\boldsymbol{c}_{rh}^{(t)})}.$$
The overall procedure of the proposed VLP scheme with IS is summarized in Algorithm 1.

Tables Icon

Algorithm 1. The proposed VLP scheme with IS

3.3 Estimation of LED parameters

To implement the proposed VLP scheme, we need to know some LED parameters, e.g., the values of $\{k_l\}_{l=1}^L$ and ratios of $\{g_l\}_{l=1}^L$. These LED parameters can be considered as prior information of VLP and can be estimated using the received signals of one PD (“PD-EST”) at $I$ pre-known positions. The detailed procedure of LED parameter estimation can be found in Ref. [26] and is omitted here. With the estimated LED parameters, robust VLP can be realized via Algorithm 1 using a positioning PD (“PD-POS”, not the same one as PD-EST) in the case of unknown receiving coefficient and receiving angle.

3.4 Complexity analysis

In this subsection, we analyze the computational complexity of the VLP schemes. The complexity of the LS scheme is $\mathcal {O} (L)$ due to the matrix computation. Letting $M$ denote the number of divided grid points in the receiver plane, the complexities of the RSSD scheme in Ref. [26] and the VLP scheme (14) with BFS are both $\mathcal {O} (M \cdot L)$. For the VLP scheme (14) with IS, it requires a complexity of $\mathcal {O} (T \cdot L)$ with $T$ being the number of generated samples. Since $T$ in IS can be much lower than $M$ in BFS, the complexity of the VLP scheme (14) with IS can be significantly reduced compared to that with BFS.

4. Simulation results and discussions

In this section, the performance of the proposed VLP scheme is verified via simulation results.

4.1 Simulation setup

The simulations are carried out using MATLAB R2020b on a personal computer with 2 GHz Intel Core i5 CPU and 16 GB memory. In the simulations, we consider the scenario where $L=5$ LEDs are installed in the ceiling of a room. The coordinates of the five LEDs are $(\pm 1, \pm 1, 2)$ m and $(0, 0, 2)$ m. The LED parameters are set to be $g_1 = \cdots = g_5 = 1$ and $k_1 = \cdots = k_5 = 5$. The PD is located in the $2\, {\rm m} \times 2 \,{\rm m}$ horizontal plane with height $z_r = 0$ m. The vertical tilting angle $\phi$ of PD is chosen from set $\{0^\circ, 5^\circ, 10^\circ, 15^\circ, 20^\circ, 25^\circ \}$, and the horizontal tilting angle $\beta$ is uniformly chosen from $[0^\circ, 360^\circ )$ at each trial. The detector area of the PD is $A_r= 7.07 \, \rm {mm}^2$ and detection gain is $B=1$. The signal-to-noise ratio (SNR) is defined as ${\rm SNR} = 10\lg \frac {BA_rg_l}{2\pi z^2 n_{r,l}}[\rm {dB}]$, which adopts the receiver point directly below each LED as the reference point. Unless otherwise specified, the SNR is set to be 50 dB. The number of grid points in BFS is $M=40000$, and the sample size in IS is $T=1000$. For comparison, we also present the results of the LS scheme and RSSD scheme mentioned in Section 2.2. In this work, the mean positioning error is defined using mean absolute deviation, i.e., $MPE=\mathbb {E}\left \{\sqrt {(x_r-\hat {x}_r)^2+(y_r-\hat {y}_r)^2} \right \}$.

4.2 Simulation results

We first investigate the positioning accuracy of the proposed VLP scheme using BFS and IS methods. Figure 2 compares the CDF of positioning errors using the two methods (labeled as “Prop-BFS” and “Prop-IS”). We consider two cases of PD vertical tilting angles, namely $\phi =0^\circ$ and $\phi =15^\circ$. The CDF results are obtained using 100 PD locations which are uniformly distributed in the 2 m $\times$ 2 m horizontal plane. At each test point, 100 trials are adopted to obtain a statistical result. It can be observed that the IS method has a similar CDF of positioning errors as the BFS method. The mean positioning errors of BFS and IS methods are 3.7 cm and 4.0 cm (4.3 cm and 4.5 cm) when $\phi =0^\circ$ ($\phi =15^\circ$). It implies that the BFS and IS methods yield close positioning accuracy. In the simulations, the average running time of the BFS method is 0.79 s for each positioning point, while that of the IS method is 0.03 s for each positioning point. Compared to the BFS method, the IS method can achieve a close positioning accuracy with much lower complexity. On the whole, the IS method can achieve a good complexity-performance tradeoff and is appealing for practical applications.

 figure: Fig. 2.

Fig. 2. CDF of positioning errors with BFS and IS methods under different tilting angles.

Download Full Size | PDF

To investigate the robustness of the proposed VLP scheme to the variation of receiving angle, Fig. 3 plots the simulated positioning error versus PD vertical tilting angle $\phi$. Since actual value of $\phi$ is unknown, assume that LS and RSSD schemes adopt $\phi = 0^\circ$. It can be seen that the mean positioning error of the proposed scheme (with BFS or IS) increases slowly with $\phi$, which results from the decrease of the received SNR. On the contrary, the positioning errors of the LS and RSSD schemes increase rapidly as $\phi$ increases, due to imperfect knowledge of the receiver orientation angle. The results imply that the proposed VLP scheme is much more robust to the receiving angle than the LS and RSSD schemes. We can also see that under small $\phi$, the mean positioning error of the proposed scheme is a little higher than those of the LS and RSSD schemes. The reason is that (approximately) perfect knowledge of the PD tilting angle by the LS or RSSD scheme is beneficial for improving the positioning accuracy.

 figure: Fig. 3.

