Skip to main content
Log in

Optimality analysis of range sensor placement under constrained deployment region

  • Original Paper
  • Published:
Wireless Networks Aims and scope Submit manuscript

Abstract

Source localization is a critical issue in various wireless sensor network applications. However, communication and concealment constraints often restrict sensor placement, resulting in non-arbitrary sensor deployment regions. To further enhance localization accuracy, this paper presents an optimality analysis of range sensor placement under constrained deployment regions, focusing on optimal geometries rather than specific localization algorithms. The optimality analysis is formulated as a constrained optimisation problem that maximizes the determinant of the Fisher information matrix, also known as D-optimality, while taking into account the constraints imposed by the deployment region. To simplify the analysis, we introduce the concepts of maximum feasible angle and separation angle, which are used to express the objective function and constraints in equivalent forms. By comparing the maximum feasible angle with the optimal separation angles in unconstrained cases, our method will be applicable to both circular constrained regions and general irregular regions. The conclusions we have reached are comprehensive and intuitive, and they differ significantly from the conventional uniform angular geometry. The proposed range sensor-source geometries are verified through theoretical analysis and simulations.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Fig. 1
Fig. 2
Fig. 3
Fig. 4
Fig. 5
Fig. 6
Fig. 7
Fig. 8
Fig. 9
Fig. 10
Fig. 11
Fig. 12

Similar content being viewed by others

References

  1. Saeedi, I. D. I., & Al-Qurabat, A. K. M. (2022). Perceptually important points-based data aggregation method for wireless sensor networks. Baghdad Science Journal, 35, 0875.

    Article  Google Scholar 

  2. Lu, R., Liu, X., & Chen, X. (2023). Localization and tracking of multiple fast moving targets in bistatic MIMO radar. Signal Processing, 203, 108780.

    Article  Google Scholar 

  3. Al-Qurabat, A. K. M., & Idrees, A. K. (2019). Two level data aggregation protocol for prolonging lifetime of periodic sensor networks. Wireless Networks, 25(6), 3623–3641.

    Article  Google Scholar 

  4. Chan, Y. T., & Ho, K. C. (1994). A simple and efficient estimator for hyperbolic location. IEEE Transactions on signal processing, 42(8), 1905–1915.

    Article  Google Scholar 

  5. Huang, Y., Benesty, J., Elko, G. W., & Mersereati, R. M. (2001). Real-time passive source localization: A practical linear-correction least-squares approach. IEEE transactions on Speech and Audio Processing, 9(8), 943–956.

    Article  Google Scholar 

  6. Simonetto, A., & Leus, G. (2014). Distributed maximum likelihood sensor network localization. IEEE Transactions on Signal Processing, 62(6), 1424–1437.

    Article  MathSciNet  MATH  Google Scholar 

  7. Huang, Y., Bai, M., Li, Y., Zhang, Y., & Chambers, J. (2021). An improved variational adaptive Kalman filter for cooperative localization. IEEE Sensors Journal, 21(9), 10 775-10 786.

    Article  Google Scholar 

  8. Raja, G., Suresh, S., Anbalagan, S., Ganapathisubramaniyan, A., & Kumar, N. (2021). PFIN: An efficient particle filter-based indoor navigation framework for UAVs. IEEE Transactions on Vehicular Technology, 70(5), 4984–4992.

    Article  Google Scholar 

  9. Liu, J., & Guo, G. (2021). Vehicle localization during GPS outages with extended Kalman filter and deep learning. IEEE Transactions on Instrumentation and Measurement, 70, 1–10.

    Article  Google Scholar 

  10. Nguyen, N. H., & Doğançay, K. (2016). Optimal geometry analysis for multistatic TOA localization. IEEE Transactions on Signal Processing, 64(16), 4180–4193.

    Article  MathSciNet  MATH  Google Scholar 

  11. Heydari, A., Aghabozorgi, M., & Biguesh, M. (2020). Optimal sensor placement for source localization based on RSSD. Wireless Networks, 26(7), 5151–5162.

    Article  Google Scholar 

  12. Huang, B., Li, T., Anderson, B. D., & Yu, C. (2013). Performance limits in sensor localization. Automatica, 49(2), 503–509.

    Article  MathSciNet  MATH  Google Scholar 

  13. Meng, W., Xie, L., & Xiao, W. (2013). Optimality analysis of sensor-source geometries in heterogeneous sensor networks. IEEE Transactions on Wireless Communications, 12(4), 1958–1967.

