Skip to main content

An Optimal Control Approach to Mapping GPS-Denied Environments Using a Stochastic Robotic Swarm

  • Chapter
  • First Online:
Robotics Research

Part of the book series: Springer Proceedings in Advanced Robotics ((SPAR,volume 2))

Abstract

This paper presents an approach to mapping a region of interest using observations from a robotic swarm without localization. The robots have local sensing capabilities and no communication, and they exhibit stochasticity in their motion. We model the swarm population dynamics with a set of advection-diffusion-reaction partial differential equations (PDEs). The map of the environment is incorporated into this model using a spatially-dependent indicator function that marks the presence or absence of the region of interest throughout the domain. To estimate this indicator function, we define it as the solution of an optimization problem in which we minimize an objective functional that is based on temporal robot data. The optimization is performed numerically offline using a standard gradient descent algorithm. Simulations show that our approach can produce fairly accurate estimates of the positions and geometries of different types of regions in an unknown environment.

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

Access this chapter

Chapter
USD 29.95
Price excludes VAT (USA)
  • Available as PDF
  • Read on any device
  • Instant download
  • Own it forever
eBook
USD 169.00
Price excludes VAT (USA)
  • Available as EPUB and PDF
  • Read on any device
  • Instant download
  • Own it forever
Softcover Book
USD 219.99
Price excludes VAT (USA)
  • Compact, lightweight edition
  • Dispatched in 3 to 5 business days
  • Free shipping worldwide - see info
Hardcover Book
USD 219.99
Price excludes VAT (USA)
  • Durable hardcover edition
  • Dispatched in 3 to 5 business days
  • Free shipping worldwide - see info

Tax calculation will be finalised at checkout

Purchases are for personal use only

Institutional subscriptions

References

  1. Ammari, H.: An Introduction to Mathematics of Emerging Biomedical Imaging, vol. 62. Springer, Berlin (2008)

    MATH  Google Scholar 

  2. Belmiloudi, A.: Stabilization, Optimal and Robust Control: Theory and Applications in Biological and Physical Sciences. Springer, London (2008)

    MATH  Google Scholar 

  3. Biswas, R., Limketkai, B., Sanner, S., Thrun, S.: Towards object mapping in dynamic environments with mobile robots. In: International Conference on Intelligent Robots and Systems (IROS) (2002)

    Google Scholar 

  4. Borzì, A., Schulz, V.: Computational optimization of systems governed by partial differential equations. In: Society for Industrial and Applied Mathematics (SIAM), Philadelphia (2012)

    Google Scholar 

  5. Correll, N., Hamann, H.: Probabilistic modeling of swarming systems. In: J Kacprzyk WP (ed) Springer Handbook of Computational Intelligence, pp 1423–1431. Springer, Heidelberg (2015)

    Google Scholar 

  6. Dirafzoon, A., Lobaton, E.: Topological mapping of unknown environments using an unlocalized robotic swarm. In: International Conference on Intelligent Robots and Systems (IROS) (2013)

    Google Scholar 

  7. Dirafzoon, A., Betthauser, J., Schornick, J., Benavides, D., Lobaton, E.: Mapping of unknown environments using minimal sensing from a stochastic swarm. In: International Conference on Intelligent Robots and Systems (IROS) (2014)

    Google Scholar 

  8. Elamvazhuthi, K.: A variational approach to planning, allocation and mapping in robot swarms using infinite dimensional models. Master’s thesis, Arizona State University (2014)

    Google Scholar 

  9. Elamvazhuthi, K., Berman, S.: Optimal control of stochastic coverage strategies for robotic swarms. In: International Conference on Robotics and Automation (ICRA) (2015)

    Google Scholar 

  10. Fattorini, H.O.: Infinite Dimensional Optimization and Control Theory, vol 54. Cambridge University Press, Cambridge (1999)

    Google Scholar 

  11. Gardiner, C.W.: Stochastic Methods: A Handbook for the Natural and Social Sciences, 4th edn. Springer, Evanston, IL, USA (2009)

    MATH  Google Scholar 

  12. Hinze, M., Pinnau, R., Ulbrich, M., Ulbrich, S.: Optimization with PDE Constraints, vol. 23. Springer, Netherlands (2009)

    Book  MATH  Google Scholar 

  13. Horning, M., Lin, M., Siddarth, S., Zou, S., Haberland, M., Yin, K., Bertozzi, A.L.: Compressed sensing environmental mapping by an autonomous robot. In: Proceedings of the Second International Workshop on Robotic Sensor Networks. Seattle, WA (2015)