Fig. 3. Simulated positioning error versus PD vertical tilting angle $\phi$.

Download Full Size | PDF

To investigate the effect of receiver coefficient mismatch on the positioning accuracy, we consider the scalar error model of the receiver characteristic vector: ${\boldsymbol {\mu }}_r = \alpha {\boldsymbol {\mu }}_{r0}$, where ${\boldsymbol {\mu }}_{r0}$ is a known vector, $\alpha$ is an unknown scalar, and ${\boldsymbol {\mu }}_{r}$ is the unknown true vector. In the positioning, ${\boldsymbol {\mu }}_{r0}$ is used instead of ${\boldsymbol {\mu }}_{r}$. Figure 4 shows the simulated positioning error versus the unknown scalar $\alpha$ under $\phi = 0^\circ$. It can be observed that the positioning error of the proposed scheme (with BFS or IS) decreases with $\alpha$ due to the increase of received SNR. Similar results are also observed for RSSD scheme. In contrast, the positioning accuracy of the LS scheme deteriorates sharply under severe mismatch of scalar $\alpha$. The results in Fig. 3 and Fig. 4 demonstrate the robustness of the proposed scheme to the receiver characteristics including the receiving angle and receiving coefficient.

 figure: Fig. 4.

Fig. 4. Simulated positioning error versus the unknown scalar $\alpha$ under $\phi = 0^\circ$.

Download Full Size | PDF

Figure 5 shows the estimated positions of 100 test points under $\phi =0^\circ$ and $L = 4, 5$. When $L =4$, the coordinates of the LEDs are $(\pm 1, \pm 1, 2)\, {\rm m}$; When $L =5$, the coordinates of the LEDs are $(\pm 1, \pm 1, 2)\, {\rm m}$ and $(0, 0, 2)\, {\rm m}$. In Fig. 5, 100 trials are adopted at each test point to obtain a statistical result. Note that ML estimator is an asymptotically unbiased estimator (e.g., as the number of samples approaches infinity) rather than an exactly unbiased one [38]. Therefore, there exists estimation biases in VLP for small number of LEDs. Similar results can also be observed in Refs. [3941]. It is seen that the positioning accuracy using 4 LEDs is much worse than that using 5 LEDs. Similar results can be obtained for other tilting angles. The results provide guidance for us to adopt 5 LEDs in real positioning scenarios.

 figure: Fig. 5.

Fig. 5. Estimated positions of 100 test points under $\phi =0^\circ$. (a) BFS. (b) IS.

Download Full Size | PDF

5. Experimental results and discussions

In this section, experimental results are presented to further evaluate the performance of the proposed VLP scheme.

5.1 Experimental setup

The experimental environment is shown in Fig. 6. In the experiments, we adopt 5 LEDs (Kecent SL-COB) with coordinates $(\pm 0.3, \pm 0.3, 1.5)$ m and $(0, 0, 1.5)$ m. All LEDs are driven by the same DC bias and sinusoid signal using a bias-tee circuit. An AC-coupled avalanche photodiode (APD, C12702) receives the optical signals and outputs the electrical AC signals. A real-time oscilloscope (Agilent Technologies MSO-X 6004A) is used to collect the output AC signals of APD for further processing performed with MATLAB. We first estimate the LED parameters $\{g_l\}_{l=1}^L$ and $\{k_l\}_{l=1}^L$ using one APD (“PD-EST”) with $\phi = 0^\circ$. Then, we locate the position of one test APD (“PD-POS”) based on the estimated LED parameters. The VLP experiments are conducted under two vertical tilting angles of PD-POS, namely $\phi = 0^\circ$ (Case 1) and $\phi = 13^\circ$ (Case 2), and the horizontal tilting angle $\beta$ is not fixed at each test point as it is difficult to accurately measure $\beta$ in real scenarios. The SNR in the experiments is about 35 dB.

 figure: Fig. 6.

Fig. 6. The experimental environment.

Download Full Size | PDF

We measure the output AC signals at 49 receiver points of PD-POS. Figure 7 depicts the horizontal positions of LEDs and test receiver points. The relevant experimental parameters are listed in Table 1.

 figure: Fig. 7.

Fig. 7. The horizontal positions of LEDs and test receiver points.

Download Full Size | PDF

Tables Icon

Table 1. The Experimental Parameters for Indoor VLP System

5.2 Experimental results

Recall that the parameter $I$ represents the number of pre-known positions of PD-EST used for LED parameter estimation. Table 2 lists the estimated values of LED parameters under different $I$. To study the impact of number $I$ on the positioning accuracy, Fig. 8 shows the mean positioning errors of the proposed scheme (with BFS or IS) under different $I$. It can be seen that the mean positioning error decreases with $I$. Specifically, when $I=13$, the mean positioning errors of the proposed scheme are both below 7 cm in Case 1 and Case 2. To guarantee a high positioning accuracy, we adopt $I=13$ by default in the following results.

 figure: Fig. 8.

Fig. 8. Mean positioning errors of the proposed scheme under different $I$.

Download Full Size | PDF

Tables Icon

Table 2. Estimated Values of LED Parameters under different $I$

Table 3 illustrates the difference between the normalized RSS in Case 1 and fitting values using the estimated LED parameters with $I = 13$. We can see that the normalized RSS is very close to the fitting values with a maximum difference below 0.03. It verifies the accuracy of the signal model given by Eq. (4).

