Submitted Successfully!
To reward your contribution, here is a gift for you: A free trial for our video production service.
Thank you for your contribution! You can also upload a video entry or images related to this topic.
Version Summary Created by Modification Content Size Created at Operation
1 -- 1823 2023-07-27 13:48:12 |
2 format correct Meta information modification 1823 2023-07-28 02:59:16 |

Video Upload Options

Do you have a full video?

Confirm

Are you sure to Delete?
Cite
If you have any further questions, please contact Encyclopedia Editorial Office.
Beck, E.; Shin, B.; Wang, S.; Wiedemann, T.; Shutin, D.; Dekorsy, A. Swarm Exploration and Communications. Encyclopedia. Available online: https://encyclopedia.pub/entry/47357 (accessed on 04 July 2024).
Beck E, Shin B, Wang S, Wiedemann T, Shutin D, Dekorsy A. Swarm Exploration and Communications. Encyclopedia. Available at: https://encyclopedia.pub/entry/47357. Accessed July 04, 2024.
Beck, Edgar, Ban-Sok Shin, Shengdi Wang, Thomas Wiedemann, Dmitriy Shutin, Armin Dekorsy. "Swarm Exploration and Communications" Encyclopedia, https://encyclopedia.pub/entry/47357 (accessed July 04, 2024).
Beck, E., Shin, B., Wang, S., Wiedemann, T., Shutin, D., & Dekorsy, A. (2023, July 27). Swarm Exploration and Communications. In Encyclopedia. https://encyclopedia.pub/entry/47357
Beck, Edgar, et al. "Swarm Exploration and Communications." Encyclopedia. Web. 27 July, 2023.
Swarm Exploration and Communications
Edit

Swarm exploration by multi-agent systems relies on stable inter-agent communication. However, so far both exploration and communication have been mainly considered separately despite their strong inter-dependency in such systems. By a semantic communication design, communication efficiency in terms of latency, required data rate, energy, and complexity may be improved. 

distributed exploration probabilistic factor graphs machine learning semantic communication goal-oriented communication

1. Introduction

In hazardous or inhospitable environments, exploration, and monitoring tasks impose high risks on human operators. Typical examples include emergency scenarios caused by nuclear or toxic accidents, as well as exploration scenarios in extraterrestrial environments [1][2]. Here, the use of mobile robotic systems is required. Cooperation in a multi-agent system, such as a swarm, is able to accelerate such reconnaissance missions or mapping tasks significantly [3]. An example of swarm exploration on an extraterrestrial surface, e.g., on Mars, is shown in Figure 1: Agents distribute and process sensed data along the arrows with the aim to reconstruct an unknown physical or chemical process 𝑢(𝜹,𝑡) of interest at position 𝜹 and time t or relevant parameters of such processes in the domain ΩΩ. For instance, a process of interest can be the spatio-temporal distribution of gas concentration. There, a relevant process parameter is the location of gas sources.
Figure 1. A swarm of autonomous agents explores an unknown physical process 𝑢(𝜹,𝑡) over spatial coordinate 𝜹 and time t in the spatial domain ΩΩ.
To achieve this goal, swarm exploration incorporates methods for distributed sensing, optimized (intelligent) information gathering [4], and agent movement/action coordination (exploitation). In particular, it requires the communication of locally and instantaneously available exploration measurements between agents. The underlying communication network acts as a data exchange backbone and is the tool that eventually enables the “diffusion” of local information to all agents and, hence, assists global decision-making. Communication is therefore always an integral part of a swarm exploration.
Swarm exploration often considers reliable and error-free communications, i.e., ideal links. However, communication systems do add uncertainty to the exchanged information. This means that studies so far paint an optimistic picture of the exploration performance metric. For instance, communication uncertainty needs to be considered when predicting new sampling positions for agents, since locations causing severe communication degradation will be useless for distributed information processing/exploration purposes.