    Google Scholar 

  14. Kirsch, A.: An Introduction to the Mathematical Theory of Inverse Problems, 2nd edn. vol 120. Springer, New York (2011)

    Google Scholar 

  15. Kuipers, B., Byun, Y.T.: A robot exploration and mapping strategy based on a semantic hierarchy of spatial representations. Int. J. Robot. Auton. Syst. 8(1–2), 47–63 (1991)

    Article  Google Scholar 

  16. Liu, S., Mohta, K., Shen, S., Kumar, V.: Towards collaborative mapping and exploration using multiple micro aerial robots. In: International Symposium on Experimental Robotics (ISER) (2014)

    Google Scholar 

  17. Robb, R.A.: Biomedical Imaging, Visualization, and Analysis. Wiley Inc, New York, NY, USA (1999)

    Google Scholar 

  18. Robertson, P., Angermann, M., Krach, B.: Simultaneous localization and mapping for pedestrians using only foot-mounted inertial sensors. In: Proceedings of the 11th International Conference on Ubiquitous Computing (Ubicomp) (2009)

    Google Scholar 

  19. Thrun, S.: A probabilistic online mapping algorithm for teams of mobile robots. Int. J. Robot. Res. 20(5), 335–363 (2001)

    Article  Google Scholar 

  20. Thrun, S., Bücken, A.: Integrating grid-based and topological maps for mobile robot navigation. In: Proceedings of the AAAI 13th National Conference on Artificial Intelligence (1996)

    Google Scholar 

  21. Tong, S., Fine, E.J., Lin, Y., Cradick, T.J., Bao, G.: Nanomedicine: tiny particles and machines give huge gains. Ann. Biomed. Eng. 42(2), 243–259 (2014)

    Article  Google Scholar 

  22. Tröltzsch, F.: Optimal Control of Partial Differential Equations: Theory, Methods, and Applications, vol. 112. American Mathematical Society, Providence, RI, USA (2010)

    MATH  Google Scholar 

  23. Tuchin, V.V.: Tissue Optics, Light Scattering Methods and Instruments for Medical Diagnosis, 3rd edn. vol. PM254. Spie Press Book (2015)

    Google Scholar 

  24. Wilson, S., Pavlic, T.P., Kumar, G.P., Buffin, A., Pratt, S.C., Berman, S.: Design of ant-inspired stochastic control policies for collective transport by robotic swarms. Swarm Intell. 8(4), 303–327 (2014)

    Article  Google Scholar 

Download references

Acknowledgements

This work was supported by NSF Awards CMMI-1363499 and CMMI-1436960.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Ragesh K. Ramachandran .

Editor information

Editors and Affiliations

Appendices

Appendix 1: Mathematical Preliminaries

We study the solution to PDEs in the weak sense, which can be found in the Sobolev space \( H^1(\varOmega ) = \left\{ y \in L^2(\varOmega ) : \frac{\partial y}{\partial x_1} \in L^2(\varOmega ),\right. \left. \ \frac{\partial y}{\partial x_2} \in L^2(\varOmega ) \right\} \). Here, the spatial derivative is to be understood as a weak derivative defined in the distributional sense. The space is equipped with the common Sobolev space norm, \( \left\| y \right\| _{H^1(\varOmega )} = \sqrt{\left( \left\| y \right\| ^2_{L^2(\varOmega )} + \sum _{i =1}^{2} \left\| \frac{\partial y}{\partial x_i} \right\| ^2_{L^2(\varOmega )} \right) } \). We also define \( V = H^1(\varOmega ) \), which has the dual space \( V^* = H^1(\varOmega )^* \).

We consider the general system for Eqs. (3)–(5):

$$\begin{aligned} \frac{\partial u}{\partial t} = Au + \sum _{i=1}^{2}v_iB_iu -K(\mathbf {x})u + f\ ~~~in\ ~L, \nonumber \\ \mathbf {n} \cdot (D\nabla u - \mathbf {v} u) = g\ ~~~on\ ~\varGamma , \nonumber \\ u(\mathbf {x},0) = u_0,\qquad \end{aligned}$$
(9)

where A is a formal operator and \(B_i \) is an operator defined as \(B_i : L^2(0,T;V) \rightarrow L^2(0,T;L^2(\varOmega )) \), \( K(\mathbf {x}) \in L^2(\varOmega )\), \( f \in F = L^2(0,T;L^2(\varOmega )) \) is the forcing function in the system, \( g \in G = L^2(0,T;L^2(\partial \varOmega )) \), and \( u_0 \in L^2(\varOmega ) \). The variational form of the operator A, called \( A_g \), is defined as \( A_g : L^2(0,T;V) \rightarrow L^2(0,T;V^*) \). The solution of the system in the weak sense is given by \( u \in U = L^2(0,T;V) \) with \( u_t \in U^* = L^2(0,T;V^*) \) if it satisfies the equation:

$$\begin{aligned} \left\langle \frac{\partial u}{\partial t},\phi \right\rangle _{U^*,U} = \left\langle A_g,\phi \right\rangle _{U^*,U} + \sum _{i=1}^{2} \left\langle v_iB_iu, \phi \right\rangle _F -\left\langle K(\mathbf {x})u, \phi \right\rangle _F + \left\langle f,\phi \right\rangle _ F \end{aligned}$$
(10)

for all \( \phi \in L^2(0,T;V) \). The boundary conditions are equipped with \( A_g \) in the variational formulation using Green’s theorem. This is essentially the variational form of the Laplacian,

$$\begin{aligned} \left\langle A_g u,\phi \right\rangle _{U^*,U} = -\left\langle D \nabla u, \nabla \phi \right\rangle _{L^2(\varOmega )} + \int _{\partial \varOmega } \left( g + \mathbf {n} \cdot \mathbf {v} u \right) \phi dx. \end{aligned}$$
(11)

In the macroscopic model Eqs. (3)–(5), we define \( A = \nabla ^2\), \( B_i = \frac{\partial }{\partial x_i} \), \( f = 0\), and \(g = 0\).

Appendix 2: Adjoint Equations

The adjoint equation \(\nabla _\mathbf {u} \mathscr {L} = 0\) implies that \([\nabla _{u_{1}} \mathscr {L},\ldots ,\nabla _{u_{i}} \mathscr {L},\ldots ,\nabla _{u_{N}} \mathscr {L} ] = 0.\) From Eq. (8),

$$\begin{aligned} \nabla _{u_{i}}\mathscr {L}= & {} \nabla _{u_{i}}\mathbf {J}(\mathbf {u},K) + \nabla _{u_{i}} \sum _{j = 1}^{N} \langle p_j, \varPsi _j(u_j,K) \rangle \nonumber \\= & {} \nabla _{u_{i}}\mathbf {J}(u_{i},K) + \nabla _{u_{i}} \langle p_i, \varPsi _i(u_i,K) \rangle , \end{aligned}$$
(12)

since a term in the sum is a function of \(u_i\) only when \( i=j \). By Eq. (7),

$$\begin{aligned} \nabla _{u_{i}}\mathbf {J}(u_{i},K) = \nabla _{u_{i}} \sum _{j = 1}^{N} W_j J_j(u_j) = W_i \nabla _{u_{i}} J_i(u_i). \end{aligned}$$
(13)

From Eq. (6),

$$\begin{aligned} \nabla _{u_{i}} J_i(u_i) = \nabla _{u_{i}} \left( \frac{1}{2}\left\| (D u_i)(t) - g_i(t)\right\| _{L^2([0,T])} ^2\right) , \end{aligned}$$
(14)

where \(D := U \rightarrow L^2([0,T])\) and \((D u_i)(t) = \int _{\varOmega } u_i(\mathbf {x},t) d\mathbf {x}\). Then, by the chain rule of differentiation [4, 12], the directional derivative of \(J_i(u_i)\), \( \nabla _{u_{i}} J_i(u_i) \), is given by

$$\begin{aligned} \langle \nabla _{u_{i}} J_i(u_i) , s \rangle _{U} = \langle (D u_i)(t) - g_i(t), Ds \rangle _{L^2([0,T])} = \langle D^*((D u_i)(t) - g_i(t)), s \rangle _{U}. \end{aligned}$$
(15)

Here, \( D^* := L^2([0,T]) \rightarrow U \) and \( (D^*f)(t) = f(t)\cdot \mathbf {1}_\varOmega (\mathbf {x})\), where \(f(t) \in L^2([0,T])\) and \(\mathbf {1}_\varOmega \) is the indicator function of \(\varOmega \subset \mathbb {R}^2\). We can show that \( \langle D y,f \rangle = \langle y,D^*f \rangle \ \forall y \in U , ~f \in L^2([0,T]) \). Therefore,

$$\begin{aligned} \nabla _{u_{i}} J_i(u_i) = D^*((D u_i)(t) - g_i(t)). \end{aligned}$$
(16)