Tables Icon

Table 3. Difference between the normalized RSS in Case 1 and fitting values using the estimated LED parameters with $I = 13$

In the following, we compare the positioning accuracy of different VLP schemes. The LS and RSSD schemes simply assume $\phi = 0^\circ$ in Cases 1 and 2. Figure 9 shows the CDF of positioning errors in Cases 1 and 2. The CDF results are obtained using 49 PD locations shown in Fig. 7. At each test point, 10 trials are adopted to obtain a statistical result. We can see that the LS scheme and RSSD scheme are both infeasible in Case 2 with imperfect knowledge of the receiver tilting angle. Specifically, the maximum positioning errors of the LS scheme and RSSD scheme are both about 45 cm. In contrast, the positioning accuracy of the proposed scheme (with BFS or IS) in Case 2 is close to that in Case 1 with a maximum positioning error of 12 cm, demonstrating its robustness to the receiving angle. Note that there exists a performance gap between the experimental results and simulation results, which can be explained as follows. Firstly, the received SNR in the simulations is higher than that in the experiments, yielding higher positioning accuracy. Secondly, there can be measurement error and parameter estimation error in the experiments, resulting in a positioning accuracy degradation. Therefore, it is reasonable that the positioning accuracy in the experiments is worse than that in the simulations. Similar results can be observed in Refs. [26,42,43].

 figure: Fig. 9.

Fig. 9. CDF of positioning errors in Cases 1 and 2.

Download Full Size | PDF

To better compare the simulation results and experimental results, Table 4 lists the mean positioning errors obtained in the simulation and experiment under experimental parameters. The coordinates of LEDs are ($\pm$0.3, $\pm$0.3, 1.5) m and (0, 0, 1.5) m. The SNR is about 35 dB. It is seen from Table 4 that the positioning errors in the experiment are a little larger than those in the simulation due to the measurement error and parameter estimation error in the experiment.

Tables Icon

Table 4. Mean positioning errors in the simulation and experiment under experimental parameters

We further investigate the effects of receiver height and angle perturbations on the positioning accuracy. With actual transceiver height difference $z = 150$ cm, Fig. 10 shows the mean positioning errors under different transceiver height assumptions in Cases 1 and 2. It can be observed that the positioning errors of all VLP schemes vary little under small mismatch of transceiver height difference, implying that the receiver height disturbance has small influence on the positioning accuracy. Also note that, the proposed scheme (with BFS or IS) retains a stable positioning accuracy below 7 cm in Cases 1 and 2. In contrast, the mean positioning error of the LS or RSSD scheme in Case 2 is much higher than that in Case 1. It implies that the LS and RSSD schemes are very sensitive to the variation of receiver tilting angle. The results demonstrate the robustness of the proposed VLP scheme under receiver height and angle disturbances.

 figure: Fig. 10.

Fig. 10. Mean positioning errors under different transceiver height assumptions in Cases 1 and 2.

Download Full Size | PDF

6. Conclusion

In this paper, we have proposed a novel RSS-based VLP scheme with unknown receiving angles. By jointly estimating the receiver coordinate and receiver characteristic vector, the proposed scheme is robust to the receiving coefficient and receiving angle. To further reduce the computational complexity, we have also proposed an IS method for solving the VLP problem. Simulation and experimental results demonstrate the robustness of the proposed scheme. Overall, the proposed scheme can provide a good performance-complexity tradeoff and is appealing for practical applications. Note that this work assumes linear response of PD. In practical scenarios, the PD may suffer nonlinear behavior under dynamic input signals. Robust design of RSS-based VLP under PD nonlinearity remains for our future work.

7. Appendix: proof of theorem 1

According to the definition of $\boldsymbol {S}$ in Section 3, we have

$$\boldsymbol{S} = \begin{bmatrix} g_1 \frac{z^{k_1}}{d_{r,1}^{k_1}} (x_1 - x_r) & g_1 \frac{z^{k_1}}{d_{r,1}^{k_1}} (y_1 - y_r) & g_1 \frac{z^{k_1}}{d_{r,1}^{k_1}} \cdot z\\ \vdots & \vdots & \vdots \\ g_L \frac{z^{k_L}}{d_{r,L}^{k_L}} (x_L - x_r) & g_L \frac{z^{k_L}}{d_{r,L}^{k_L}} (y_L - y_r) & g_L \frac{z^{k_L}}{d_{r,L}^{k_L}} \cdot z \end{bmatrix} = \begin{bmatrix} g_1 \frac{z^{k_1}}{d_{r,1}^{k_1}} & & \\ & \ddots & \\ & & g_L \frac{z^{k_L}}{d_{r,L}^{k_L}} \end{bmatrix} \cdot \widetilde{\boldsymbol{S}},$$
where $\widetilde {\boldsymbol {S}}$ is given by Eq. (21). Using (25), we obtain ${\rm rank}(\boldsymbol {S}) = {\rm rank}(\widetilde {\boldsymbol {S}})$.

