Skip to main content

Advertisement

Log in

Time-optimal path planning in dynamic flows using level set equations: theory and schemes

  • Published:
Ocean Dynamics Aims and scope Submit manuscript

Abstract

We develop an accurate partial differential equation-based methodology that predicts the time-optimal paths of autonomous vehicles navigating in any continuous, strong, and dynamic ocean currents, obviating the need for heuristics. The goal is to predict a sequence of steering directions so that vehicles can best utilize or avoid currents to minimize their travel time. Inspired by the level set method, we derive and demonstrate that a modified level set equation governs the time-optimal path in any continuous flow. We show that our algorithm is computationally efficient and apply it to a number of experiments. First, we validate our approach through a simple benchmark application in a Rankine vortex flow for which an analytical solution is available. Next, we apply our methodology to more complex, simulated flow fields such as unsteady double-gyre flows driven by wind stress and flows behind a circular island. These examples show that time-optimal paths for multiple vehicles can be planned even in the presence of complex flows in domains with obstacles. Finally, we present and support through illustrations several remarks that describe specific features of our methodology.

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.

Institutional subscriptions

Fig. 1
Fig. 2
Fig. 3
Fig. 4
Fig. 5
Fig. 6
Fig. 7
Fig. 8
Fig. 9
Fig. 10
Fig. 11
Fig. 12
Fig. 13
Fig. 14
Fig. 15

Similar content being viewed by others

References

  • Adalsteinsson D, Sethian JA (1999) The fast construction of extension velocities in level set methods. J Comput Phys 148(1):2–22

    Article  Google Scholar 

  • Adalsteinsson D, Sethian JA (1995) A fast level set method for propagating interfaces. J Comput Phys 118(2):269–277

    Article  Google Scholar 

  • Alvarez A, Caiti A, Onken R (2004) Evolutionary path planning for autonomous underwater vehicles in a variable ocean. IEEE J Ocean Eng 29(2):418–429

    Article  Google Scholar 

  • Athans M, Falb PL (2006) Optimal control: an introduction to the theory and its applications. Dover Books on Engineering Series. Dover Publications

  • Bahr A, Leonard JJ, Fallon MF (2009) Cooperative localization for autonomous underwater vehicles. Int J Robot Res 28(6):714–728

    Article  Google Scholar 

  • Bakolas E, Tsiotras P (2010) The Zermelo-Voronoi diagram: a dynamic partition problem. Automatica 46(12):2059–2067

    Article  Google Scholar 

  • Bardi M, Capuzzo-Dolcetta I (2008) Optimal control and viscosity solutions of Hamilton-Jacobi-Bellman equations. Modern Birkhäuser Classics. Birkhäuser Boston

  • Barraquand J, Langlois B, Latombe JC (1992) Numerical potential field techniques for robot path planning. IEEE Trans Syst Man Cybern 22(2):224–241

    Article  Google Scholar 

  • Bhatta P, Fiorelli E, Lekien F, Leonard N E, Paley DA, Zhang F, Bachmayer R, Davis RE, Fratantoni DM, Sepulchre R (2005) Coordination of an underwater glider fleet for adaptive sampling. In: Proceedings of international workshop on underwater robotics, pp 61–69

  • Bianchini S, Tonon D (2012) SBV Regularity for Hamilton–Jacobi equations with Hamiltonian depending on (t,x). SIAM J Math Anal 44(3):2179–2203

    Article  Google Scholar 

  • Binney J, Krause A, Sukhatme GS (2010) Informative path planning for an autonomous underwater vehicle. In: Robotics and automation (icra), 2010 IEEE international conference on, pp 4791–4796

  • Bokanowski O, Forcadel N, Zidani H (2010) Reachability and minimal times for state constrained nonlinear problems without any controllability assumption. SIAM J Optim Control 48(7):4292–4316

    Article  Google Scholar 

  • Bressan A (2011) Viscosity solutions of Hamilton-Jacobi equations and optimal control problems. An Illustrated Tutorial, Penn State University

  • Bruce J, Veloso M (2002) Real-time randomized path planning for robot navigation. In: Proceedings of IROS-2002, Switzerland

  • Bryson AE, Ho YC (1975) Applied optimal control: optimization, estimation and control. CRC Press, New York

    Google Scholar 

  • Cannarsa P, Sinestrari C (2004) Semiconcave functions, Hamilton-Jacobi equations, and optimal control. Progress in Nonlinear Differential Equations and Their Applications. Birkhäuser Boston

  • Canny JF (1988) The complexity of robot motion planning. MIT Press, Cambridge, MA, USA

    Google Scholar 

  • Carroll KP, McClaran SR, Nelson EL, Barnett DM, Friesen DK, William GN (1992) AUV path planning: an A* approach to path planning with consideration of variable vehicle speeds and multiple, overlapping, time-dependent exclusion zones. In: Proceedings of the symposium on autonomous underwater vehicle technology, pp 79–84

  • Cheung MY, Hover FS (2013) A multi-armed bandit formulation with switching costs for autonomous adaptive acoustic relay positioning. In: International symposium on untethered unmanned submersible technology (UUST), Portsmouth

  • Cheung MY, Leighton J, Hover FS (2013) Autonomous mobile acoustic relay positioning as a multi-armed bandit with switching costs. In: International conference on intelligent robotic systems (IROS) (to appear), Tokyo

  • Choi H-L, How JP (2010) Continuous trajectory planning of mobile sensors for informative forecasting. Automatica 46(8):1266–1275

    Article  Google Scholar 

  • Chopp DL (August 2009) Another look at velocity extensions in the level set method. SIAM J Sci Comput 31(5):3255–3273

    Article  Google Scholar 

  • Chopp DL (1993) Computing minimal surfaces via level set curvature flow. J Comput Phys 106(1):77–91

    Article  Google Scholar 

  • Clarke FH, Ledyaev YS., Stern RJ, Wolenski PR (1998) Nonsmooth analysis and control theory. Springer, New York

    Google Scholar 

  • Cushman-Roisin B, Beckers JM (2010) Introduction to geophysical fluid dynamics. Physical and numerical aspects. Academic Press

  • Davis RE, Leonard NE, Fratantoni DM (2009) Routing strategies for underwater gliders. Deep Sea Res Part II: Top Stud Oceanogr 56(35):173–187

    Article  Google Scholar 

  • Dijkstra H, Katsman C (1997) Temporal variability of the wind-driven quasi-geostrophic double gyre ocean circulation: basic bifurcation diagrams. Geophys Astrophys Fluid Dyn 85:195–232

    Article  Google Scholar 

  • Fiorelli E, Leonard NE, Bhatta P, Paley DA, Bachmayer R, Fratantoni DM (2004) Multi-AUV control and adaptive sampling in Monterey Bay. IEEE J Ocean Eng 34:935–948

    Google Scholar 

  • Frankowska H (1989) Hamilton-Jacobi equations: viscosity Solutions and Generalized Gradients. J Math Anal Appl 141(1):21–26

    Article  Google Scholar 

  • Garau B, Bonet M, Alvarez A, Ruiz S, Pascual A (2009) Path planning for autonomous underwater vehicles in realistic oceanic current fields: application to gliders in the Western Mediterranean Sea. J Marit Res 6(2):5–22

    Google Scholar 

  • Garrido S, Moreno L, Blanco D (2006) Voronoi diagram and fast marching applied to path planning. In: Proceedings of IEEE international conference on Robotics and automation, pp 3049–3054

  • Haley PJ Jr., Lermusiaux PFJ, Robinson AR, Leslie WG, Logoutov O, Cossarini G, Liang XS, Moreno P, Ramp SR, Doyle J D, Bellingham J, Chavez F, Johnston S (2009) Forecasting and reanalysis in the Monterey Bay/California current region for the Autonomous Ocean Sampling Network-ii experiment. Deep Sea Res II: Top Stud in Oceanogr 56(35):127–148

    Article  Google Scholar 

  • Haley PJ Jr., Lermusiaux PFJ (2010) Multiscale two-way embedding schemes for free-surface primitive equations in the multidisciplinary simulation, estimation and assimilation system. Ocean Dyn 60(6):1497–1537

    Article  Google Scholar 

  • Heaney KD, Gawarkiewicz G, Duda TF, Lermusiaux PFJ (2007) Nonlinear optimization of autonomous undersea vehicle sampling strategies for oceanographic data-assimilation. J Field Robot 24(6):437–448

    Article  Google Scholar 

  • Hollinger G, Choudhary S, Qarabaqi P, Murphy C, Mitra U, Sukhatme G, Stojanovic M, Singh H, Hover F (2012) Underwater data collection using robotic sensor networks. IEEE J Sel Areas in Commun 30 (5):899–911