2. Distributed Multi-Agent Exploration

Distributed exploration requires cooperative computational techniques, which are also referred to as “in-network processing” [5]. The estimation is done such that each node conducts “local” computations and shares intermediate results with its neighboring nodes. The key to these computations is a decomposition of a network-global objective function into a sum of “local” sub-objectives, typically with additional constraints that ensure a network-wide convergence to a specific solution. A special class of such algorithms is called consensus-based algorithms, see, e.g., [6][7][8][9][10][11]. This class of algorithms enforces consensus over the whole network, i.e., each node converges to the same solution. Here, the Alternating Directions Method of Multipliers (ADMM) [12] has gained popularity for in-network processing due to its ability to handle different types of constraints on model parameters.
As an alternative, diffusion-based approaches (see e.g., [13][14][15] and references therein) have been proposed that estimate a quantity in a distributed fashion within a network without enforcing consensus. Such approaches are also based on solving an optimization problem that permits a decomposability of the network objective function. One of the applications of interest for swarm exploration is seismic imaging of subsurface structures. In particular, distributed subsurface imaging techniques based on the full waveform inversion and the traveltime tomography have been proposed recently that can be directly applied to decentralized multi-agent networks, s. [15][16]. Full waveform inversion is a high-resolution geophysical imaging method based on the wave equation [17]. For a distributed implementation of this method, a global cost function is decomposed over the receivers and local gradients and subsurface images are computed. Following the diffusion-based information exchange, these gradients, and images are exchanged among the receivers in order to obtain a global estimate of the subsurface image.
For the exploration of complex physical processes that are described in terms of Partial Differential Equations (PDEs), classical approaches typically do not provide a direct assessment of statistical information about the quality of estimated parameters. In contrast, Bayesian inference methods postulate randomness of the parameters of interest and are from the domain of machine learning [18]. As such, instead of a point estimate, parameter distributions are computed. FGs can be used to describe probabilistic relationships between all model parameters [19] and parameter estimation is then realized using message passing schemes [20]. Bayesian tools have been used in the past for inverse PDE problems (see, e.g., [3][21]). In [3], the authors use FGs for inverse PDE modeling in a distributed setting and to localize gas sources based on concentration measurement samples. In essence, random variables are used to represent the gas concentration distribution in each mesh cell of the discretized PDE. An FG is then applied to capture temporal and spatial dependencies between concentration variables.
Having inferred the model parameters, one can then design a movement planning strategy that exploits the statistics of the estimated model parameters to optimally guide agents to new, more informative sampling locations to accelerate the exploration process. The work of [22] proposes information-driven approaches that guide agents based on mutual information or entropy. Furthermore, some swarm exploration approaches make use of (deep) Reinforcement Learning (RL) for the movement strategy of the agents [23][24][25]. However, the success of these methods relies heavily on the availability of suitable training data to learn an adequate movement strategy. Especially in applications with scarce training data, such approaches are likely to fail or perform unreliably in real environments: The use of synthetic training data introduces a model mismatch that is learned by the system. Furthermore, the learned behavior cannot be easily corrected a-posteriori due to the structure of the Deep Neural Network (DNN) that cannot be interpreted.
All aforementioned methods for distributed exploration and path planning heavily rely on agent-to-agent communication of the exchanged data or messages. Hence, the quality of the inter-agent communication links has a direct impact on the exploration result. However, the majority of state-of-the-art methods for distributed exploration do not sufficiently take into account the erroneous nature of the communication links. Most studies consider erroneous inter-agent links by integrating noise and link failures into the link model, see, e.g., [26][27]. The algorithmic solutions are then adapted to these erroneous communication links.