Subtracting the first row of $\widetilde {\boldsymbol {S}}$ from the last $L-1$ ones, it yields

$$ \widetilde{\boldsymbol{S}}=\left[\begin{array}{ccc} x_1-x_r & y_1-y_r & z \\ x_2-x_r & y_2-y_r & z \\ \vdots & \vdots & \vdots \\ x_L-x_r & y_L-y_r & z \end{array}\right] \rightarrow\left[\begin{array}{cc|c} x_1-x_r & y_1-y_r & z \\ \hline x_2-x_1 & y_2-y_1 & 0 \\ \vdots & \vdots & \vdots \\ x_L-x_1 & y_L-y_1 & 0 \end{array}\right] \triangleq \widetilde{\boldsymbol{S}^{\prime}}=\left[\begin{array}{c|c} \widetilde{\boldsymbol{S}_1^{\prime}} & z \\ \hline \widetilde{\widetilde{S}_2^{\prime}} & \mathbf{0} \end{array}\right] . $$
Using (26), we obtain ${\rm rank}(\widetilde {\boldsymbol {S}}) = {\rm rank}(\widetilde {\boldsymbol {S}}^\prime ) = {\rm rank}(\widetilde {\boldsymbol {S}}^\prime _2) + 1$. When the $L$ LEDs are collinear, we have ${\rm rank}(\widetilde {\boldsymbol {S}}^\prime _2) = 1$ and ${\rm rank}(\boldsymbol {S}) ={\rm rank}(\widetilde {\boldsymbol {S}}) = 2$; When the $L$ LEDs are non-collinear, we have ${\rm rank}(\widetilde {\boldsymbol {S}}^\prime _2) = 2$ and ${\rm rank}(\boldsymbol {S}) ={\rm rank}(\widetilde {\boldsymbol {S}}) = 3$. The two cases of collinear and non-collinear LEDs are illustrated in Fig. 11(a) and Fig. 11(b), respectively. Therefore, ${\boldsymbol {S}}$ is full column rank if and only if LEDs are non-collinear.

 figure: Fig. 11.

Fig. 11. Illustration of the LED layout. (a) All LEDs are collinear. (b) LEDs are non-collinear.

Download Full Size | PDF

Funding

National Natural Science Foundation of China (62101526, 62171428); Key Program of National Natural Science Foundation of China (61631018); Key Research Program of Frontier Sciences of CAS (QYZDY-SSW-JSC003); Fundamental Research Funds for the Central Universities (KY2100000118); open research fund of National Mobile Communications Research Laboratory Southeast University (2019D14).

Disclosures

The authors declare no conflicts of interest.

Data Availability

Data underlying the results presented in this paper are not publicly available at this time but may be obtained from the authors upon reasonable request.

References

1. Y. Chen, T. Li, Y. Cui, Z. Li, M. Gao, and L. Zhou, “Research on indoor and outdoor integrated location service technology,” Proc. International Conference on Intelligent Transportation, Big Data & Smart City (ICITBS), (2018), pp. 117–120.

2. N. Saeed, H. Nam, T. Y. Al-Naffouri, and M.-S. Alouini, “A state-of-the-art survey on multidimensional scaling-based localization techniques,” IEEE Commun. Surv. Tutor. 21(4), 3565–3583 (2019). [CrossRef]  

3. O. O. Erunkulu, A. M. Zungeru, C. K. Lebekwe, M. Mosalaosi, and J. M. Chuma, “5G mobile communication applications: A survey and comparison of use cases,” IEEE Access 9, 97251–97295 (2021). [CrossRef]  

4. J. Du and M. J. Barth, “Next-generation automated vehicle location systems: Positioning at the lane level,” IEEE Trans. Intell. Transp. Syst. 9(1), 48–57 (2008). [CrossRef]  

5. M. Ochiai, M. Fujii, A. Ito, Y. Watanabe, and H. Hatano, “A study on indoor position estimation based on fingerprinting using GPS signals,” Proc. International Conference on Indoor Positioning and Indoor Navigation (IPIN), (2014), pp. 727–728.

6. Y. Zhuang, Z. Syed, Y. Li, and N. El-Sheimy, “Evaluation of two WiFi positioning systems based on autonomous crowdsourcing of handheld devices for indoor navigation,” IEEE Trans. Mob. Comput. 15(8), 1982–1995 (2016). [CrossRef]  

7. N. Jovanovic, T. Ozcelebi, and J. Lukkien, “Indoor user positioning using infrared LEDs and sensors,” Proc. International Conference on Indoor Positioning and Indoor Navigation (IPIN), (2014), pp. 400–406.

8. A. De Angelis, A. Moschitta, P. Carbone, M. Calderini, S. Neri, R. Borgna, and M. Peppucci, “Design and characterization of a portable ultrasonic indoor 3-D positioning system,” IEEE Trans. Instrum. Meas. 64(10), 2616–2625 (2015). [CrossRef]  

9. R. Faragher and R. Harle, “Location fingerprinting with bluetooth low energy beacons,” IEEE J. Sel. Areas Commun. 33(11), 2418–2428 (2015). [CrossRef]  

10. X. Zhu, J. Yi, J. Cheng, and L. He, “Adapted error map based mobile robot UWB indoor positioning,” IEEE Trans. Instrum. Meas. 69(9), 6336–6350 (2020). [CrossRef]  

11. C.-Y. Yao and W.-C. Hsia, “An indoor positioning system based on the dual-channel passive RFID technology,” IEEE Sens. J. 18(11), 4654–4663 (2018). [CrossRef]  

12. P. H. Pathak, X. Feng, P. Hu, and P. Mohapatra, “Visible light communication, networking, and sensing: A survey, potential and challenges,” IEEE Commun. Surv. Tutor. 17(4), 2047–2077 (2015). [CrossRef]  