    Article  Google Scholar 

  • Isern-Gonzalez J, Hernandez-Sosa D, Fernasndez-Perdomo E, Cabrera-Gomez J, Dominguez-Brito AC, Prieto-Maran V (2012) Obstacle avoidance in underwater glider path planning. J Phys Agents 6 (1):11–20

    Google Scholar 

  • Kruger D, Stolkin R, Blum A, Briganti J (2007) Optimal AUV path planning for extended missions in complex, fast-flowing estuarine environments. In: IEEE international conference on Robotics and automation, pp 4265–4270

  • Kuffner JJ, LaValle SM (2000) RRT-connect: an efficient approach to single-query path planning Proceedings of IEEE international confernce on robotics and automation, pp 995–1001

  • Latombe JC (1991) Robot motion planning. Kluwer Academic, Boston

    Book  Google Scholar 

  • Lavalle SM (1998) Rapidly-exploring Random Trees: a new tool for path planning.Tech. rep., Iowa State University

  • LaValle S M (2006) Planning algorithms. Cambridge University, Cambridge

    Book  Google Scholar 

  • Leonard NE, Paley D, Lekien F, Sepulchre R, Fratantoni D, Davis R (2007) Collective motion, sensor networks, and ocean sampling. Proceedings of the IEEE, special issue on the emerging technology of networked control systems 95: 48–74

  • Leonard N, Fiorelli E (2001) Virtual leaders, artificial potentials and coordinated control of groups Proceedings of the 40th IEEE conference on Decision and control, vol 3, pp 2968–2973

  • Lermusiaux PFJ (1997) Error subspace data assimilation methods for ocean field estimation: theory, validation and applications. PhD thesis, Harvard University

  • Lermusiaux PFJ (2006) Uncertainty estimation and prediction for interdisciplinary ocean dynamics. J Comput Phys 217:176–199

    Article  Google Scholar 

  • Lermusiaux PFJ (2007) Adaptive modeling, adaptive data assimilation and adaptive sampling. Physica D Nonlinear Phenom 230:172–196

    Article  Google Scholar 

  • Lermusiaux PFJ, Chiu CS, Gawarkiewicz GG, Abbot P, Robinson AR, Miller RN, Haley PJ Jr., Leslie WG, Majumdar SJ, Pang A, Lekien F (2006) Quantifying uncertainties in ocean predictions. Oceanography 19(1):90–103

    Article  Google Scholar 

  • Lermusiaux PFJ, Lolla T, Haley PJ Jr., Yigit K, Ueckermann MP, Sondergaard T, Leslie WG (2014) Science of Autonomy: time-optimal path planning and adaptive sampling for swarms of ocean vehicles. Chapter 11, Springer Handbook of Ocean Engineering: Autonomous Ocean Vehicles, Subsystems and Control, Tom Curtin (Ed.) in press.

  • Lolla T, Haley PJ Jr., Lermusiaux PFJ (2014) Path planning in multiscale ocean flows: coordination, pattern formation and dynamic obstacles. Ocean Model. (to be submitted)

  • Lolla T, Haley PJ Jr., Lermusiaux PFJ (2014) Time-optimal path planning in dynamic flows using level set equations: realistic applications. Ocean Dyn. doi:10.1007/s10236-014-0760-3

  • Lolla T, Lermusiaux PFJ, Ueckermann MP, Haley PJ Jr. (2014c) Modified level set approaches for the planning of timeoptimal paths for swarms of ocean vehicles, MSEAS report-14. Department of Mechanical Engineering, Massachusetts Institute of Technology, Cambridge, MA, USA

  • Lolla T, Ueckermann MP, Yigit K, Haley PJ Jr., Lermusiaux PFJ (2012) Path planning in time dependent flow fields using level set methods. In Proceedings of IEEE international conference on robotics and automation, pp 166– 173

  • Lolla T (2012) Path planning in time dependent flows using level set methods. Masters thesis, Department of Mechanical Engineering, Massachusetts Institute of Technology

  • Melchior NA, Simmons R (2007) Particle RRT for path planning with uncertainty, pp 1617–1624

  • Mitchell IM, Bayen AM, Tomlin CJ (2005) A time-dependent Hamilton-Jacobi formulation of reachable sets for continuous dynamic games. IEEE Trans Autom Control 50(7):947–957

    Article  Google Scholar 

  • Mulder W, Osher S, Sethian JA (1992) Computing interface motion in compressible gas dynamics. J Comput Phys 100:209–228

    Article  Google Scholar 

  • Osher S, Fedkiw R (2003) Level set methods and dynamic implicit surfaces. Springer Verlag, Berlin Heidelberg New York

    Book  Google Scholar 

  • Osher S, Sethian JA (1988) Fronts propagating with curvature-dependent speed: algorithms based on Hamilton-Jacobi formulations. J Contemp Math 79(1):12–49

    Google Scholar 

  • Paley DA, Zhang F, Leonard NE (2008) Cooperative control for ocean sampling: the glider coordinated control system. IEEE Trans Consum Electron 16(4):735–744

    Google Scholar 

  • Pedlosky J (1998) Ocean Circulation Theory. Springer-Verlag, Berlin Heidelberg New York

    Google Scholar 

  • Pereira A, Binney J, Hollinger GA, Sukhatme GS (2013) Risk-aware path planning for autonomous underwater vehicles using predictive ocean models. J Field Robot 30(5):741–762. http://robotics.usc.edu/publications/801/

    Article  Google Scholar 

  • Petres C, Pailhas Y, Patron P, Petillot Y, Evans J, Lane D (2007) Path planning for autonomous underwater vehicles. IEEE Trans Robot 23(2):331–341

    Article  Google Scholar 

  • Ramp S, Davis R, Leonard N, Shulman I, Chao Y, Robinson A, Marsden J, Lermusiaux P, Fratantoni D, Paduan J, Chavez F, Bahr F, Liang S, Leslie W, Li Z (2009) Preparing to predict: the second autonomous ocean sampling network (AOSN-II) experiment in the Monterey Bay. Deep Sea Res II: Top Stud Oceanogr 56(35):68–86

    Article  Google Scholar 

  • Rao D, Williams SB (2009) Large-scale path planning for underwater gliders in ocean currents. In: Proceedings of australasian conference on robotics and automation

  • Rhoads B, Mezic I, Poje A (2010) Minimum time feedback control of autonomous underwater vehicles 49 th IEEE conference on Decision and control (CDC), pp 5828–5834

  • Russo G, Smereka P (2000) A remark on computing distance functions. J Comput Phys 163:51–67

    Article  Google Scholar 

  • Sapsis TP, Lermusiaux PFJ (2009) Dynamically orthogonal field equations for continuous stochastic dynamical systems. Physica D: Nonlinear Phenom 238(23–24):2347–2360

    Article  Google Scholar 

  • Schneider T, Schmidt H (2010) Unified command and control for heterogeneous marine sensing networks. J Field Robot 27(6):876–889

    Article  Google Scholar 

  • Schofield O, Glenn S, Orcutt J, Arrott M, Meisinger M, Gangopadhyay A, Brown W, Signell R, Moline M, Chao Y, Chien S, Thompson D, Balasuriya A, Lermusiaux P, Oliver M (2010) Automated sensor network to advance ocean science. EOS, Trans Am Geophys Union 91(39):345–346

    Article  Google Scholar 

  • Sethian JA (1999a) Fast marching methods. SIAM Rev 41(2):199–235

    Article  Google Scholar 

  • Sethian JA (1999b) Level set methods and fast marching methods: evolving interfaces in computational geometry, fluid mechanics, computer vision, and materials science. Cambridge University, Cambridge

    Google Scholar 

  • Simmonet E, Dijkstra H, Ghil M (2009) Bifurcation analysis of ocean, atmosphere, and climate models ‘Computational methods for the atmosphere and the oceans’, vol XIV, pp 187–229. Handbook of Numerical Analysis

  • Smith RN, Chao Y, Li PP, Caron DA, Jones BH, Sukhatme G S (2010) Planning and implementing trajectories for autonomous underwater vehicles to track evolving ocean processes based on predictions from a regional ocean model. Int J Robot Res 29(12):1475–1497

    Article  Google Scholar 

  • Soulignac M, Taillibert P, Rueher M (2009) Time-minimal path planning in dynamic current fields. In: IEEE international conference on Robotics and automation ICRA ’09, pp 2473–2479

  • Subramani D (2014) Energy optimal path planning in time-dependent flow fields using dynamically orthogonal level-set equations. Masters thesis, Computation for Design and Optimization (CDO) Program, Massachusetts Institute of Technology

  • Sussman M, Smereka P, Osher S (1994) A level set approach for computing solutions to incompressible two-phase flow. J Comput Phys 114:146–159