3. Machine Learning for Communications

The probabilistic view often used in exploration is vital for the field of communications. Since Claude Elwood Shannon laid the theoretical foundation of communications and information theory [28], probabilistic models have found their way not only into exploration but also into one prominent example of recent research interest: Artificial Intelligence (AI), in particular its subdomain Machine Learning (ML).
In the last decade, ML saw the emergence of powerful (probabilistic) models known as Deep Neural Networks (DNNs). Thanks to its ability to approximate arbitrarily well and to learn abstract features, it has led to several breakthroughs in research areas where there is no explicit domain knowledge but data to be collected, e.g., pattern recognition, generative modeling, and RL [29]. Previously considered intractable to optimize, automatic differentiation on dedicated Graphics Processing Units (GPUs) and innovative architectures now enable data-driven training of DNNs.
The impressive results showing equal or superhuman performance have not gone unnoticed by the communications community. Thus, much of the recent literature focuses on the data-driven design of the physical layer with DNNs, e.g., for wireless, molecular, and fiber-optical channels [29]. One prominent early example of such an approach is the Auto Encoder (AE) where a complete communication system is interpreted as one DNN and trained end-to-end [30].
In wireless communications, a number of channel models have been proposed and are widely used, so that key gains from using ML are expected in approximating optimal algorithmic structures that are otherwise numerically too complex (algorithm deficit) to be realized. For example, the computational complexity of Maximum A-Posteriori (MAP) decoding of large block-length codes or MAP detection, e.g., in massive Multiple Input Multiple Output (MIMO) systems, grows exponentially with code/system dimensions. In fact, e.g., using plain DNNs for decoding enables lowering of decoding complexity while approximately maintaining MAP error rate [31]. To improve generalization and reduce training complexity, more recent works focus on the idea of deep unfolding [32][33][34]. In deep unfolding, the parameters of a model-based iterative algorithm with a fixed number of iterations are untied and enriched with additional weights as well as non-linearities. The resulting DNN can be optimized for performance improvements in MIMO detection [33][35] and belief propagation decoding [33]. An example of an algorithm deficit on a higher level beyond the physical layer is resource allocation, where it is difficult to analytically express the true objective function or to find the global optimum. Thus, Deep RL has proven to be a proper means [36].

Semantic Communication

In contrast to wireless channels, a model deficit holds for molecular and fiber-optical channels. Note that it applies in particular to the example of this article: integration of semantic context, here exploration, into communication system design. The idea of semantic communication emerged in the early 1950s [37][38][39] but has seen a lot of research interest only recently with the rise of ML application to the physical layer [40][41][42][43][44].
Its notion traces back to Weaver [37] who reviewed Shannon’s information theory [28] in 1949 and amended considerations w.r.t. semantic content of messages. Oftentimes quoted is his statement that “there seem to be [communication] problems at three levels[37]:
  • How accurately can the symbols of communication be transmitted? (The technical problem).
  • How precisely do the transmitted symbols convey the desired meaning? (The semantic problem).
  • How effectively does the received meaning affect conduct in the desired way? (The effectiveness problem).
Weaver saw the broad applicability of Shannon’s theory back in 1949 and argued for the generality of the theory at Level A for all levels [45].
The generic model of Weaver was revisited by Bao, Basu et al. in [39][46] where the authors define semantic information sources and semantic channels. In [39], the authors consider joint semantic compression and channel coding at Level B with the classic transmission system, i.e., Level A, as the (semantic) channel. By this means, the authors can derive semantic counterparts of the source and channel coding theorems.
Recently, drawing inspiration from Weaver, Bao, Basu et al. [37][39][46] and enabled by the rise of ML in communications research, DNN-based natural language processing techniques, i.e., transformer networks, were introduced in AEs for the task of text and speech transmission [47][48][49][50]. The aim of these techniques is to learn compressed hidden representations of the semantic content of sentences to improve communication efficiency, but exact recovery of the source (text) is the main objective. This leads to performance improvements in semantic metrics, especially at low Signal-to-Noise Ratio (SNR) compared to classical digital transmissions.
As a result, semantic communication is still a nascent field: It remains still unclear what this term exactly means and especially its distinction from Joint Source-Channel Coding (JSCC) [44][48][51]. As a result, many survey papers aim to provide an interpretation, see, e.g., [40][41][42][43][44].