13. Y. Zhuang, L. Hua, L. Qi, J. Yang, P. Cao, Y. Cao, Y. Wu, J. Thompson, and H. Haas, “A survey of positioning systems using visible LED lights,” IEEE Commun. Surv. Tutor. 20(3), 1963–1988 (2018). [CrossRef]  

14. Z. Cao, M. Cheng, Q. Yang, M. Tang, D. Liu, and L. Deng, “Experimental investigation of environmental interference mitigation and blocked LEDs using a memory-artificial neural network in 3D indoor visible light positioning systems,” Opt. Express 29(21), 33937–33953 (2021). [CrossRef]  

15. Q. Liang, Y. Sun, C. Liu, M. Liu, and L. Wang, “Ledmapper: Toward efficient and accurate LED mapping for visible light positioning at scale,” IEEE Trans. Instrum. Meas. 71, 1–12 (2022). [CrossRef]  

16. J. He and B. Zhou, “Vehicle positioning scheme based on visible light communication using a CMOS camera,” Opt. Express 29(17), 27278–27290 (2021). [CrossRef]  

17. I. Sharp and K. Yu, “Indoor TOA error measurement, modeling, and analysis,” IEEE Trans. Instrum. Meas. 63(9), 2129–2144 (2014). [CrossRef]  

18. P. Du, S. Zhang, C. Chen, A. Alphones, and W.-D. Zhong, “Demonstration of a low-complexity indoor visible light positioning system using an enhanced TDOA scheme,” IEEE Photonics J. 10(4), 1–10 (2018). [CrossRef]  

19. C.-Y. Hong, Y.-C. Wu, Y. Liu, C.-W. Chow, C.-H. Yeh, K.-L. Hsu, D.-C. Lin, X.-L. Liao, K.-H. Lin, and Y.-Y. Chen, “Angle-of-arrival (AOA) visible light positioning (VLP) system using solar cells with third-order regression and ridge regression algorithms,” IEEE Photonics J. 12(3), 1–5 (2020). [CrossRef]  

20. Y. Yu, B. Zhu, Z. Zhang, L. Wang, L. Wu, and J. Dang, “Indoor visible light localization algorithm with the optimal optical angle-of-arrival estimator,” Proc. 2nd Information Communication Technologies Conference (ICTC), (2021), pp. 194–198.

21. L. Bai, Y. Yang, C. Feng, and C. Guo, “Received signal strength assisted perspective-three-point algorithm for indoor visible light positioning,” Opt. Express 28(19), 28045–28059 (2020). [CrossRef]  

22. W. Xu, J. Wang, H. Shen, H. Zhang, and X. You, “Indoor positioning for multiphotodiode device using visible-light communications,” IEEE Photonics J. 8(1), 1–11 (2016). [CrossRef]  

23. M. F. Keskin and S. Gezici, “Comparative theoretical analysis of distance estimation in visible light positioning systems,” J. Lightwave Technol. 34(3), 854–865 (2016). [CrossRef]  

24. W. Gu, M. Aminikashani, P. Deng, and M. Kavehrad, “Impact of multipath reflections on the performance of indoor visible light positioning systems,” J. Lightwave Technol. 34(10), 2578–2587 (2016). [CrossRef]  

25. B. Lin, X. Tang, Z. Ghassemlooy, C. Lin, and Y. Li, “Experimental demonstration of an indoor VLC positioning system based on OFDMA,” IEEE Photonics J. 9, 1–9 (2017). [CrossRef]  

26. N. Huang, C. Gong, J. Luo, and Z. Xu, “Design and demonstration of robust visible light positioning based on received signal strength,” J. Lightwave Technol. 38(20), 5695–5707 (2020). [CrossRef]  

27. E.-M. Jeong, S.-H. Yang, H.-S. Kim, and S.-K. Han, “Tilted receiver angle error compensated indoor positioning system based on visible light communication,” Electron. Lett. 49(14), 890–892 (2013). [CrossRef]  

28. H. Steendam, T. Q. Wang, and J. Armstrong, “Theoretical lower bound for indoor visible light positioning using received signal strength measurements and an aperture-based receiver,” J. Lightwave Technol. 35(2), 309–319 (2017). [CrossRef]  

29. C. Carre no, F. Seguel, P. Charpentier, and N. Krommenacker, “Polynomial fitting functions for visible light positioning based on RSS with tilted receiver,” 11th International Conference on Indoor Positioning and Indoor Navigation, IPIN 2021, (2021).

30. S. Li, S. Shen, and H. Steendam, “A positioning algorithm for VLP in the presence of orientation uncertainty,” Sig. Process. 160, 13–20 (2019). [CrossRef]  

31. J. N. Franklin, Matrix theory (Courier Corporation, 2012).

32. D. P. Palomar and M. Chiang, “A tutorial on decomposition methods for network utility maximization,” IEEE J. Select. Areas Commun. 24(8), 1439–1451 (2006). [CrossRef]  

33. M. Pincus, “A closed form solution of certain programming problems,” Oper. Res. 16(3), 690–694 (1968). [CrossRef]  

34. S. Kay and S. Saha, “Mean likelihood frequency estimation,” IEEE Trans. Signal Process. 48(7), 1937–1946 (2000). [CrossRef]  

35. H. Wang, S. Kay, and S. Saha, “An importance sampling maximum likelihood direction of arrival estimator,” IEEE Trans. Signal Process. 56(10), 5082–5092 (2008). [CrossRef]  