    Article  Google Scholar 

  • Thompson DR, Chien SA, Chao Y, Li P, Arrott M, Meisinger M, Balasuriya AP, Petillo S, Schofield O (2009) Glider Mission Planning in a Dynamic Ocean Sensorweb. In: SPARK workshop on scheduling and planning applications, international conference on automated planning and scheduling

  • Thompson DR, Chien SA, Chao Y, Li P, Cahill B, Levin J, Schofield O, Balasuriya AP, Petillo S, Arrott M, Meisinger M (2010) Spatiotemporal path planning in strong, dynamic, uncertain currents. In: Proceedings of IEEE international conference on robotics and automation, pp 4778–4783

  • Tonon D (2011) Regularity results for Hamilton-Jacobi Equations. PhD thesis, Scuola Internazionale Superiore di Studi Avanzati

  • Ueckermann MP, Lermusiaux PFJ (2011) 2.29 Finite Volume MATLAB Framework Documentation. Tech. rep., MSEAS Group, Massachusetts Institute of Technology, Cambridge, MA USA

  • Ueckermann M, Lermusiaux P, Sapsis T (2013) Numerical schemes for dynamically orthogonal equations of stochastic fluid and ocean flows. J Comput Phys 233:272–294

    Article  Google Scholar 

  • Van Leer B (1977) Towards the ultimate conservative difference scheme. IV. A new approach to numerical convection. J Comput Phys 23(3):276–299

    Article  Google Scholar 

  • Vasudevan C, Ganesan K (1996) Case-based path planning for autonomous underwater vehicles. Auton Robot 3:79–89

    Article  Google Scholar 

  • Wang D, Lermusiaux PFJ, Haley PJ, Eickstedt D, Leslie WG, Schmidt H (2009) Acoustically focused adaptive sampling and on-board routing for marine rapid environmental assessment. J Mar Syst 78:S393–S407

    Article  Google Scholar 

  • Warren CW (1990) A technique for autonomous underwater vehicle route planning. In: Proceedings of the symposium on Autonomous underwater vehicle technology, pp 201–205

  • Witt J, Dunbabin M (2008) Go with the flow: Optimal auv path planning in coastal environments. In: Proceedings of australasian conference on robotics and automation

  • Yang K, Gan S, Sukkarieh S (2010) An efficient path planning and control algorithm for ruavs in unknown and cluttered environments. J Intell Robot Syst 57:101–122

    Article  Google Scholar 

  • Yigit K (2011) Path planning methods for autonomous underwater vehicles. Masters thesis, Department of Mechanical Engineering, Massachusetts Institute of Technology

  • Yilmaz NK, Evangelinos C, Lermusiaux P, Patrikalakis NM (2008) Path planning of autonomous underwater vehicles for adaptive sampling using mixed integer linear programming. IEEE J Ocean Eng 33(4):522–537

    Article  Google Scholar 

  • Zhang F, Fratantoni DM, Paley D, Lund J, Leonard NE (2007) Control of coordinated patterns for ocean sampling. Int J Control 80(7):1186–1199

    Article  Google Scholar 

  • Zhang W, Inane T, Ober-Blobaum S, Marsden JE (2008) Optimal trajectory generation for a glider in time-varying 2D ocean flows B-spline model. In:, pp 1083–1088

Download references

Acknowledgments

We are very thankful to the MSEAS group in particular Mr. K. Yiğit and Mr. T. Sondergaard for the helpful discussions. We thank Dr. T. Duda for the suggestions on an earlier version of this manuscript and two anonymous reviewers for their comments. We are grateful to the Office of Naval Research for support under grants N00014-09-1-0676 (Science of Autonomy A-MISSION) and N00014-12-1-0944 (ONR6.2) to the Massachusetts Institute of Technology (MIT). MPU and PFJL also thank the Natural Sciences and Engineering Research Council (NSERC) of Canada for the Postgraduate Scholarship partially supporting the graduate studies and research of MPU at MIT.

Author information

Authors and Affiliations

Authors

Corresponding authors

Correspondence to Tapovan Lolla or Pierre F. J. Lermusiaux.

Additional information

Responsible Editor: Jörg-Olaf Wolff

Appendices

Appendix A: Preliminaries

In this Appendix, we describe some of the relevant definitions and terminology needed for the theoretical results. Most of the material presented in this Appendix may be found in the study of Bardi and Capuzzo-Dolcetta (2008), Clarke et al. (1998), Cannarsa and Sinestrari (2004), Frankowska (1989), and Bressan (2011). In what follows, we let \(n \in \mathbb {N}\), \({\Omega } \subseteq \mathbb {R}^{n}\) be an open set and \(\xi : {\Omega } \rightarrow \mathbb {R}\).

Remark 1

Let \(\xi \in \mathcal {C}({\Omega })\). Let + ξ(x 0) and ξ(x 0) denote the sets of super- and sub-differentials (Bardi and Capuzzo-Dolcetta 2008; Clarke et al. 1998) of ξ at x 0. Then, q + ξ(x 0) (resp. ξ(x 0)) if and only if there exists a function \(\gamma \in \mathcal {C}^{1}({\Omega })\) such that γ(x 0)=ξ(x 0), ∇γ(x 0)=q and the function γξ has a strict local minima (resp. maxima) at x 0.

Definition 1 (Generalized gradient)

Let ξ be locally Lipschitz at x 0. For any \(\mathbf {u} \in \mathbb {R}^{n}\), let \(\xi ^{g}(\mathbf {x}_{0};\mathbf {u} )\) denote the generalized directional derivative of ξ at x 0 (Clarke et al. 1998). The set of generalized gradients of ξ at x 0 is the non-empty set

$$\begin{array}{@{}rcl@{}} \partial \xi(\mathbf{x}_{0}) = \left \{ \mathbf{q} \in \mathbb{R}^{n}: \forall \, \mathbf{u} \in \mathbb{R}^{n}, \mathbf{q} \cdot \mathbf{u} \le \xi^{g}(\mathbf{x}_{0};\mathbf{u}) \right \}\, . \end{array} $$
(20)

Definition 2 (Regular function)

ξ is said to be regular at x 0∈Ω if it is Lipschitz near x 0 and admits directional derivatives \(\xi ^{d}(\mathbf {x}_{0};\mathbf {u})\) for all \(\mathbf {u} \in \mathbb {R}^{n}\), with ξ g(x 0;u)=ξ d(x 0;u).

Properties of regular functions

  1. 1.

    If ξ is continuously differentiable at x 0, then it is regular at x 0. Furthermore, ξ d(x 0;u)=∇ξ(x 0)⋅u=ξ g(x 0;u) for all \(\mathbf {u} \in \mathbb {R}^{n}\).

  2. 2.

    If ξ is convex and Lipschitz near x 0, then it is regular at x_0.

  3. 3.

    Let ξ be regular at x 0∈Ω. Then,

    $$ \partial_{-}\xi(\mathbf{x}_{0}) = \partial \xi(\mathbf{x}_{0}) \, . $$
    (21)

Definition 3 (Viscosity solution)

Let F≥0 and let V(x,t) satisfy assumptions 34. Consider the Hamilton-Jacobi equation

$$ \frac{\partial \phi}{\partial t} + F |\nabla \phi| + \mathbf{V}(\mathbf{x,t}) \cdot \nabla \phi = 0\quad\text{in}\quad{\Omega} \times (0,\infty) \, , $$
(22)

with initial conditions

$$ \phi(\mathbf{x},0) = \nu(\mathbf{x}) \, , $$
(23)

where \(\nu :{\Omega }\rightarrow \mathbb {R}\) is Lipschitz continuous. A function \(\phi \in \mathcal {C}({\Omega } \times [0,\infty ))\) is a viscosity subsolution of Eq. 22 if for every \((\mathbf {x,t})\in {\Omega } \times (0,\infty )\) and (q , p)∈ + ϕ(x , t),

$$ p + F |\mathbf{q}| + \mathbf{V}(\mathbf{x},t) \cdot \mathbf{q} \le 0 \, . $$
(24)

A function \(\phi \in \mathcal {C}({\Omega } \times [0,\infty ))\) is a viscosity supersolution of Eq. 22 if for every \((\mathbf {x,t})\in {\Omega }\times (0,\infty )\) and (q , p)∈ ϕ(x , t),

$$ p + F |\mathbf{q}| + \mathbf{V}(\mathbf{x},t) \cdot \mathbf{q} \ge 0 \, . $$
(25)

ϕ is said to be a viscosity solution of Eq. 22 if it is both a viscosity subsolution and a viscosity supersolution.