References

  1. Burgués, J.; Marco, S. Environmental chemical sensing using small drones: A review. Sci. Total Environ. 2020, 748, 141172.
  2. Tzoumas, G.; Pitonakova, L.; Salinas, L.; Scales, C.; Richardson, T.; Hauert, S. Wildfire detection in large-scale environments using force-based control for swarms of UAVs. Swarm Intell. 2022, 17, 89–115.
  3. Wiedemann, T.; Manss, C.; Shutin, D. Multi-agent exploration of spatial dynamical processes under sparsity constraints. Auton. Agents Multi-Agent Syst. 2018, 32, 134–162.
  4. Viseras, A. Distributed Multi-Robot Exploration under Complex Constraints. Ph.D. Thesis, Universidad Pablo de Olavide, Seville, Spain, 2018.
  5. Schizas, I.D.; Mateos, G.; Giannakis, G.B. Distributed LMS for consensus-based in-network adaptive processing. IEEE Trans. Signal Process. 2009, 57, 2365–2382.
  6. Kar, S.; Moura, J.M. Distributed consensus algorithms in sensor networks with imperfect communication: Link failures and channel noise. IEEE Trans. Signal Process. 2008, 57, 355–369.
  7. Pereira, S.S. Distributed Consensus Algorithms for Wireless Sensor Networks: Convergence Analysis and Optimization. Ph.D. Thesis, Universitat Politècnica de Catalunya-Barcelona Tech, Barcelona, Spain, 2012.
  8. Talebi, S.P.; Werner, S. Distributed Kalman Filtering and Control Through Embedded Average Consensus Information Fusion. IEEE Trans. Autom. Control 2019, 64, 4396–4403.
  9. Wang, S.; Shin, B.S.; Shutin, D.; Dekorsy, A. Diffusion Field Estimation Using Decentralized Kernel Kalman Filter with Parameter Learning over Hierarchical Sensor Networks. In Proceedings of the IEEE MLSP, Espoo, Finland, 21–24 September 2020.
  10. Shin, B.S.; Shutin, D. Distributed blind deconvolution of seismic signals under sparsity constraints in sensor networks. In Proceedings of the IEEE International Workshop on Machine Learning for Signal Processing, MLSP, Espoo, Finland, 21–24 September 2020.
  11. Shutin, D.; Shin, B.S. Variational Bayesian Learning for Decentralized Blind Deconvolution of Seismic Signals Over Sensor Networks. IEEE Access 2021, 9, 164316–164330.
  12. Boyd, S.; Parikh, N.; Chu, E. Distributed Optimization and Statistical Learning via the Alternating Direction Method of Multipliers; Now Publishers Inc.: Delft, The Netherlands, 2011.
  13. Sayed, A.H. Adaptation, learning, and optimization over networks. Found. Trends Mach. Learn. 2014, 7, 311–801.
  14. Shin, B.S.; Yukawa, M.; Cavalcante, R.L.G.; Dekorsy, A. Distributed adaptive learning with multiple kernels in diffusion networks. IEEE Trans. Signal Process. 2018, 66, 5505–5519.
  15. Shin, B.S.; Shutin, D. Adapt-then-combine full waveform inversion for distributed subsurface imaging in seismic networks. In Proceedings of the ICASSP, IEEE International Conference on Acoustics, Speech and Signal Processing—Proceedings, 2021, Toronto, ON, Canada, 6–11 June 2021; pp. 4700–4704.
  16. Shin, B.S.; Shutin, D. Distributed Traveltime Tomography Using Kernel-based Regression in Seismic Networks. IEEE Geosci. Remote Sens. Lett. 2022, 19, 1–5.
  17. Fichtner, A. Full Seismic Waveform Modelling and Inversion; Springer: Berlin/Heidelberg, Germany, 2011.
  18. Bishop, C.M. Pattern Recognition and Machine Learning; Springer: Berlin/Heidelberg, Germany, 2006.