36. G. Wang and H. Chen, “An importance sampling method for TDOA-based source localization,” IEEE Trans. Wirel. Commun. 10(5), 1560–1568 (2011). [CrossRef]  

37. R. Wicklin, Simulating data with SAS (SAS Institute, 2013).

38. H. V. Poor, An introduction to signal detection and estimation (Springer Science & Business Media, 1998).

39. D. Kim, J. K. Park, and J. T. Kim, “Three-dimensional VLC positioning system model and method considering receiver tilt,” IEEE Access 7, 132205–132216 (2019). [CrossRef]  

40. L.-S. Hsu, D.-C. Tsai, H. M. Chen, Y.-H. Chang, Y. Liu, C.-W. Chow, S.-H. Song, and C.-H. Yeh, “Using Received-Signal-Strength (RSS) Pre-Processing and Convolutional Neural Network (CNN) to Enhance Position Accuracy in Visible Light Positioning (VLP),” Optical Fiber Communication Conference, (Optica Publishing Group, 2022), pp. W3I–6.

41. Y.-C. Chuang, Z.-Q. Li, C.-W. Hsu, Y. Liu, and C.-W. Chow, “Visible light communication and positioning using positioning cells and machine learning algorithms,” Opt. Express 27(11), 16377–16383 (2019). [CrossRef]  

42. M. F. Keskin, S. Gezici, and O. Arikan, “Direct and two-step positioning in visible light systems,” IEEE Trans. Commun. 66(1), 239–254 (2018). [CrossRef]  

43. D. Plets, S. Bastiaens, M. Ijaz, Y. Almadani, L. Martens, W. Raes, N. Stevens, and W. Joseph, “Three-dimensional visible light positioning: An experimental assessment of the importance of the LEDs’ locations,” 2019 International Conference on Indoor Positioning and Indoor Navigation (IPIN), (IEEE, 2019), pp. 1–6.

Data Availability

Data underlying the results presented in this paper are not publicly available at this time but may be obtained from the authors upon reasonable request.

Cited By

Optica participates in Crossref's Cited-By Linking service. Citing articles from Optica Publishing Group journals and other participating publishers are listed here.

Alert me when this article is cited.


Figures (11)

Fig. 1.
Fig. 1. Illustration of the considered indoor VLP system. (a) System configuration. (b) Signal flow.
Fig. 2.
Fig. 2. CDF of positioning errors with BFS and IS methods under different tilting angles.
Fig. 3.
Fig. 3. Simulated positioning error versus PD vertical tilting angle $\phi$.
Fig. 4.
Fig. 4. Simulated positioning error versus the unknown scalar $\alpha$ under $\phi = 0^\circ$.
Fig. 5.
Fig. 5. Estimated positions of 100 test points under $\phi =0^\circ$. (a) BFS. (b) IS.
Fig. 6.
Fig. 6. The experimental environment.
Fig. 7.
Fig. 7. The horizontal positions of LEDs and test receiver points.
Fig. 8.
Fig. 8. Mean positioning errors of the proposed scheme under different $I$.
Fig. 9.
Fig. 9. CDF of positioning errors in Cases 1 and 2.
Fig. 10.
Fig. 10. Mean positioning errors under different transceiver height assumptions in Cases 1 and 2.
Fig. 11.
Fig. 11. Illustration of the LED layout. (a) All LEDs are collinear. (b) LEDs are non-collinear.

Tables (5)

Tables Icon

Algorithm 1. The proposed VLP scheme with IS

Tables Icon

Table 1. The Experimental Parameters for Indoor VLP System

Tables Icon

Table 2. Estimated Values of LED Parameters under different $I$

Tables Icon

Table 3. Difference between the normalized RSS in Case 1 and fitting values using the estimated LED parameters with $I = 13$

Tables Icon

Table 4. Mean positioning errors in the simulation and experiment under experimental parameters

Equations (26)

Equations on this page are rendered with MathJax. Learn more.