Theorem 1

(Frankowska 1989 ) A locally Lipschitz function \(\phi :{\Omega } \times (0,\infty )\rightarrow \mathbb {R}\) is a viscosity solution to Eq. 22 if and only if for every \((\mathbf {x,t}) \in {\Omega }\times (0,\infty )\) ,

$$ \max_{(\mathbf{q,p}) \in \partial \phi(\mathbf{x,t})} \{ p + F|\mathbf{q}| + \mathbf{V}(\mathbf{x},t) \cdot \mathbf{q} \}= 0\, $$
(26)

and for all (q,p)∈∂ ϕ(x,t),

$$ p + F|\mathbf{q}| + \mathbf{V}(\mathbf{x},t) \cdot \mathbf{q = 0} \, . $$
(27)

Theorem 2

(Clarke et al. 1998 ) [Lebourg’s Mean Value Theorem] Let \(\mathbb {S} \subseteq \mathbb {R}\) be an open set. Let \(x, y \in \mathbb {S}\) and suppose that \(f:\mathbb {S}\rightarrow \mathbb {R}\) is Lipschitz on an open set containing the segment [x,y]. Then, there exists 0<λ<1 such that

$$ f(y) - f(x) = g \times (y-x) \, , $$
(28)

for some g∈∂f(z), where z=λx+(1−λ)y.

Theorem 3

(Clarke et al. 1998 )[Chain Rule] Let \({\Omega }_{1} \subseteq \mathbb {R}^{n}\) and \({\Omega }_{2} \subseteq \mathbb {R}^{m}\) be two open sets with \(m, n \in \mathbb {N}\). Let \(\mathbf {g}:{\Omega }_{1}\rightarrow {\Omega }_{2}\) be continuously differentiable near x∈Ω1, and let \(F:{\Omega }_{2}\rightarrow \mathbb {R}\) be Lipschitz near g(x). Then, f :=Fg is Lipschitz near x and

$$ \partial f(\mathbf{x}) \subseteq \left (\mathbf{g}^{\prime}(\mathbf{x}) \right )^{*} \partial F \left (\mathbf{g}(\mathbf{x}) \right ) \, , $$
(29)

where denotes the adjoint.

Appendix B: Theoretical results

We now state a lemma that provides a monotonicity result related to ϕ, the viscosity solution of the Hamilton-Jacobi Equation (22). According to this result, the generalized gradient of ϕ is non-positive on trajectories \(\widetilde {\mathbf {X}}_{P}(\mathbf {y}_{\mathbf {s}},t)\), along the direction \(\left (\frac {\text {d} \widetilde {\mathbf {X}}_{P}(\mathbf {y}_{\mathbf {s}},t)}{\text {d} t},1\right )\) for t>0. This lemma is then used to prove Theorem 4, which establishes the relationship between reachable sets and the viscosity solution of a modified level set equation.

Lemma 1

Let \({\Omega } \subseteq \mathbb {R}^{n}\) be open, F>0 and let V(x,t) satisfy assumptions (3 )–( 4 ). Let ϕ be the viscosity solution to Eq. 22 . Let the trajectory \(\widetilde {\mathbf {X}}_{P}(\mathbf {y}_{\mathbf {s}},t)\) satisfy ( 1 ) with initial conditions \(\widetilde {\mathbf {X}}_{P}(\mathbf {y}_{\mathbf {s}},0) = \mathbf {y}_{\mathbf {s}}\). Then,

  1. 1.
    $$ p + \frac{\text{d} \widetilde{\mathbf{X}}_{P}(\mathbf{y}_{\mathbf{s}},t)}{\text{d} t} \cdot \mathbf{q} \le 0 \, \quad \forall \, (\mathbf{q,p}) \in \partial \phi(\widetilde{\mathbf{X}}_{P}(\mathbf{y}_{\mathbf{s}},t),t) $$
    (30)
  2. 2.
    $$ {\phi}^{g} \left (\widetilde{\mathbf{X}}_{P}(\mathbf{y}_{\mathbf{s}},t),t;\left (\frac{\text{d} \widetilde{\mathbf{X}}_{P}(\mathbf{y}_{\mathbf{s}},t)}{\text{d} t},1 \right ) \right ) \le 0 \quad \forall \, t>0 \, . $$
    (31)

The proof of this Lemma may be found in Lolla et al. (2014).

Theorem 4

Let \({\Omega } \subseteq \mathbb {R}^{n}\) be an open set, \(\mathbf {V}(\mathbf {x},t):{\Omega } \times [0,\infty ) \rightarrow \mathbb {R}^{n}\) satisfy ( 3 )–( 4 ), and F≥0. Let \({T}^{o}(\mathbf {y}):{\Omega } \rightarrow \mathbb {R}\) denote the optimal first arrival time at y. Let the trajectory \(\widetilde {\mathbf {X}}_{P}(\mathbf {y}_{\mathbf {s}},t)\) satisfy (1) with initial conditions \(\widetilde {\mathbf {X}}_{P}(\mathbf {y}_{\mathbf {s}},0) = \mathbf {y}_{\mathbf {s}}\) . Let ϕ o(x , t) be the viscosity solution to the Hamilton-Jacobi equation (9) with initial condition (10). Then,

  1. 1.

    \(\phi ^{o}(\widetilde {\mathbf {X}}_{P}(\mathbf {y}_{\mathbf {s}},t),t) \le 0\) for all t≥0.

  2. 2.

    If ϕ o is regular at \((\mathbf {X}^{o}_{P}(\mathbf {y}_{\mathbf {s}},t),t)\) for all t>0 and \(\mathbf {X}^{o}_{P}\) satisfies

    $$ \frac{\text{d} \mathbf{X}^{o}_{P}}{\text{d} t} = F \frac{\mathbf{q}^{o}}{|\mathbf{q}^{o}|} + \mathbf{V}(\mathbf{X}^{o}_{P}(\mathbf{y}_{\mathbf{s}},t),t), \quad t > 0, $$
    (32)

    for some \((\mathbf {q}^{o},p^{o}) \in \partial \phi ^{o}(\mathbf {X}^{o}_{P}(\mathbf {y}_{\mathbf {s}},t),t)\) , then

    $$ \phi^{o}(\mathbf{X}^{o}_{P}(\mathbf{y}_{\mathbf{s}},t),t) = 0 \quad \forall \, t\ge 0 \, . $$
  3. 3.
    $$ {T}^{o}(\mathbf{y}) = \inf_{t\ge 0} \{t:\phi^{o}(\mathbf{y,t}) = 0\} \,, $$
    (33)

    where \(\inf \) denotes the infimum.

  4. 4.

    The optimal trajectory to y f ∈Ω satisfies (11) whenever ϕ o is differentiable at \((\mathbf {X}^{\star }_{P}(\mathbf {y}_{\mathbf {s}},t),t)\) and \(|\nabla \phi ^{o} (\mathbf {X}^{\star }_{P},t)| \ne 0\).

  5. 5.

    If \(F > \max \limits _{\mathbf {x} \in {\Omega }, t\ge 0} |\mathbf {V}(\mathbf {x},t)|\), then T o(y) is the viscosity solution of the modified Eikonal equation

    $$ F |\nabla {T}^{o}(\mathbf{y})| + \mathbf{V}(\mathbf{y},{T}^{o}(\mathbf{y})) \cdot \nabla {T}^{o}(\mathbf{y}) - 1= 0 \, , \mathbf{y} \in {\Omega} \, . $$
    (34)

Proof

  • 1. The viscosity solution to Eq. 9 is locally Lipschitz (see Tonon (2011), Bianchini and Tonon (2012), and Cannarsa and Sinestrari (2004)). We now argue that \({\phi ^{o}_{P}}(t) := \phi ^{o}(\widetilde {\mathbf {X}}_{P}(\mathbf {y}_{\mathbf {s}},t),t)\) is locally Lipschitz for all t≥0. Observe that \({\phi ^{o}_{P}}(t) = \phi ^{o}(\mathbf {g}_{P}(t))\) where \(\mathbf {g}_{P}(t) := (\widetilde {\mathbf {X}}_{P}(\mathbf {y}_{\mathbf {s}},t),t)\). Since g P (t) is continuously differentiable in \((0,\infty )\) with \(\frac {\text {d} \mathbf {g}_{P}(t)}{\text {d} t} = \left (\frac {\text {d} \widetilde {\mathbf {X}}_{P}}{\text {d} t},1 \right )\) and ϕ o is locally Lipschitz, \({\phi ^{o}_{P}}(t)\) is also locally Lipschitz in \((0,\infty )\) by the chain rule stated in Theorem 3.