By definition,

$$\begin{aligned} \langle p_i,\nabla _{u_{i}} \varPsi _i(u_i,K)s \rangle = \langle \nabla _{u_{i}} \varPsi _i(u_i,K)^* p_i,s \rangle ~~ \forall s \in U, \end{aligned}$$
(17)

where \( \nabla _{u_{i}} \varPsi _i(u_i,K)^* \) is the adjoint operator of \( \nabla _{u_{i}} \varPsi _i(u_i,K) \) corresponding to the inner product of the Hilbert space. Now, by taking the directional derivative of \(\varPsi _i(u_i,K) \) at \( u_i \) in the direction of s, we obtain

$$\begin{aligned} \nabla _{u_{i}} \varPsi _i(u_i,K)s = \frac{\partial s}{\partial t} -( \nabla \cdot (D\nabla s - \mathbf {v}_i(t)s) -kK(\mathbf {x})s). \end{aligned}$$
(18)

Substituting Eq. (18) into Eq. (17) yields

$$\begin{aligned} \langle p_i,\nabla _{u_{i}} \varPsi _i(u_i,K)s \rangle = \int _{0}^{T}\langle p_i,\frac{\partial s}{\partial t} \rangle _{L^2(\varOmega )} - \langle p_i,D\nabla ^2s \rangle + \langle p_i,\nabla \cdot \mathbf {v}_i(t)s \rangle + \langle p_i,kK(\mathbf {x})s \rangle . \end{aligned}$$

Using integration by parts on the integral term in the equation above, we get

$$\begin{aligned} \int _{0}^{T}\langle p_i,\frac{\partial s}{\partial t} \rangle _{L^2(\varOmega )} = \langle p_i(T),s(T)\rangle - \langle p_i(0),s(0)\rangle -\int _{0}^{T}\langle s,\frac{\partial p_i}{\partial t} \rangle _{L^2(\varOmega )}. \end{aligned}$$

As this is true for all \( s \in U\), we could choose the s with \( s(0) = 0\) and construct \( p_i(T) \) such that \( \int _{0}^{T}\langle p_i,\frac{\partial s}{\partial t} \rangle _{L^2(\varOmega )} = \int _{0}^{T}\langle -\frac{\partial p_i}{\partial t},s \rangle _{L^2(\varOmega )} \). Thus, we choose the final condition of the adjoint equation as \( p_i(T) =0\). We now make use of the following lemma:

Lemma 1 Let L and \( L^* \) be operators defined by \( L : L^2(0,T;V) \rightarrow L^2(0,T;V^*)\) and \( L^* : L^2(0,T;V) \rightarrow L^2(0,T;V^*)\), respectively. The variational form of L is:

$$\begin{aligned} \langle Lu,\phi \rangle _{V^*,V} = -\left\langle D \nabla u, \nabla \phi \right\rangle _{L^2(\varOmega )} - \langle \mathbf {v} \cdot \nabla u,\phi \rangle _{L^2(\varOmega )} + \int _{\partial \varOmega } \mathbf {n} \cdot (\mathbf {v}u \phi ) dx \end{aligned}$$

\( \forall \phi \in V \). Also, by Lagrange’s identity, \( \langle Lu,p \rangle _{V^*,V} = \langle u,L^*p \rangle _{V,V^*} ~\forall u,p \in L^2 (0,T;V) \). We use the zero-flux boundary condition in Eq. (4) to compute the variational form of the operator \( L^* \) to be \(\langle L^*p,\phi \rangle _{V^*,V} = -\left\langle D \nabla p, \nabla \phi \right\rangle _{L^2(\varOmega )} + \langle \mathbf {v}\cdot \nabla p,\phi \rangle _{L^2(\varOmega )}\) \( \forall p \in L^2(0,T;V)\) and \(\forall \phi \in V \).

Using the variational form of the Laplacian as in Eq. (11) and applying Lemma 1 and integration by parts, we can show that \( -\langle p_i,D\nabla ^2s \rangle + \langle p_i,\nabla \cdot \mathbf {v}_i(t)s \rangle \) can be transformed into \(~ -\langle D\nabla ^2p_i,s \rangle - \langle \nabla \cdot \mathbf {v}_i(t)p_i,s \rangle \) with the boundary condition \( \mathbf {n}\cdot \nabla p_i = 0\). Finally, we observe that \( \langle p_i,K(\mathbf {x})s \rangle = \langle p_i K(\mathbf {x}),s \rangle \). By combining these results with Eqs. (12), (15), and (17), we obtain

$$\begin{aligned} \langle \nabla _{u_{i}} J_i(u_i) , s \rangle + \langle -\frac{\partial p_i}{\partial t} - D\nabla ^2p_i -\nabla \cdot \mathbf {v}_i(t)p_i + p_i kK(\mathbf {x}), s \rangle = 0. \end{aligned}$$