    Article  Google Scholar 

  14. Bishop, A. N., Fidan, B., Doğançay, K., Anderson, B. D., & Pathirana, P. N. (2008). Exploiting geometry for improved hybrid AOA/TDOA-based localization. Signal Processing, 88(7), 1775–1791.

    Article  MATH  Google Scholar 

  15. Kim, Y. H., Kim, D. G., Han, J. W., Song, K. H., & Kim, H. N. (2017). Analysis of sensor-emitter geometry for emitter localisation using TDOA and FDOA measurements. IET Radar, Sonar and Navigation, 11(2), 341–349.

    Article  Google Scholar 

  16. Xu, S. (2020). Optimal sensor placement for target localization using hybrid RSS, AOA and TOA measurements. IEEE Communications Letters, 24(9), 1966–1970.

    Article  Google Scholar 

  17. Xu, S., & Doğançay, K. (2017). Optimal sensor placement for 3-D angle-of-arrival target localization. IEEE Transactions on Aerospace and Electronic Systems, 53(3), 1196–1211.

    Article  Google Scholar 

  18. Xu, B., Fei, Y., Wang, X., Tang, J., & Razzaqi, A. A. (2023). Optimal topology design of multi-target AUVs for 3D cooperative localization formation based on angle of arrival measurement. Ocean Engineering, 271, 113758.

    Article  Google Scholar 

  19. Zhao, S., Chen, B. M., & Lee, T. H. (2013). Optimal sensor placement for target localisation and tracking in 2D and 3D. International Journal of Control, 86(10), 1687–1704.

    Article  MathSciNet  MATH  Google Scholar 

  20. Fang, X., & Li, J. (2018). Frame theory for optimal sensor augmentation problem of AOA localization. IEEE Signal Processing Letters, 25(9), 1310–1314.

    Article  Google Scholar 

  21. Golihaghighi, N., & Bighesh, M. B. (2021). Using frame theory for optimal placement of new added anchors in location estimation wireless networks. IEEE Transactions on Aerospace and Electronic Systems, 58(3), 1858–1867.

    Article  Google Scholar 

  22. Meng, W., Xie, L., & Xiao, W. (2016). Optimal TDOA sensor-pair placement with uncertainty in source location. IEEE Transactions on Vehicular Technology, 65(11), 9260–9271.

    Article  Google Scholar 

  23. Fang, X., Yan, W., Zhang, F., & Li, J. (2015). Optimal sensor placement for range-based dynamic random localization. IEEE Geoscience and Remote Sensing Letters, 12(12), 2393–2397.

    Article  Google Scholar 

  24. Huang, B., Xie, L., & Yang, Z. (2014). TDOA-based source localization with distance-dependent noises. IEEE Transactions on Wireless Communications, 14(1), 468–480.

    Article  Google Scholar 

  25. Fang, X., Yan, W., & Chen, W. (2016). Sensor placement for underwater source localization with fixed distances. IEEE Geoscience and Remote Sensing Letters, 13(9), 1379–1383.

    Article  Google Scholar 

  26. Yang, Y., Zheng, J., Liu, H., Ho, K., Chen, Y., & Yang, Z. (2022). Optimal sensor placement for source tracking under synchronization offsets and sensor location errors with distance-dependent noises. Signal Processing, 193, 108399.

    Article  Google Scholar 

  27. Bo, X., Razzaqi, A. A., & Farid, G. (2019). A review on optimal placement of sensors for cooperative localization of AUVs. Journal of Sensors, 2019.

  28. Sadeghi, M., Behnia, F., & Amiri, R. (2021). Optimal geometry analysis for elliptic localization in multistatic radars with two transmitters. IEEE Transactions on Aerospace and Electronic Systems, 58(1), 697–703.

    Article  Google Scholar 

  29. Sadeghi, M., Behnia, F., & Amiri, R. (2020). Optimal sensor placement for 2-D range-only target localization in constrained sensor geometry. IEEE Transactions on Signal Processing, 68, 2316–2327.

    Article  MathSciNet  MATH  Google Scholar 

  30. Cheng, X., Shu, F., Li, Y., Zhuang, Z., Wu, D., & Wang, J. (2022). Optimal measurement of drone swarm in RSS-based passive localization with region constraints. IEEE Open Journal of Vehicular Technology, 4, 1–11.

    Article  Google Scholar 

  31. Tzoreff, E., & Weiss, A. J. (2017). Path design for best emitter location using two mobile sensors. IEEE Transactions on Signal Processing, 65(19), 5249–5261.