    Let t 1 > 0 be fixed. Since \({\phi ^{o}_{P}}\) is locally Lipschitz, there exists an open interval around t 1 in which \({\phi ^{o}_{P}}\) is Lipschitz. Thus, for any t 2>t 1 in this interval, Lebourg’s Mean Value Theorem (Theorem 2) implies there exist t 3∈(t 1,t 2) and \(s \in \partial {\phi ^{o}_{P}} (t_{3})\) such that

    $$ {\phi^{o}_{P}}(t_{2}) - {\phi^{o}_{P}}(t_{1}) = s \times (t_{2} - t_{1}) \, . $$
    (35)

    Using the chain rule of Theorem 3 again ( denotes the adjoint),

    $$\begin{array}{@{}rcl@{}} \partial {\phi^{o}_{P}}(t_{3}) &\subseteq \, \left (\mathbf{g}_{P}^{\prime}(t_{3}) \right )^{*} \partial \phi^{o} \left (\mathbf{g}_{P}(t_{3}) \right ) \\ & = \left \{ p + \mathbf{q}\cdot \frac{\text{d} \widetilde{\mathbf{X}}_{P}(\mathbf{y}_{\mathbf{s}},t_{3})}{\text{d} t} \right . \\ &\qquad :\left . (\mathbf{q,p}) \in \partial \phi^{o}(\widetilde{\mathbf{X}}_{P}(\mathbf{y}_{\mathbf{s}},t_{3}),t_{3}) \right \}\, . \end{array} $$
    (36)

    Hence, any \(s \in \partial {\phi ^{o}_{P}}(t_{3})\) can be written as

    $$ s = p + \mathbf{q} \cdot \frac{\text{d} \widetilde{\mathbf{X}}_{P}(\mathbf{y}_{\mathbf{s}},t_{3})}{\text{d} t} \, . $$
    (37)

    for some \((\mathbf {q,p}) \in \partial \phi ^{o}(\widetilde {\mathbf {X}}_{P}(\mathbf {y}_{\mathbf {s}},t_{3}),t_{3})\). From Eq. 30,

    $$\begin{array}{@{}rcl@{}} p + \mathbf{q} \cdot \frac{\text{d} \widetilde{\mathbf{X}}_{P}(\mathbf{y}_{\mathbf{s}},t_{3})}{\text{d} t} &\le 0 \, \end{array} $$

    implying that for any \(s \in \partial {\phi ^{o}_{P}}(t_{3})\), s≤0. Using this result in Eq 35 yields \({\phi ^{o}_{P}}(t_{2}) \le {\phi ^{o}_{P}}(t_{1})\) for all t 1,t 2. Since \({\phi ^{o}_{P}}\) is locally Lipschitz in \((0,\infty )\), we conclude that \({\phi ^{o}_{P}}(t)\) is non-increasing on \((0,\infty )\). Moreover, since \({\phi ^{o}_{P}}\) is continuous on \([0,\infty )\), with \({\phi ^{o}_{P}}(0) = 0\) (from Eq. 10) and non-increasing in \((0,\infty )\), we have \({\phi ^{o}_{P}}(t) = \phi (\widetilde {\mathbf {X}}_{P}(\mathbf {y}_{\mathbf {s}},t),t) \le 0\) for all t≥0.

  • When the trajectory \(\mathbf {X}^{o}_{P}(\mathbf {y}_{\mathbf {s}},t)\) is regular, i.e., when ϕ o is regular at points \((\mathbf {X}^{o}_{P}(\mathbf {y}_{\mathbf {s}},t),t)\) for all t > 0, Eq. 21 implies \(\partial _{-} \phi ^{o}(\mathbf {X}^{o}_{P}(\mathbf {y}_{\mathbf {s}},t),t) = \partial \phi ^{o}(\mathbf {X}^{o}_{P}(\mathbf {y}_{\mathbf {s}},t),t)\) for all t > 0. Since ϕ o is the viscosity solution to Eq. 9, we obtain from Theorem 1 that for any t > 0, \(p + F|\mathbf {q}| + \mathbf {V}(\mathbf {X}^{o}_{P}(\mathbf {y}_{\mathbf {s}},t),t) \cdot \mathbf {q} = 0\) for all \((\mathbf {q,p}) \in \partial _{-} \phi ^{o}(\mathbf {X}^{o}_{P}(\mathbf {y}_{\mathbf {s}},t),t) = \partial \phi ^{o}(\mathbf {X}^{o}_{P}(\mathbf {y}_{\mathbf {s}},t),t)\). Specifically, for the member (q o,p o) of \(\partial \phi ^{o}(\mathbf {X}^{o}_{P}(\mathbf {y}_{\mathbf {s}},t),t)\) that satisfies (32), the definition of generalized gradient (20) implies

    $$\begin{array}{@{}rcl@{}} {\phi^{o}}^{g} &\left (\mathbf{X}^{o}_{P}(\mathbf{y}_{\mathbf{s}},t),t;\left (\frac{\text{d} \mathbf{X}^{o}_{P}(\mathbf{y}_{\mathbf{s}},t)}{\text{d} t},1\right ) \right ) \\ &\ge p^{o} + \mathbf{q}^{o} \cdot \frac{\text{d} \mathbf{X}^{o}_{P}(\mathbf{y}_{\mathbf{s}},t)}{\text{d} t} \\ &= p^{o} + F |\mathbf{q}^{o}| + \mathbf{V}(\mathbf{X}^{o}_{P}(\mathbf{y}_{\mathbf{s}},t),t) \cdot \mathbf{q}^{o} \\ &= 0 \,. \end{array} $$
    (38)

    Combining this result with Eq. 31, we obtain

    $${\phi^{o}}^{g}\left (\mathbf{X}^{o}_{P}(\mathbf{y}_{\mathbf{s}},t),t;\left (\frac{\text{d} \mathbf{X}^{o}_{P}(\mathbf{y}_{\mathbf{s}},t)}{\text{d} t},1\right ) \right ) = 0 \, . $$

    Since ϕ o is regular at \(\left (\mathbf {X}^{o}_{P}(\mathbf {y}_{\mathbf {s}},t),t \right )\) by assumption, we then also have

    $$ {\phi^{o}}^{d}\left (\mathbf{X}^{o}_{P}(\mathbf{y}_{\mathbf{s}},t),t;\left (\frac{\text{d} \mathbf{X}^{o}_{P}(\mathbf{y}_{\mathbf{s}},t)}{\text{d} t},1\right ) \right ) = 0 \, . $$
    (39)

    For any h > 0, the definition of \({\phi ^{o}_{P}}\) implies

    $$\begin{array}{@{}rcl@{}} & \left | \frac{{\phi^{o}_{P}}(t+h) - {\phi^{o}_{P}}(t)}{h} \right | \\ &=\left | \frac{\phi^{o} \left (\mathbf{X}^{o}_{P}(\mathbf{y}_{\mathbf{s}},t+h),t+h \right ) - \phi^{o} \left (\mathbf{X}^{o}_{P}(\mathbf{y}_{\mathbf{s}},t),t\right )}{h} \right | \, . \end{array} $$
    (40)

    Since ϕ o is locally Lipschitz, ∃ C > 0 such that for h > 0 small enough,

    $$\begin{array}{@{}rcl@{}} &\left | \phi \left (\mathbf{X}^{o}_{P}(\mathbf{y}_{\mathbf{s}},t+h),t+h \right ) - \phi^{o} \left (\mathbf{X}^{o}_{P}(\mathbf{y}_{\mathbf{s}},t) + h\frac{\text{d} \mathbf{X}^{o}_{P}}{\text{d} t},t+h \right ) \right | \\ &\le C \left | \mathbf{X}^{o}_{P}(\mathbf{y}_{\mathbf{s}},t+h) - \mathbf{X}^{o}_{P}(\mathbf{y}_{\mathbf{s}},t) - h\frac{\text{d} \mathbf{X}^{o}_{P}}{\text{d} t}\right | \\ &= C \left | \mathbf{o}(h) \right | \, , \end{array} $$
    (41)

    where \(\mathbf {o}(h) \in \mathbb {R}^{n}\) denotes a vector whose individual terms are o(h). Adding and subtracting \(\phi ^{o} \left (\mathbf {X}^{o}_{P} + h\frac {\text {d} \mathbf {X}^{o}_{P}}{\text {d} t},t+h \right )\) from the numerator of Eq. 40 and using the triangle inequality, we obtain

    $$\begin{array}{cl} &\left | \frac{{\phi^{o}_{P}}(t+h) - {\phi^{o}_{P}}(t)}{h} \right | \\ &\le \left | \frac{\phi^{o} \left (\mathbf{X}^{o}_{P}(\mathbf{y}_{\mathbf{s}},t) + h\frac{\text{d} \mathbf{X}^{o}_{P}(\mathbf{y}_{\mathbf{s}},t)}{\text{d} t},t+h \right ) - \phi^{o} \left (\mathbf{X}^{o}_{P}(\mathbf{y}_{\mathbf{s}},t),t\right )}{h} \right | \\ &\qquad+ C\left | \frac{\mathbf{o}(h)}{h}\right |. \end{array} $$