$$p_{r,l} = B H_{r,l} p_l + n_{r,l},$$
$$H_{r,l} = \frac{A_r (m_l + 1)}{2 \pi \|\boldsymbol{c}_l - \boldsymbol{c}_r\|_2^2} \cos^{m_l}{(\theta_l)} \cos{(\varphi_l)} {\rm rect}\Big(\frac{\varphi_{l}}{\Phi}\Big),$$
$$p_{r,l} = \frac{B A_r (m_l + 1) p_l}{2 \pi} \cdot \frac{z^{m_l} (\boldsymbol{c}_l - \boldsymbol{c}_r)^\top \boldsymbol{u}_r}{\|\boldsymbol{c}_l - \boldsymbol{c}_r\|_2^{m_l + 3}} + n_{r,l}.$$
$$p_{r,l} = \frac{g_l z^{k_l} (\boldsymbol{c}_l - \boldsymbol{c}_r)^\top \boldsymbol{\mu}_r}{\|\boldsymbol{c}_l - \boldsymbol{c}_r\|_2^{k_l}} + n_{r,l},$$
$$d_{r,l} \approx z \left(\frac{BA_r g_l}{2\pi z^2 p_{r,l}} \right)^{\frac{1}{k_l}}, \quad l = 1,\ldots, L.$$
$$\begin{aligned} (x_1 - x_r)^2 + (y_1 - y_r)^2 + z^2 &\approx z^2 \left(\frac{BA_r g_1}{2\pi z^2 p_{r,1}} \right)^{\frac{2}{k_1}},\\ (x_2 - x_r)^2 + (y_2 - y_r)^2 + z^2 &\approx z^2 \left(\frac{BA_r g_2}{2\pi z^2 p_{r,2}} \right)^{\frac{2}{k_2}},\\ &\vdots\\ (x_L - x_r)^2 + (y_L - y_r)^2 + z^2 &\approx z^2 \left(\frac{BA_r g_L}{2\pi z^2 p_{r,L}} \right)^{\frac{2}{k_L}}. \end{aligned}$$
$$\boldsymbol{A} \boldsymbol{c}_{rh} = \boldsymbol{b},$$
$$\begin{aligned} &\boldsymbol{A} = \begin{bmatrix} 2(x_1 - x_2), 2(y_1 - y_2) \\ \cdots \\ 2(x_1 - x_L), 2(y_1 - y_L) \end{bmatrix}, \boldsymbol{c}_{rh} = \begin{bmatrix} x_r \\ y_r \end{bmatrix},\\ &\boldsymbol{b} = \begin{bmatrix} v_{r,2}^2 - v_{r,1}^2 + x_1^2 + y_1^2 - x_2^2 - y_2^2 \\ v_{r,3}^2 - v_{r,1}^2 + x_1^2 + y_1^2 - x_3^2 - y_3^2 \\ \cdots \\ v_{r,L}^2 - v_{r,1}^2 + x_1^2 + y_1^2 - x_L^2 - y_L^2 \end{bmatrix}. \end{aligned}$$
$$\begin{aligned} (\hat{x}_r, \hat{y}_r, \hat{\boldsymbol{\mu}}_r) & = \arg\max_{x_r, y_r, \boldsymbol{\mu}_r} \prod_{l=1}^{L} f (p_{r,l}|x_r, y_r, \boldsymbol{\mu}_r)\\ & = \arg\max_{x_r, y_r, \boldsymbol{\mu}_r} \sum_{l=1}^{L} \ln f (p_{r,l}|x_r, y_r, \boldsymbol{\mu}_r), \end{aligned}$$
$$f(p_{r,l}|x_r, y_r, \boldsymbol{\mu}_r) \propto \exp \left\{-\frac{(p_{r, l} - \boldsymbol{s}_l \boldsymbol{\mu}_r)^2}{2 \sigma_n^2} \right\},$$
$$(\hat{x}_r, \hat{y}_r, \hat{\boldsymbol{\mu}}_r) = \arg\min_{x_r, y_r, \boldsymbol{\mu}_r} \| \boldsymbol{p}_r - \boldsymbol{S} \boldsymbol{\mu}_r\|_2^2.$$
$$\begin{aligned} (\hat{x}_r, \hat{y}_r, \hat{\boldsymbol{\mu}}_r) &= \arg\min_{x_r, y_r, \boldsymbol{\mu}_r} \| \boldsymbol{p}_r - \boldsymbol{S} \boldsymbol{\mu}_r\|_2^2\\ &= \arg\min_{x_r, y_r}\min_{\boldsymbol{\mu}_r} \| \boldsymbol{p}_r - \boldsymbol{S} \boldsymbol{\mu}_r\|_2^2. \end{aligned}$$
$$\boldsymbol{\mu}_r^* = \arg\min_{\boldsymbol{\mu}_r} \| \boldsymbol{p}_r - \boldsymbol{S} \boldsymbol{\mu}_r\|_2^2.$$
$$\begin{aligned} (\hat{x}_r, \hat{y}_r) &= \arg\min_{x_r, y_r} \| \boldsymbol{p}_r - \boldsymbol{S} \boldsymbol{\mu}_r^*\|_2^2\\ &= \arg\min_{x_r, y_r} \| \boldsymbol{p}_r - \boldsymbol{S} (\boldsymbol{S}^\top \boldsymbol{S})^{{-}1} \boldsymbol{S}^\top \boldsymbol{p}_r \|_2^2\\ &= \arg\max_{x_r, y_r} \| \boldsymbol{S} (\boldsymbol{S}^\top \boldsymbol{S})^{{-}1} \boldsymbol{S}^\top \boldsymbol{p}_r \|_2^2. \end{aligned}$$
$$\| \boldsymbol{S} (\boldsymbol{S}^\top \boldsymbol{S})^{{-}1} \boldsymbol{S}^\top \boldsymbol{p}_r \|_2^2 = \| \boldsymbol{S} (\boldsymbol{S}^{{-}1} \boldsymbol{S}^{-\top}) \boldsymbol{S}^\top \boldsymbol{p}_r \|_2^2 = \| \boldsymbol{p}_r\|_2^2,$$
$$\hat{c}_i = \lim_{\lambda \rightarrow \infty} \frac{\int \cdots \int_S c_i \exp[\lambda F(\boldsymbol{c})] {\text{d}}\boldsymbol{c}}{\int \cdots \int_S \exp[\lambda F(\boldsymbol{c})] {\text{d}}\boldsymbol{c}}, \quad i=1, \ldots, n.$$