    Article  MathSciNet  MATH  Google Scholar 

  32. Li, Y., Qi, G., & Sheng, A. (2019). Optimal deployment of vehicles with circular formation for bearings-only multi-target localization. Automatica, 105, 347–355.

    Article  MathSciNet  MATH  Google Scholar 

  33. Yoo, K., & Chun, J. (2020). Analysis of optimal range sensor placement for tracking a moving target. IEEE Communications Letters, 24(8), 1700–1704.

    Article  Google Scholar 

  34. Zheng, Y., Liu, J., Sheng, M., Han, S., Shi, Y., & Valaee, S. (2020). Toward practical access point deployment for angle-of-arrival based localization. IEEE Transactions on Communications, 69(3), 2002–2014.

    Article  Google Scholar 

  35. Bishop, A. N., Fidan, B., Anderson, B. D., Doğançay, K., & Pathirana, P. N. (2010). Optimality analysis of sensor-target localization geometries. Automatica, 46(3), 479–492.

    Article  MathSciNet  MATH  Google Scholar 

  36. Doğançay, K., & Hmam, H. (2008). Optimal angular sensor separation for AOA localization. Signal Processing, 88(5), 1248–1260.

    Article  MATH  Google Scholar 

Download references

Acknowledgements

This work was supported in part by Key Research and Development Program of Shaanxi (Grant No. 2022GY-240).

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Xinpeng Fang.

Additional information

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Appendices

Appendix A

In the case of \(\varphi <\pi /3\), an equivalent problem to the optimality criterion (21) is

$$\begin{aligned} \begin{aligned}&\max \,f=\sigma ^{-4}\left[ \sin ^2 \theta _{13}+\sin ^2 \theta _{12}+\sin ^2 \left( \theta _{13}-\theta _{12} \right) \right] \\&\text {s.t.}\left\{ \begin{array}{l} \theta _{12},\theta _{13},\theta _{13}-\theta _{12}\in \left[ 0,2\varphi \right] \\ \varphi <\pi /3\\ \end{array} \right. \end{aligned} \end{aligned}$$
(24)

Taking derivative of objective function in (24) with respect to \(\theta _{13}\) and \(\theta _{12}\), respectively, yields

$$\begin{aligned}{} & {} \frac{\partial f}{\partial \theta _{13}}=2\sigma ^{-4}\sin \left( 2\theta _{13}-\theta _{12} \right) \cos \theta _{12} \\{} & {} \frac{\partial f}{\partial \theta _{12}}=2\sigma ^{-4}\cos \theta _{13} \sin \left( 2\theta _{12}-\theta _{13} \right) \end{aligned}$$

\(a)\,\varphi <\pi /4\)

In the case of \(\varphi <\pi /4\), we can easily obtain

$$\begin{aligned} \begin{aligned}&\theta _{12}\le 2\varphi<\pi /2\\&2\theta _{13}-\theta _{12}\le 2\theta _{13}\le 4\varphi <\pi \end{aligned} \end{aligned}$$

then, \(\cos \theta _{12}>0\), \(\sin \left( 2\theta _{13}-\theta _{12}\right) >0\). Obviously, \(\frac{\partial f}{\partial \theta _{13}}>0\).

According to \(\theta _{13}\le 2\varphi\), we know the objective function gets its maximum at \(\theta _{13}=2\varphi\). Then the optimal geometry is that any two sensors are located at the two tangent points, respectively. Therefore, we can respectively write

$$\begin{aligned} \begin{aligned} f&=\sigma ^{-4}\left[ \sin ^2 2\varphi +\sin ^2\theta _{12}+\sin ^2\left( 2\varphi -\theta _{12}\right) \right] \\ \frac{\partial f}{\partial \theta _{12}}&= 2\sigma ^{-4}\cos 2\varphi \sin \left( 2\theta _{12}-2\varphi \right) \end{aligned} \end{aligned}$$

Applying \(2\varphi <\pi /2\) yields \(\cos 2\varphi >0\). Then f achieves its minimum at \(\theta _{12}=\varphi\), and its maximum is obtained at \(\theta _{12}=0\) or \(2\varphi\), it is easy to verify that the maximum is the same, this indicates that the third sensor should be placed at any tangent point, as shown in Fig. 5(a).

The maximum for \(\varphi <\pi /4\) is \(f= 2\sigma ^{-4}\sin ^2 2\varphi\).