    The first term on the right converges to:\({\phi ^{o}}^{d}\left (\mathbf {X}^{o}_{P},t;\left (\frac {\text {d} \mathbf {X}^{o}_{P}}{\text {d} t},1\right ) \right )\) as h 0 and by Eq. 39, its value is zero. The second term uniformly converges to zero as h 0, by definition. This implies

    $$\lim\limits_{h \downarrow 0} \left | \frac{{\phi^{o}_{P}}(t+h) - {\phi^{o}_{P}}(t)}{h} \right | = 0 \, , $$

    and consequently that

    $$ \lim\limits_{h \downarrow 0} \frac{{\phi^{o}_{P}}(t+h) - {\phi^{o}_{P}}(t)}{h} = 0 \, . $$
    (42)

    Since Eq. 42 holds for all t > 0, \({\phi ^{o}_{P}}\) is right differentiable in \((0,\infty )\) and the value of the right-derivative is zero for all t > 0. This implies that \({\phi ^{o}_{P}}\) is constant in \((0,\infty )\). Since \({\phi ^{o}_{P}}(0) = 0\), we obtain \({\phi ^{o}_{P}}(t) = \phi ^{o}(\mathbf {X}^{o}_{P}(\mathbf {y}_{\mathbf {s}},t),t)=0\) for all t≥0. Therefore, trajectories \(\mathbf {X}^{o}_{P}(\mathbf {y}_{\mathbf {s}},t)\) that are regular and satisfy (32) always remain on the zero-level set of ϕ o.

  • It has been shown in part (1) that \(\phi ^{o}(\widetilde {\mathbf {X}}_{P}(\mathbf {y}_{\mathbf {s}},t),t) \le 0\) for all t≥0 for any trajectory \(\widetilde {\mathbf {X}}_{P}(\mathbf {y}_{\mathbf {s}},t)\) that satisfies (1) and the initial conditions \(\widetilde {\mathbf {X}}_{P}(\mathbf {y}_{\mathbf {s}},0) = \mathbf {y}_{\mathbf {s}}\). Therefore, for a trajectory X P (y s ,t) that reaches a given end point y∈Ω at time \(\widetilde {T}(\mathbf {y})\) (not necessarily optimal),

    $$ \phi^{o}(\mathbf{y}, \widetilde{T}(\mathbf{y})) = \phi^{o}(\widetilde{\mathbf{X}}_{P}(\mathbf{y}_{\mathbf{s}},\widetilde{T}(\mathbf{y})),\widetilde{T}(\mathbf{y}))\le 0 \, . $$
    (43)

    Since this inequality holds for any arbitrary arrival time \(\widetilde {T}(\mathbf {y})\), it will also hold for the optimal arrival time T o(y), implying

    $$ \phi^{o}(\mathbf{y}, {T}^{o}(\mathbf{y})) \le 0 \,~ \text{for all}\quad \mathbf{y} \in {\Omega} \, . $$
    (44)

    For y=y s , Eq. 33 holds trivially. For any yy s , ϕ o(y,0) > 0 by Eq. 10. The continuity of ϕ o and Eq. 44 together then yield

    $$ {T}^{o}(\mathbf{y}) \ge \inf_{t\ge 0}\{t: \phi^{o}(\mathbf{y,t}) = 0\} \, . $$
    (45)

    In part (2), we showed the existence of trajectories that always remain on the zero level set of ϕ o. Furthermore, any point on the zero level set of ϕ o belongs to a characteristics of Eq. 9 emanating from y s since y s is the only point in Ω where ϕ o is initially zero. Therefore, when the zero level set reaches y for the first time, it implies the existence of a trajectory \(\mathbf {X}^{o}_{P}(\mathbf {y}_{\mathbf {s}},t)\) with \(\mathbf {X}^{o}_{P}(\mathbf {y}_{\mathbf {s}},0) = \mathbf {y}_{\mathbf {s}}\) that satisfies (1). For this trajectory, (45) holds with an equality, thereby establishing (33). Physically, this means that fastest arrival time at any end point y∈Ω is when the zero level set of ϕ o reaches y for the first time, and equivalently that the reachability front \(\partial \mathcal {R}(\mathbf {y}_{\mathbf {s}},t)\) coincides with the zero level set of ϕ o at time t.

  • Let y f ∈Ω be fixed. From part (3), the optimal trajectory to y f satisfies \(\phi ^{o}(\mathbf {X}^{\star }_{P}(\mathbf {y}_{\mathbf {s}},t),t) = 0\) for all t≥0. Hence, \({\phi ^{o}_{P}}(t) := \phi ^{o}(\mathbf {X}^{\star }_{P}(\mathbf {y}_{\mathbf {s}},t),t)\) equals zero for all \(0 \le t \le {T}^{\star }(\mathbf {y}_{\mathbf {f}})\). Let us fix a time \(0 < t < {T}^{\star }(\mathbf {y}_{\mathbf {f}})\) such that ϕ o is differentiable at \((\mathbf {X}^{\star }_{P}(\mathbf {y}_{\mathbf {s}},t),t)\). The usual chain rule then yields

    $$ 0 = \frac{\text{d} {\phi^{o}_{P}}(t)}{\text{d} t} = \frac{\partial \phi^{o}}{\partial t} + \nabla \phi^{o} \cdot \frac{\text{d} \mathbf{X}^{\star}_{P}(\mathbf{y}_{\mathbf{s}},t)}{\text{d} t} \, , $$
    (46)

    where the derivatives of ϕ o are evaluated at \((\mathbf {X}^{\star }_{P}(\mathbf {y}_{\mathbf {s}},t),t)\). Since ϕ o is assumed to be differentiable at this point, Eq. 9 holds in the classical sense and \(\frac {\partial \phi ^{o}}{\partial t} = - F |\nabla \phi ^{o} (\mathbf {X}^{\star }_{P},t)| - \mathbf {V}(\mathbf {X}^{\star }_{P},t)\cdot \nabla \phi ^{o} (\mathbf {X}^{\star }_{P},t)\). Substituting this in Eq. 46 gives

    $$ \frac{\text{d} \mathbf{X}^{\star}_{P}}{\text{d} t} \cdot \nabla \phi^{o}(\mathbf{X}^{\star}_{P},t) = F |\nabla \phi^{o} (\mathbf{X}^{\star}_{P},t)| + \mathbf{V}(\mathbf{X}^{\star}_{P},t) \cdot \nabla\phi^{o} (\mathbf{X}^{\star}_{P},t) \, . $$
    (47)

    Using Eq. 1,

    $$\begin{array}{cl} &\frac{\text{d} \mathbf{X}^{\star}_{P}}{\text{d} t} \cdot \nabla \phi^{o}(\mathbf{X}^{\star}_{P},t) \\ &= F_{P}^{\star}(t)\, \hat{h}^{\star}(t) \cdot \nabla \phi^{o}(\mathbf{X}^{\star}_{P},t) + \mathbf{V}(\mathbf{X}^{\star}_{P},t) \cdot \nabla\phi^{o} (\mathbf{X}^{\star}_{P},t) \, \\ &\le F |\nabla \phi^{o} (\mathbf{X}^{\star}_{P},t)| + \mathbf{V}(\mathbf{X}^{\star}_{P},t) \cdot \nabla\phi^{o} (\mathbf{X}^{\star}_{P},t) \, , \end{array} $$

    equality holding iff \(F_{P}^{\star }(t) = F\) and \(\hat {h}^{\star }(t) = \frac {\nabla \phi ^{o} (\mathbf {X}^{\star }_{P},t)}{|\nabla \phi ^{o} (\mathbf {X}^{\star }_{P},t)|}\), for \(|\nabla \phi ^{o} (\mathbf {X}^{\star }_{P},t)| \ne 0\). Using this result in Eq. 47 yields

    $$ \frac{\text{d} \mathbf{X}^{\star}_{P}}{\text{d} t} = F \, \frac{\nabla \phi^{o} (\mathbf{X}^{\star}_{P},t)}{|\nabla \phi^{o} (\mathbf{X}^{\star}_{P},t)|} + \mathbf{V}(\mathbf{X}^{\star}_{P},t) \, . $$
  • Under the assumption \(F > \sup _{\mathbf {x} \in {\Omega }, t\ge 0} \{ |\mathbf {V}(\mathbf {x},t)| \}\), the start point y s belongs to the interior of the reachable set \(\mathcal {R}(\mathbf {y}_{\mathbf {s}},t)\) for all t>0, i.e., for any t > 0, there exists 𝜖 t > 0 such that all points \(\mathbf {x}^{\prime }\) satisfying \(|\mathbf {y}_{\mathbf {s}} - \mathbf {x}^{\prime }| < \epsilon _{t}\) are members of \(\mathcal {R}(\mathbf {y}_{\mathbf {s}},t)\). This condition is equivalent to the “small time local controllability” condition discussed in the study of Bardi and Capuzzo-Dolcetta (2008) as a result of which, T o is continuous in Ω. See (Bardi and Capuzzo-Dolcetta 2008) for a formal proof of this statement. Let us fix y∈Ω. By definition, T o(y) satisfies