  19. Loeliger, H.A. An introduction to factor graphs. IEEE Signal Process. Mag. 2004, 21, 28–41.
  20. Kschischang, F.R.; Frey, B.J.; Loeliger, H.A. Factor graphs and the sum-product algorithm. IEEE Trans. Inf. Theory 2001, 47, 498–519.
  21. Wang, J.; Zabaras, N. Hierarchical Bayesian models for inverse problems in heat conduction. Inverse Probl. 2004, 21, 183.
  22. Julian, B.J.; Angermann, M.; Schwager, M.; Rus, D. Distributed robotic sensor networks: An information-theoretic approach. Int. J. Robot. Res. 2012, 31, 1134–1154.
  23. Hüttenrauch, M.; Šošić, A.; Neumann, G. Deep Reinforcement Learning for Swarm Systems. J. Mach. Learn. Res. 2019, 20, 1–31.
  24. Zhu, X.; Zhang, F.; Li, H. Swarm Deep Reinforcement Learning for Robotic Manipulation. Procedia Comput. Sci. 2022, 198, 472–479.
  25. Kakish, Z.; Elamvazhuthi, K.; Berman, S. Using Reinforcement Learning to Herd a Robotic Swarm to a Target Distribution. In Distributed Autonomous Robotic Systems, Proceedings of the 15th International Symposium, Brussels, Belgium, 19–24 September 2022; Springer International Publishing: Berlin, Germany, 2022; pp. 401–414.
  26. Schizas, I.D.; Ribeiro, A.; Giannakis, G.B. Consensus in ad hoc WSNs with noisy links—Part I: Distributed estimation of deterministic signals. IEEE Trans. Signal Process. 2008, 56, 350–364.
  27. Zhao, X.; Tu, S.; Sayed, A. Diffusion Adaptation over Networks Under Imperfect Information Exchange and Non-Stationary Data. IEEE Trans. Signal Process. 2012, 60, 3460–3475.
  28. Shannon, C.E. A Mathematical Theory of Communication. Bell Syst. Tech. J. 1948, 27, 379–423.
  29. Simeone, O. A Very Brief Introduction to Machine Learning with Applications to Communication Systems. IEEE Trans. Cogn. Commun. Netw. 2018, 4, 648–664.
  30. O’Shea, T.; Hoydis, J. An Introduction to Deep Learning for the Physical Layer. IEEE Trans. Cogn. Commun. Netw. 2017, 3, 563–575.
  31. Gruber, T.; Cammerer, S.; Hoydis, J.; Brink, S.t. On deep learning-based channel decoding. In Proceedings of the 51st Annual Conference on Information Sciences and Systems (CISS 2017), Baltimore, MD, USA, 22–24 March 2017; pp. 1–6.
  32. Monga, V.; Li, Y.; Eldar, Y.C. Algorithm Unrolling: Interpretable, Efficient Deep Learning for Signal and Image Processing. IEEE Signal Process. Mag. 2021, 38, 18–44.
  33. Balatsoukas-Stimming, A.; Studer, C. Deep Unfolding for Communications Systems: A Survey and Some New Directions. In Proceedings of the IEEE International Workshop on Signal Processing Systems (SiPS 2019), Nanjing, China, 20–23 October 2019; pp. 266–271.
  34. Farsad, N.; Shlezinger, N.; Goldsmith, A.J.; Eldar, Y.C. Data-Driven Symbol Detection Via Model-Based Machine Learning. In Proceedings of the 2021 IEEE Statistical Signal Processing Workshop (SSP), Rio de Janeiro, Brazil, 11–14 July 2021; pp. 571–575.
  35. Beck, E.; Bockelmann, C.; Dekorsy, A. CMDNet: Learning a Probabilistic Relaxation of Discrete Variables for Soft Detection with Low Complexity. IEEE Trans. Commun. 2021, 69, 8214–8227.
  36. Gracla, S.; Beck, E.; Bockelmann, C.; Dekorsy, A. Robust Deep Reinforcement Learning Scheduling via Weight Anchoring. IEEE Commun. Lett. 2023, 27, 210–213.