\(b)\,\pi /4<\varphi <\pi /3\)

In the case of \(\pi /4<\varphi <\pi /3\), we have \(\theta _{13}\le 2\varphi <2\pi /3\).

If \(\theta _{13}\le \pi /2\), then \(\cos \theta _{13}>0\), the objective function f achieves its minimum at \(\theta _{12}=\theta _{13}/2\), and gets the same maximum \(2\sigma ^{-4}\sin ^2\theta _{13}\) at \(\theta _{12}=0\) or \(\theta _{13}\), the maximum is \(f= 2\sigma ^{-4}\).

That is, when \(\theta _{13}\le \pi /2\), the optimality criterion gets the maximum at \(\theta _{13}=\pi /2\), \(\theta _{12}=0\) or \(\pi /2\), the maximum is \(2\sigma ^{-4}\).

If \(\pi /2<\theta _{13}<2\pi /3\), then \(\cos \theta _{13}<0\), the objective function f achieves its maximum at \(\theta _{12}=\theta _{13}/2\). Then

$$\begin{aligned} \frac{\partial f}{\partial \theta _{13}}=2\sigma ^{-4}\left[ \sin (3\theta _{13}/2) \cos (\theta _{13}/2)+\sin \theta _{13}\right] \end{aligned}$$

It follows from \(\pi /2<\theta _{13}<2\pi /3\) that

$$\begin{aligned} \sin (3\theta _{13}/2)>0,\cos (\theta _{13}/2)>0,\sin \theta _{13}>0 \end{aligned}$$

That is, when \(\pi /2<\theta _{13}<2\pi /3\), the optimality criterion gets the maximum at \(\theta _{13}=2\varphi\), \(\theta _{12}=\theta _{13}/2=\varphi\), the maximum is \(\sigma ^{-4}\left[ \sin ^2 2\varphi +2\sin ^2\varphi \right]\).

We have to compare these two maximums to get the maximum when \(\pi /4<\varphi <\pi /3\).

Denote \(g=\sigma ^{-4}\left[ \sin ^2 2\varphi +2\sin ^2\varphi \right] -2\sigma ^{-4}\), then

$$\begin{aligned} \frac{\partial g}{\partial \varphi }=\sigma ^{-4}\left[ \sin 4\varphi +2\sin 2\varphi \right] >0 \end{aligned}$$

and \(g>g(\pi /4)=0\).

The maximum for \(\pi /4<\varphi <\pi /3\) is

$$\begin{aligned} \sigma ^{-4}\left[ \sin ^2 2\varphi +2\sin ^2\varphi \right] \end{aligned}$$

The optimal geometry is that any two sensors are placed at two tangent points, respectively, the third sensor is placed at line SO, as shown in Fig. 5(b).

\(c)\,\varphi =\pi /4\)

In the case of \(\varphi =\pi /4\), we have

$$\begin{aligned} \begin{aligned}&2\sigma ^{-4}\sin ^2 2\varphi =2\sigma ^{-4}\\&\sigma ^{-4}\left[ \sin ^2 2\varphi +2\sin ^2\varphi \right] =2\sigma ^{-4} \end{aligned} \end{aligned}$$

Then the optimal geometries in Fig. 5(a) and (b) are equivalent.

Appendix B

As described in Sect. 5, the n sensors, \(s_1,s_2,\ldots ,s_n\), are divided into n/2 or \((n+1)/2\) groups according to whether n is even or odd.

$$\begin{aligned} {\mathcal {G}}_j = \left\{ {\begin{array}{ll} \left\{ {{s_{2j - 1}},{s_{2j}}} \right\} 1 \le j \le n/2 &\quad\text{n is even}\\ {\left. {\begin{array}{l} {\left\{ {{s_{2j - 1}},{s_{2j}}} \right\} }{1 \le j \le (n - 1)/2}\\ {\left\{ {{s_n}} \right\} }\qquad{j = (n + 1)/2} \end{array}} \right\} }&\quad\text{n is odd} \end{array}} \right. \end{aligned}$$

a) n is even

If every group of \({\mathcal {G}}_1,{\mathcal {G}}_2,\ldots ,{\mathcal {G}}_{n/2}\) is placed as \(C_2\) in Table 1, the separation angle \(\theta _{2j-1,2j}=\pi /2\), and \({\varvec{v}}_{2j-1}+{\varvec{v}}_{2j}={\varvec{0}}\) holds for every group.