    $$ {T}^{o}(\mathbf{y}) = \inf_{h>0} \{ {T}^{o}(\widetilde{\mathbf{y}}) + h\} \, , $$
    (48)

    where \(\widetilde {\mathbf {y}} \in {\Omega }\) is a point such that there exists a trajectory \(\mathbf {\widetilde {X}}_{P}(\mathbf {y}_{\mathbf {s}},t)\) satisfying (1) and the limiting conditions

    $$ \mathbf{\widetilde{X}}_{P}(\mathbf{y}_{\mathbf{s}},{T}^{o}(\mathbf{\widetilde{y}})) = \mathbf{\widetilde{y}},~~~ \mathbf{\widetilde{X}}_{P}(\mathbf{y}_{\mathbf{s}},{T}^{o}(\mathbf{\widetilde{y}})+h) = \mathbf{y} \, . $$
    (49)

    In order to show that T o is a viscosity solution to Eq. 34, we show that it is both a viscosity subsolution and a supersolution to Eq. 34.

    Viscosity subsolution From Definition 3 and Remark 1, \({T}^{o} \in \mathcal {C}({\Omega })\) is a viscosity subsolution to Eq. 34 if at every y∈Ω and for every \(\mathcal {C}^{1}\) function \(\tau _{s}:{\Omega } \rightarrow \mathbb {R}\) such that \(\tau _{s}(\mathbf {y}) = {T}^{o}(\mathbf {y})\) and \(\tau _{s} - {T}^{o}\) has a local minima at y,

    $$ F \, |\nabla \tau_{s} (\mathbf{y})| + \mathbf{V}(\mathbf{y}, {T}^{o}(\mathbf{y})) \cdot \nabla \tau_{s}(\mathbf{y}) - 1 \le 0 \, . $$
    (50)

    Since \(\tau _{s} \ge {T}^{o}\) in a neighborhood of y, we obtain for h>0 small enough,

    $$\tau_{s}(\mathbf{y}) - \tau_{s}(\widetilde{\mathbf{y}}) \le {T}^{o}(\mathbf{y}) - {T}^{o}(\widetilde{\mathbf{y}}) \, . $$

    Moreover, for this choice of h and the resulting \(\widetilde {\mathbf {y}}\), (48) implies

    $${T}^{o}(\mathbf{y}) \le {T}^{o}(\widetilde{\mathbf{y}}) + h \, . $$

    Combining the above two inequalities yields

    $$ \tau_{s}(\mathbf{y}) - \tau_{s}(\widetilde{\mathbf{y}}) \le {T}^{o}(\mathbf{y}) - {T}^{o}(\widetilde{\mathbf{y}}) \le h \, . $$
    (51)

    Since τ s is differentiable at y, Taylor’s theorem may be used to expand \(\tau _{s}(\widetilde {\mathbf {y}})\) near y.

    $$\begin{array}{@{}rcl@{}} \tau_{s}(\widetilde{\mathbf{y}}) &= \tau_{s}(\mathbf{y}) + \nabla \tau_{s}(\mathbf{y}) \cdot (\widetilde{\mathbf{y}} -\mathbf{y}) + \mathbf{o} \left (|\widetilde{\mathbf{y}} - \mathbf{y}| \right ) \, \\ &= \tau_{s}(\mathbf{y}) - {\int}_{{T}^{o}(\widetilde{\mathbf{y}})}^{{T}^{o}(\widetilde{\mathbf{y}})+h} \nabla \tau_{s}(\mathbf{y}) \cdot \frac{\text{d} \widetilde{\mathbf{X}}_{P}}{\text{d} t} \, \text{d}t + \mathbf{o} \left (|\widetilde{\mathbf{y}} - \mathbf{y}| \right ).\\ \end{array} $$
    (52)

    Inserting Eq. 52 in Eq. 51 and dividing by h,

    $$\frac{1}{h} {\int}_{{T}^{o}(\widetilde{\mathbf{y}})}^{{T}^{o}(\widetilde{\mathbf{y}})+h} \nabla \tau_{s}(\mathbf{y}) \cdot \frac{\text{d} \widetilde{\mathbf{X}}_{P}}{\text{d} t} \, \text{d}t + \frac{\mathbf{o} \left (|\widetilde{\mathbf{y}} - \mathbf{y}| \right )}{h} \le 1 \, . $$

    As h 0 and after noting that the second term on the left vanishes under this limit, we obtain

    $$ \nabla \tau_{s}(\mathbf{y}) \cdot \frac{\text{d} \widetilde{\mathbf{X}}_{P}}{\text{d} t}(\mathbf{y}_{\mathbf{s}},{T}^{o}(\mathbf{y})) \le 1 \, . $$
    (53)

    One can see that Eq. 50 is satisfied trivially when |∇τ s (y)|=0. Thus, we may assume |∇τ s (y)|≠0. Since (53) holds for any valid choice of \(\frac {\text {d} \widetilde {\mathbf {X}}_{P}}{\text {d} t}\), we may choose \(\frac {\text {d} \widetilde {\mathbf {X}}_{P}}{\text {d} t} (\mathbf {y}_{\mathbf {s}},{T}^{o}(\mathbf {y}))= F \frac {\nabla \tau _{s}(\mathbf {y})}{|\nabla \tau _{s}(\mathbf {y})|} + \mathbf {V}(\mathbf {y},{T}^{o}(\mathbf {y}))\) to obtain

    $$\begin{array}{cl} &\nabla \tau_{s}(\mathbf{y}) \cdot \left (F \frac{\nabla \tau_{s}(\mathbf{y})}{|\nabla \tau_{s}(\mathbf{y})|} + \mathbf V(\mathbf{y},{T}^{o}(\mathbf{y})) \right ) \\ &= F|\nabla \tau_{s}(\mathbf{y})| + \mathbf{V}(\mathbf{y},{T}^{o}(\mathbf{y})) \cdot \nabla \tau_{s}(\mathbf{y}) \le 1 \, , \end{array} $$

    thereby establishing (50). Therefore, T o is a viscosity subsolution to Eq. 34.

    Viscosity supersolution T o is a viscosity supersolution to Eq. 34 if at any y∈Ω and for every \(\mathcal {C}^{1}\) function \(\tau ^{s}:{\Omega } \rightarrow \mathbb {R}\) such that τ s(y)=T o(y) and τ sT o has a local maxima at y,

    $$ F |\nabla \tau^{s} (\mathbf{y})| + \mathbf{V}(\mathbf{y}, {T}^{o}(\mathbf{y})) \cdot \nabla \tau^{s}(\mathbf{y}) - 1 \ge 0 \, . $$
    (54)

    For any 0 < h < T o(y), there exists \(\widehat {\mathbf y} \in {\Omega }\) satisfying \({T}^{o}(\widehat {\mathbf {y}}) + h = {T}^{o}(\mathbf {y})\) and a trajectory \(\widehat {\mathbf X}_{P}(\mathbf {y}_{\mathbf {s}},t)\) satisfying (1) and the limiting conditions

    $$ \widehat{\mathbf{X}}_{P}(\mathbf{y}_{\mathbf{s}},{T}^{o}(\widehat{\mathbf y})) = \widehat{\mathbf{y}},~~~ \widehat{\mathbf{X}}_{P}(\mathbf{y}_{\mathbf{s}},{T}^{o}(\mathbf{y})) = \mathbf{y} \, . $$
    (55)

    Of course, the optimal trajectory leading to y is a valid choice for \(\widehat {\mathbf X}_{P}(\mathbf {y}_{\mathbf {s}},t)\). For h > 0 small enough,

    $$ h = {T}^{o}(\mathbf{y}) - {T}^{o}(\widehat{\mathbf{y}}) \le \tau^{s}(\mathbf{y}) - \tau^{s}(\widehat{\mathbf{y}}) \, . $$
    (56)