Thus, the set of adjoint equations for the system defined by the \(i^{th}\) set of constraints, \(\varPsi _i(u_i,K)\), with respect to the objective functional, \( \mathbf {J} \), is given by

$$\begin{aligned} -\frac{\partial p_i}{\partial t} = \nabla \cdot (D\nabla p_i +\mathbf {v}_i(t)p_i) - p_i k K(\mathbf {x}) - \nabla _{u_{i}} J_i(u_i) ~~\ in ~\ L \end{aligned}$$
(19)

with the Neumann boundary conditions

$$\begin{aligned} \mathbf {n}\cdot \nabla p_i = 0 ~~\ on\ ~\varGamma , ~~~ p_i(T) = 0, ~~~i = 1,\ldots ,N. \end{aligned}$$
(20)

Here, Eq. (19) with Eq. (20) has a solution in the weak sense.

Appendix 3: Gradient Equation

Using a similar analysis to the one in Appendix 2, we find that \( \nabla _K \mathscr {L} \) reduces to

$$\begin{aligned} \nabla _K \mathscr {L} = \nabla _{K}\mathbf {J}(\mathbf {u},K) + \sum _{i = 1}^{N} \nabla _{K} \langle p_i, \varPsi _i(u_i,K) \rangle . \end{aligned}$$
(21)

From Eq. (7), we can derive the following expressions:

$$\begin{aligned} \nabla _{K}\mathbf {J}(\mathbf {u},K) = \nabla _{K} \frac{\lambda }{2}\Vert K(\mathbf {x})\Vert ^2_{L^2(\varOmega )}, ~~~ \langle \nabla _{K}\mathbf {J}(\mathbf {u},K),s \rangle = \langle \lambda K(\mathbf {x}),s \rangle . \end{aligned}$$
(22)

As in Appendix 2, we could express \( \langle p_i, \nabla _{K} \varPsi _i(u_i,K) s \rangle \) as \( \langle \nabla _{K} \varPsi _i(u_i,K)^* p_i,s \rangle \ \forall s \in L^2(\varOmega )\), where \( \nabla _{K} \varPsi _i(u_i,K)^* \) is the adjoint operator of \( \nabla _{K} \varPsi _i(u_i,K) \) corresponding to the inner product of the Hilbert space. Now, by taking the directional derivative of \(\varPsi _i(u_i,K) \) at K in the direction of s, we find that \( \nabla _{K} \varPsi _i(u_i,K)s = ku_is \). Therefore, with further simplification, we can show that

$$\begin{aligned} \langle \nabla _{K} \varPsi _i(u_i,K)^* p_i,s \rangle = \langle (\varXi (ku_ip_i))(\mathbf {x}),s\rangle _{L^2(\varOmega )}, \end{aligned}$$
(23)

where \( \varXi := L^2(0,T;\varOmega ) \rightarrow L^2(\varOmega ) \) and \( (\varXi f)(\mathbf {x}) = \int _{0}^{T}fdt\) for all \(f \in L^2([0,T];\varOmega ) \) and \( \mathbf {x} \in \varOmega \). By combining Eqs. (21)–(23), we formulate the objective functional derivative as

$$\begin{aligned} \mathbf {J}' = \sum _{i = 1}^{N} (\varXi (ku_ip_i))(\mathbf {x}) + \lambda K(\mathbf {x}). \end{aligned}$$
(24)

Thus, the computation of \( \mathbf {J}' \) requires \(u_i\) and \( p_i \), which can be obtained by solving \(\varPsi _i(u_i,K)\) forward and solving Eqs. (19), (20) backward.

Rights and permissions

Reprints and permissions

Copyright information

© 2018 Springer International Publishing AG

About this chapter

Cite this chapter

Ramachandran, R.K., Elamvazhuthi, K., Berman, S. (2018). An Optimal Control Approach to Mapping GPS-Denied Environments Using a Stochastic Robotic Swarm. In: Bicchi, A., Burgard, W. (eds) Robotics Research. Springer Proceedings in Advanced Robotics, vol 2. Springer, Cham. https://doi.org/10.1007/978-3-319-51532-8_29

Download citation

  • DOI: https://doi.org/10.1007/978-3-319-51532-8_29

  • Published:

  • Publisher Name: Springer, Cham

  • Print ISBN: 978-3-319-51531-1

  • Online ISBN: 978-3-319-51532-8

  • eBook Packages: EngineeringEngineering (R0)

Publish with us

Policies and ethics