As a result,

$$\begin{aligned} \sum _{i=1}^n{{\varvec{v}}_i}=\sum _{j=1}^{n/2}\sum _{i\in {\mathcal {G}}_j}{{\varvec{v}}_i}={\varvec{0}} \end{aligned}$$
(25)

it is shown in (25) that the geometry is optimal.

b) n is odd

We can conclude from Sect. 4.2 that \(\sum _{i=1}^3{{\varvec{v}}_i}\ne {\varvec{0}}\) for any three sensors. For the objective function (8), we can get the following inequality after some calculations [29].

$$\begin{aligned} \begin{aligned} \det \left( {\textbf{F}} \right)&\le \cos ^2\theta _n\sum _{i=1}^{n-1}{\sin ^2\theta _i}+\sin ^2\theta _n\sum _{i=1}^{n-1}{\cos ^2\theta _i}\\&\quad +\left( \sum _{i=1}^{n-1}{\cos ^2\theta _i} \right) \left( \sum _{i=1}^{n-1}{\sin ^2\theta _i} \right) \end{aligned} \end{aligned}$$
(26)

and the equality holds when \(\sum _{i=1}^{n-1}{\sin 2\theta _i}=0\).

It is worth noting that if any two sensors \(s_{2j-1}\) and \(s_{2j}\) are placed symmetrically about the x-axis, we have \(\theta _{2j-1}=-\theta _{2j}\), and \(\sin 2\theta _{2j-1}+\sin 2\theta _{2j}=0\), this means that if we place the two sensors in every group of \({\mathcal {G}}_1,{\mathcal {G}}_2,\ldots ,{\mathcal {G}}_{(n-1)/2}\) in this way, \(\sum _{i=1}^{n-1}{\sin 2\theta _i}=0\) can be guaranteed.

In this case, the modulus and azimuth angle of \(\varvec{v'}=\sum _{i=1}^{n-1}{{\varvec{v}}_i}\) are

$$\begin{aligned} \left\| \varvec{v'} \right\| \in {\left\{ \begin{array}{ll} \left[ 0,-\left( n-1 \right) \cos 2\varphi \right] &{} \angle \varvec{v'}=0\\ \left[ 0,n-1 \right] &{} \angle \varvec{v'}=\pi \\ \end{array}\right. } \end{aligned}$$
(27)

By the definition of \({\varvec{v}}_i\), we know that

$$\begin{aligned} \begin{aligned}&\left\| {\varvec{v}}_n \right\| =1\\&\angle {\varvec{v}}_n \in \left[ \pi -2\varphi ,\pi \right] \,\text {or}\,\left( -\pi , 2\varphi -\pi \right] \end{aligned} \end{aligned}$$
(28)

It can be proved that the minimum is achieved when \(\varvec{v'}\) and \({\varvec{v}}_n\) are in the opposite direction, i.e., \(\angle \varvec{v'}=0\), \(\angle {\varvec{v}}_n=\pi\), then \(s_n\) should be placed at line SO, the positions of the sensors are determined by comparing 1 with \(-\left( n-1 \right) \cos 2\varphi\).

Let \(-\left( n-1 \right) \cos 2\varphi =1\), then

$$\begin{aligned} \varphi = \frac{1}{2}\text{arc}\cos \left( -\frac{1}{n-1} \right) \end{aligned}$$

Hence, if \(\pi /4<\varphi \le \frac{1}{2}\text{arc}\cos \left( -\frac{1}{n-1} \right)\), we have \(-\left( n-1 \right) \cos 2\varphi <1\), that is

$$\begin{aligned} \left\| {\varvec{v}}_n+\varvec{v'}\right\| =\left\| {\varvec{v}}_n\right\| -\left\| \varvec{v'}\right\| \ne {\varvec{0}} \end{aligned}$$
(29)

To achieve the minimum, \(\left\| \varvec{v'}\right\|\) should be equal to \(-\left( n-1 \right) \cos 2\varphi\), then the two sensors in \({\mathcal {G}}_1,{\mathcal {G}}_2,\ldots ,{\mathcal {G}}_{(n-1)/2}\) should be placed as \(C_1\) in Table 1, and the minimum is

$$\begin{aligned} \left\| {\varvec{v}}_n+\varvec{v'}\right\| _{\min }=1+\left( n-1 \right) \cos 2\varphi \end{aligned}$$