    As in the earlier sub-section, we may use Taylor’s theorem to expand \(\tau ^{s}(\widehat {\mathbf {y}})\) near y to obtain

    $$\begin{array}{@{}rcl@{}} \tau^{s}(\widehat{\mathbf{y}}) &= \tau^{s}(\mathbf{y}) + \nabla \tau^{s}(\mathbf{y}) \cdot (\widehat{\mathbf{y}} -\mathbf{y}) + \mathbf{o} \left (|\widehat{\mathbf{y}} - \mathbf{y}| \right ) \, \\ &= \tau^{s}(\mathbf{y}) - {\int}_{{T}^{o}(\widehat{\mathbf{y}})}^{{T}^{o}(\widehat{\mathbf{y}})+h} \nabla \tau^{s}(\mathbf{y}) \cdot \frac{\text{d} \widehat{\mathbf{X}}_{P}}{\text{d} t} \, \text{d}t + \mathbf{o} \left (|\widehat{\mathbf{y}} - \mathbf{y}| \right ). \end{array} $$
    (57)

    Inserting (57) in Eq. 56 and dividing by h,

    $$ \frac{1}{h} {\int}_{{T}^{o}(\widehat{\mathbf{y}})}^{{T}^{o}(\widehat{\mathbf{y}})+h} \nabla \tau^{s}(\mathbf{y}) \cdot \frac{\text{d} \widehat{\mathbf{X}}_{P}}{\text{d} t} \, \text{d}t + \frac{\mathbf{o} \left (|\widehat{\mathbf{y}} - \mathbf{y}| \right )}{h} \ge 1 \, . $$
    (58)

    Observe that from Eq. 1,

    $$\nabla \tau^{s}(\mathbf{y}) \cdot \frac{\text{d} \widehat{\mathbf{X}}_{P}(\mathbf{y}_{\mathbf{s}},t)}{\text{d} t} \le F |\nabla \tau^{s}(\mathbf{y})| + \mathbf{V}(\widehat{\mathbf{X}}_{P}(\mathbf{y}_{\mathbf{s}},t),t) \cdot \nabla \tau^{s}(\mathbf{y}) \, . $$

    Taking limits of Eq. 58 as h 0 gives

    $$1 \le \nabla \tau^{s}(\mathbf{y}) \cdot \frac{\text{d} \widehat{\mathbf X}_{P}}{\text{d} t}(\mathbf{y}_{\mathbf{s}},{T}^{o}(\mathbf{y})) \le F |\nabla \tau^{s}(\mathbf{y})| + \mathbf{V}(\mathbf{y},{T}^{o}(\mathbf{y})) \cdot \nabla \tau^{s}(\mathbf{y}) \, , $$

    which proves that T o is a viscosity supersolution of Eq. 34. Therefore, T o is a viscosity solution to Eq. 34. □

Appendix C: Numerical schemes

We now summarize the numerical schemes utilized to discretize and solve (9) and (12). These equations are solved using a finite volume framework implemented in MATLAB. The term |∇ϕ o| in (9) is discretized using either a first order (Sethian 1999b; Lolla 2012) or a higher order (Yigit 2011) upwind scheme and V(x,t)⋅∇ϕ o is discretized using a second order TVD scheme on a staggered C-grid (Ueckermann and Lermusiaux 2011).

1.1 C.1 Forward level set evolution

We discretize (9) in time using a fractional step method as follows:

$$\begin{array}{@{}rcl@{}} \frac{\bar{\phi} - \phi^{o}(\mathbf{x},t)}{\Delta t/2} &= -F|\nabla \phi^{o}(\mathbf{x},t)| \end{array} $$
(59)
$$\begin{array}{@{}rcl@{}} \frac{\bar{\bar{\phi}} - \bar{\phi}}{\Delta t} = -\mathbf{V}\left (\mathbf{x},t+\frac{\Delta t}{2} \right ) \cdot \nabla \bar{\phi} \end{array} $$
(60)
$$\begin{array}{@{}rcl@{}} \frac{\phi^{o}(\mathbf{x},t+{\Delta} t) - \bar{\bar{\phi}}}{\Delta t/2} &= -F |\nabla \bar{\bar{\phi}}| \end{array} $$
(61)

Equations 5961 are solved only in the interior nodes of the discretized system. For boundaries that are open inlets/outlets or side walls (i.e., not interior obstacles nor forbidden regions), open boundary conditions are used on ϕ o and on the intermediate variables \(\bar {\phi }\) and \(\bar {\bar {\phi }}\) at each time step. Specifically, a radiation boundary condition with infinite wave speed is assumed, which amounts to an internal zero normal gradient (Neumann) condition, so the boundary values are updated by replacing them with the value of the variable one cell interior to the boundary. Obstacles and forbidden regions in the domain are masked, i.e., (9) is solved only at interior nodes not lying under these regions. For points adjacent to the mask, open boundary conditions are implemented and necessary spatial gradients are evaluated using neighboring nodes that do not lie under the mask. As a result, the value of ϕ o under the mask is never used in the computation. We note that in some situations, more complex open boundary conditions could be used as done in regional ocean modeling (Lermusiaux 1997; Haley Jr. and Lermusiaux 2010). We have implemented the narrow-band scheme of Adalsteinsson and Sethian (1995) to solve (59)–(61).

The reachability front \(\partial \mathcal {R}_{\phi ^{o}}(t)\) is extracted from the ϕ o field at every time step using a contour algorithm. In a 2D problem, the amount of storage required for this is not significant because \(\partial \mathcal {R}_{\phi ^{o}}(t)\) is a 1D curve which is numerically represented by a finite number of points. We also note that this contour extraction is not needed: We could simply store the times when the zero contour of ϕ o crosses each grid point in order to compute the normals for the backtracking (Yigit 2011).

C.2 Backtracking

(12) is discretized using first order (Lolla 2012) or higher-order (Yigit 2011) time integration schemes. Ideally, it suffices to solve Eq. 9 until the level set front first reaches y f . However, due to the discrete time steps, a more convenient stopping criterion is the first time, T, when \(\phi ^{o}(\mathbf {y}_{\mathbf {f}},T) \le 0\). Due to this, y f does not lie on the final contour \(\partial \mathcal {R}_{\phi ^{o}}(t)\) exactly. Thus, we first project y f onto \(\partial \mathcal {R}_{\phi ^{o}}(t)\). The projected \(\hat {\mathbf {n}}_{p}\) is computed as the unit normal to \(\partial \mathcal {R}_{\phi ^{o}}(t)\) at the projected point. The discretized form of Eq. 12,

$$ \frac{\mathbf{X}^{\star}_{P}(\mathbf{y}_{\mathbf{s}},t-{\Delta} t) - \mathbf{X}^{\star}_{P}(\mathbf{y}_{\mathbf{s}},t)}{\Delta t} = -\mathbf{V}(\mathbf{X}^{\star}_{P},t) - F\underbrace{\frac{\nabla \phi^{o}(\mathbf{X}^{\star}_{P},t)}{|\nabla \phi^{o}(\mathbf{X}^{\star}_{P},t)|}}_{\hat{\mathbf{n}}_{p}(\mathbf{x},t)} \, , $$
(62)

is marched back in time until we reach a point on the first saved contour and this generates a discrete representation of \(\mathbf {X}^{\star }_{P}(\mathbf {y}_{\mathbf {s}},t)\). Along the way, we project each newly computed trajectory point, \(\mathbf {X}^{\star }_{P}(\mathbf {y}_{\mathbf {s}},t-{\Delta } t)\) onto the corresponding intermediate level set contour (see Lolla (2012)). Instead of performing these projections, one can use the two intermediate discrete level set contours between which an unprojected trajectory point lies to interpolate either the normal \(\hat {\mathbf {n}}_{p}\) at the trajectory point or a contour passing through the trajectory point from which \(\hat {\mathbf {n}}_{p}\) can be computed. This interpolation should be of sufficiently high order to prevent potential biases that may occur. One can also use a predictor-corrector scheme to compute \(\mathbf {X}^{\star }_{P}(\mathbf {y}_{\mathbf {s}},t-{\Delta } t)\) using the normals both at t−Δt and t.

As discussed in Section 4.1 and the Uniqueness remark of Section 3.3, multiple optimal paths exist to end points y f which lie on shock lines. However, as Eq. 9 is solved numerically, y f does not lie on shock lines exactly due to discretization errors. In fact, y f does not even lie exactly on the final level set contour \(\partial \mathcal {R}_{\phi ^{o}}(t)\), as mentioned above. Consequently, solving (12) in such cases yields only one of the optimal trajectories, depending on the numerical errors.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Lolla, T., Lermusiaux, P.F.J., Ueckermann, M.P. et al. Time-optimal path planning in dynamic flows using level set equations: theory and schemes. Ocean Dynamics 64, 1373–1397 (2014). https://doi.org/10.1007/s10236-014-0757-y

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s10236-014-0757-y

Keywords

Navigation