$$f (\boldsymbol{c}) = \frac{\exp[\lambda F(\boldsymbol{c})]}{\int \cdots \int_S \exp[\lambda F(\boldsymbol{c})] {\text{d}}\boldsymbol{c}},$$
$$\begin{aligned} F (\boldsymbol{c}_{rh}) &= \| \boldsymbol{S} (\boldsymbol{S}^\top \boldsymbol{S})^{{-}1} \boldsymbol{S}^\top \boldsymbol{p}_r \|_2^2\\ &= \boldsymbol{p}_r^\top \boldsymbol{S} (\boldsymbol{S}^\top \boldsymbol{S})^{{-}1} \boldsymbol{S}^\top \boldsymbol{p}_r. \end{aligned}$$
$$\mathbb{E}_{f} \{ \boldsymbol{c}_{rh} \} \approx \frac{\sum_{t=1}^T w(\boldsymbol{c}_{rh}^{(t)}) \boldsymbol{c}_{rh}^{(t)}}{\sum_{t=1}^T w(\boldsymbol{c}_{rh}^{(t)})},$$
$$\boldsymbol{p}_r^\top \boldsymbol{S} (\boldsymbol{S}^\top \boldsymbol{S})^{{-}1} \boldsymbol{S}^\top \boldsymbol{p}_r \stackrel{(a)}{\rightarrow} \boldsymbol{p}_r^\top \boldsymbol{S} \boldsymbol{S}^\top \boldsymbol{p}_r \stackrel{(b)}{\rightarrow} \boldsymbol{p}_r^\top \widetilde{\boldsymbol{S}} \widetilde{\boldsymbol{S}}^\top \boldsymbol{p}_r,$$
$$\begin{aligned} \widetilde{\boldsymbol{S}} = \begin{bmatrix} x_1 - x_r & y_1 - y_r & z \\ \vdots & \vdots & \vdots \\ x_L - x_r & y_L - y_r & z \end{bmatrix} \triangleq [\widetilde{\boldsymbol{s}}_{x_r}, \widetilde{\boldsymbol{s}}_{y_r}, \widetilde{\boldsymbol{s}}_z]. \end{aligned}$$
$$\begin{aligned}q(\boldsymbol{c}_{rh}) &\propto \exp[\lambda_1 \boldsymbol{p}_r^\top \widetilde{\boldsymbol{S}} \widetilde{\boldsymbol{S}}^\top \boldsymbol{p}_r]\\ &\propto \exp[ \lambda_1 |\widetilde{\boldsymbol{s}}_{x_r}^\top \boldsymbol{p}_r|^2] \cdot \exp[ \lambda_1 |\widetilde{\boldsymbol{s}}_{y_r}^\top \boldsymbol{p}_r|^2]\\ &= \widetilde{q}_x(x_r) \cdot \widetilde{q}_y(y_r), \end{aligned}$$
$$\hat{x}_r \approx \frac{\sum_{t=1}^T w (\boldsymbol{c}_{rh}^{(t)}) x_r^{(t)}}{\sum_{t=1}^T w (\boldsymbol{c}_{rh}^{(t)})}, \quad \hat{y}_r \approx \frac{\sum_{t=1}^T w (\boldsymbol{c}_{rh}^{(t)}) y_r^{(t)}}{\sum_{t=1}^T w (\boldsymbol{c}_{rh}^{(t)})},$$
$$w (\boldsymbol{c}_{rh}^{(t)}) = \frac{\exp[ \lambda {F} (\boldsymbol{c}_{rh}^{(t)})]}{\widetilde{q}_x (x_r^{(t)}) \cdot \widetilde{q}_y (y_r^{(t)})} \propto \frac{f (\boldsymbol{c}_{rh}^{(t)})}{q(\boldsymbol{c}_{rh}^{(t)})}.$$
$$\boldsymbol{S} = \begin{bmatrix} g_1 \frac{z^{k_1}}{d_{r,1}^{k_1}} (x_1 - x_r) & g_1 \frac{z^{k_1}}{d_{r,1}^{k_1}} (y_1 - y_r) & g_1 \frac{z^{k_1}}{d_{r,1}^{k_1}} \cdot z\\ \vdots & \vdots & \vdots \\ g_L \frac{z^{k_L}}{d_{r,L}^{k_L}} (x_L - x_r) & g_L \frac{z^{k_L}}{d_{r,L}^{k_L}} (y_L - y_r) & g_L \frac{z^{k_L}}{d_{r,L}^{k_L}} \cdot z \end{bmatrix} = \begin{bmatrix} g_1 \frac{z^{k_1}}{d_{r,1}^{k_1}} & & \\ & \ddots & \\ & & g_L \frac{z^{k_L}}{d_{r,L}^{k_L}} \end{bmatrix} \cdot \widetilde{\boldsymbol{S}},$$
$$ \widetilde{\boldsymbol{S}}=\left[\begin{array}{ccc} x_1-x_r & y_1-y_r & z \\ x_2-x_r & y_2-y_r & z \\ \vdots & \vdots & \vdots \\ x_L-x_r & y_L-y_r & z \end{array}\right] \rightarrow\left[\begin{array}{cc|c} x_1-x_r & y_1-y_r & z \\ \hline x_2-x_1 & y_2-y_1 & 0 \\ \vdots & \vdots & \vdots \\ x_L-x_1 & y_L-y_1 & 0 \end{array}\right] \triangleq \widetilde{\boldsymbol{S}^{\prime}}=\left[\begin{array}{c|c} \widetilde{\boldsymbol{S}_1^{\prime}} & z \\ \hline \widetilde{\widetilde{S}_2^{\prime}} & \mathbf{0} \end{array}\right] . $$
Select as filters


Select Topics Cancel
© Copyright 2024 | Optica Publishing Group. All rights reserved, including rights for text and data mining and training of artificial technologies or similar technologies.