If \(\frac{1}{2}\text{arc}\cos \left( -\frac{1}{n-1} \right)<\varphi <\pi /3\), then \(\varvec{v'}+{\varvec{v}}_n={\varvec{0}}\) can be achieved under certain conditions. We only give a sufficient but not necessary condition here. The two sensors in \({\mathcal {G}}_1,{\mathcal {G}}_2,\ldots ,{\mathcal {G}}_{(n-1)/2}\) can be placed with bearing angles \(\varphi _0=\pm \frac{1}{2}\text{arc}\cos \left( -\frac{1}{n-1} \right)\), respectively. Straightforwardly, we can easily obtain \(\left\| \varvec{v'} \right\| =1\), and

$$\begin{aligned} \left\| {\varvec{v}}_n+\varvec{v'}\right\| =\left\| {\varvec{v}}_n\right\| -\left\| \varvec{v'}\right\| = {\varvec{0}} \end{aligned}$$
(30)

It can be seen from (30) that the geometry is optimal, and the minimum is \(\left\| {\varvec{v}}_n+\varvec{v'}\right\| _{\min }=0\).

Appendix C

a) n is even

Similar to (26), we have the following inequality,

$$\begin{aligned} \det \left( {\textbf{F}} \right) \leqslant \sum _{i=1}^n{\cos ^2\theta _i}\sum _{i=1}^n{\sin ^2\theta _i} \end{aligned}$$
(31)

the equality holds when \(\sum _{i=1}^n{\sin 2\theta _i}=0\).

We can verify that if the two sensors in \({\mathcal {G}}_1,{\mathcal {G}}_2,\ldots ,{\mathcal {G}}_{n/2}\) are placed symmetrically about the x-axis, then \(\theta _{2j-1}=-\theta _{2j}\) and \(\sum _{i=1}^n{\sin 2\theta _i}=0\).

Since \(\varphi \le \pi /4\), the azimuth angle of \({\varvec{v}}_i\) is \(2\theta _i \le 2\varphi \le \pi /2\), and the modulus and azimuth angle of \(\varvec{v''}=\sum _{i=1}^n{{\varvec{v}}_i}\) are

$$\begin{aligned} \left\| \varvec{v''} \right\| \in \left[ n\cos 2\varphi ,n\right] ,\quad \angle \varvec{v''}= \pi \end{aligned}$$
(32)

We can conclude directly that the minimum can be achieved when the two sensors in every group are placed as \(C_1\) in Table 1. The minimum is \(\left\| \varvec{v''} \right\| _{\min } =n\cos 2\varphi\).

b) n is odd

Similarly, the two sensors in groups \({\mathcal {G}}_1,{\mathcal {G}}_2,\ldots ,{\mathcal {G}}_{(n-1)/2}\) should be placed symmetrically about the x-axis, and

$$\begin{aligned} \left\| \varvec{v'} \right\| \in \left[ (n-1)\cos 2\varphi ,n-1\right] ,\quad \angle \varvec{v'}= \pi \end{aligned}$$

Because \(\varvec{v'}\) is independent of \({\varvec{v}}_n\), then \(\left\| \varvec{v'} \right\|\) should be as small as possible, and its angle with \({\varvec{v}}_n\) should be as large as possible to achieve the minimum. At this time, \(\left\| \varvec{v'} \right\| =(n-1)\cos 2\varphi\), \(\angle {\varvec{v}}_n= \pm 2\varphi\). This indicates that the optimal geometry is the two sensors in \({\mathcal {G}}_1,{\mathcal {G}}_2,\ldots ,{\mathcal {G}}_{(n-1)/2}\) should be placed as \(C_1\) in Table 1, the sensor in \({\mathcal {G}}_{(n+1)/2}\) is placed at any tangent point. The minimum is

$$\begin{aligned} \left\| {\varvec{v}}_n+\varvec{v'} \right\| _{\min } =\sqrt{1+(n^2-1)\cos ^2 {2\varphi }} \end{aligned}$$

Rights and permissions

Springer Nature or its licensor (e.g. a society or other partner) holds exclusive rights to this article under a publishing agreement with the author(s) or other rightsholder(s); author self-archiving of the accepted manuscript version of this article is solely governed by the terms of such publishing agreement and applicable law.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Fang, X., He, Z. & Shi, R. Optimality analysis of range sensor placement under constrained deployment region. Wireless Netw 29, 2797–2812 (2023). https://doi.org/10.1007/s11276-023-03357-x

Download citation

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s11276-023-03357-x

Keywords

Navigation