  37. Weaver, W. Recent Contributions to the Mathematical Theory of Communication. In The Mathematical Theory of Communication; The University of Illinois Press: Champaign, IL, USA, 1949; Volume 10, pp. 261–281.
  38. Carnap, R.; Bar-Hillel, Y. An Outline of a Theory of Semantic Information; Research Laboratory of Electronics, Massachusetts Institute of Technology: Cambridge, MA, USA, 1952; p. 54.
  39. Bao, J.; Basu, P.; Dean, M.; Partridge, C.; Swami, A.; Leland, W.; Hendler, J.A. Towards a theory of semantic communication. In Proceedings of the 2011 IEEE Network Science Workshop, New York, NY, USA, 22–24 June 2011; pp. 110–117.
  40. Popovski, P.; Simeone, O.; Boccardi, F.; Gündüz, D.; Sahin, O. Semantic-Effectiveness Filtering and Control for Post-5G Wireless Connectivity. J. Indian Inst. Sci. 2020, 100, 435–443.
  41. Calvanese Strinati, E.; Barbarossa, S. 6G networks: Beyond Shannon towards semantic and goal-oriented communications. Comput. Netw. 2021, 190, 107930.
  42. Lan, Q.; Wen, D.; Zhang, Z.; Zeng, Q.; Chen, X.; Popovski, P.; Huang, K. What is Semantic Communication? A View on Conveying Meaning in the Era of Machine Intelligence. J. Commun. Inf. Netw. 2021, 6, 336–371.
  43. Uysal, E.; Kaya, O.; Ephremides, A.; Gross, J.; Codreanu, M.; Popovski, P.; Assaad, M.; Liva, G.; Munari, A.; Soret, B.; et al. Semantic Communications in Networked Systems: A Data Significance Perspective. IEEE/ACM Trans. Netw. 2022, 36, 233–240.
  44. Gündüz, D.; Qin, Z.; Aguerri, I.E.; Dhillon, H.S.; Yang, Z.; Yener, A.; Wong, K.K.; Chae, C.B. Beyond Transmitting Bits: Context, Semantics, and Task-Oriented Communications. IEEE J. Sel. Areas Commun. 2023, 41, 5–41.
  45. Beck, E.; Bockelmann, C.; Dekorsy, A. Semantic Information Recovery in Wireless Networks. arXiv 2023, arXiv:2204.13366.
  46. Basu, P.; Bao, J.; Dean, M.; Hendler, J. Preserving quality of information by using semantic relationships. Pervasive Mob. Comput. 2014, 11, 188–202.
  47. Xie, H.; Qin, Z.; Li, G.Y.; Juang, B.H. Deep Learning Enabled Semantic Communication Systems. IEEE Trans. Signal Process. 2021, 69, 2663–2675.
  48. Farsad, N.; Rao, M.; Goldsmith, A. Deep Learning for Joint Source-Channel Coding of Text. In Proceedings of the 2018 IEEE International Conference on Acoustics, Speech and Signal Processing (ICASSP), Calgary, AB, Canada, 15–20 April 2018; pp. 2326–2330.
  49. Weng, Z.; Qin, Z.; Li, G.Y. Semantic Communications for Speech Signals. In Proceedings of the 2021 IEEE International Conference on Communications (ICC), Montreal, QC, Canada, 14–18 June 2021; pp. 1–6.
  50. Sana, M.; Strinati, E.C. Learning Semantics: An Opportunity for Effective 6G Communications. In Proceedings of the 2022 IEEE 19th Annual Consumer Communications Networking Conference (CCNC), Las Vegas, NV, USA, 8–11 January 2022; pp. 631–636.
  51. Bourtsoulatze, E.; Kurka, D.B.; Gündüz, D. Deep Joint Source-Channel Coding for Wireless Image Transmission. IEEE Trans. Cogn. Commun. Netw. 2019, 5, 567–579.
More
Information
Contributors MDPI registered users' name will be linked to their SciProfiles pages. To register with us, please refer to https://encyclopedia.pub/register : , , , , ,
View Times: 232
Revisions: 2 times (View History)
Update Date: 28 Jul 2023
1000/1000
Video Production Service