Skip to main content
Log in

Method of evolving junction on optimal path planning in flows fields

  • Published:
Autonomous Robots Aims and scope Submit manuscript

Abstract

We propose an algorithm using method of evolving junctions to solve the optimal path planning problems with piece-wise constant flow fields. In such flow fields, we prove that the optimal trajectories, with respect to a convex Lagrangian in the objective function, must be formed by piece-wise constant velocity motions. Taking advantage of this property, we transform the infinite dimensional optimal control problem into a finite dimensional optimization and use intermittent diffusion to solve the problems. The algorithm is proven to be complete. At last, we demonstrate the performance of the algorithm with various simulation examples.

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

Access this article

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

Instant access to the full article PDF.

Fig. 1
Fig. 2
Fig. 3
Fig. 4
Fig. 5
Fig. 6
Fig. 7
Fig. 8

Similar content being viewed by others

References

  • Chen, Y., He, Z., & Li, S. (2019). Horizon-based lazy optimal RRT for fast, efficient replanning in dynamic environment. Autonomous Robots, 43(8), 2271–2292.

    Article  Google Scholar 

  • Chow, S.N., Yang, T.S., & Zhou, H.M. (2013). Global optimizations by intermittent diffusion. In: Chaos, CNN, Memristors and Beyond: A Festschrift for Leon Chua With DVD-ROM, composed by Eleonora Bilotta. World Scientific, pp. 466–479.

  • Forrest, J., Hirst, J., & Tomlin, J. A. (1974). Practical solution of large mixed integer programming problems with umpire. Management Science, 20(5), 736–773.

    Article  MathSciNet  MATH  Google Scholar 

  • Gammell, J. D., Barfoot, T. D., & Srinivasa, S. S. (2018). Informed sampling for asymptotically optimal path planning. IEEE Transactions on Robotics, 34(4), 966–984.

    Article  Google Scholar 

  • Hou, M., Cho, S., & Zhou, H., et al. (2021). Bounded cost path planning for underwater vehicles assisted by a time-invariant partitioned flow field model. Frontiers in Robotics and AI , 8.

  • Hou, M., Zhai, H., & Zhou, H., et al. (2019). Partitioning ocean flow field for underwater vehicle path planning. In: OCEANS 2019-Marseille, IEEE, pp. 1–8.

  • Janson, L., Schmerling, E., Clark, A., et al. (2015). Fast marching tree: A fast marching sampling-based method for optimal motion planning in many dimensions. The International Journal of Robotics Research, 34(7), 883–921.

    Article  Google Scholar 

  • Ji, D. H., Choi, H. S., Kang, J. I., et al. (2019). Design and control of hybrid underwater glider. Advances in Mechanical Engineering, 11(5), 1687814019848.

    Article  Google Scholar 

  • Kaiser, E., Noack, B. R., Cordier, L., et al. (2014). Cluster-based reduced-order modelling of a mixing layer. Journal of Fluid Mechanics, 754, 365–414.

    Article  MATH  Google Scholar 

  • Karaman, S., & Frazzoli, E. (2011). Sampling-based algorithms for optimal motion planning. The International Journal of Robotics Research, 30(7), 846–894.

    Article  MATH  Google Scholar 

  • Kuffner, J., & LaValle, S. (2000). RRT-connect: An efficient approach to single-query path planning. In: IEEE International Conference on Robotics and Automation.

  • Kularatne, D., Bhattacharya, S., & Hsieh, M. A. (2017). Optimal path planning in time-varying flows using adaptive discretization. IEEE Robotics and Automation Letters, 3(1), 458–465.

    Article  Google Scholar 

  • Kularatne, D., Bhattacharya, S., & Hsieh, M. A. (2018). Going with the flow: A graph based approach to optimal path planning in general flows. Autonomous Robots, 42(7), 1369–1387.

    Article  Google Scholar 

  • LaValle, S. M. (1998). Rapidly-exploring random trees: A new tool for path planning. Tech. rep.: Department of Computer Science, Iowa State University.

    Google Scholar 

  • Leonard, N. E., Paley, D. A., Davis, R. E., et al. (2010). Coordinated control of an underwater glider fleet in an adaptive ocean sampling field experiment in Monterey Bay. Journal of Field Robotics, 27(6), 718–740. https://doi.org/10.1002/rob.20366.

    Article  Google Scholar 

  • Li, W., Lu, J., Zhou, H., et al. (2017). Method of evolving junctions: A new approach to optimal control with constraints. Automatica, 78, 72–78.

    Article  MathSciNet  MATH  Google Scholar 

  • Lolla, S.V.T. (2016). Path planning and adaptive sampling in the coastal ocean. PhD thesis, Massachusetts Institute of Technology,.

  • Martin, P.J. (2000). Description of the navy coastal ocean model version 1.0. , Tech. Rep. NRL/FR/7322–00-9962, Naval Research Lab.

  • Noreen, I., Khan, A., & Habib, Z. (2016). Optimal path planning using RRT* based approaches: a survey and future directions. International Journal of Advanced Computer Science and Applications, 7(11), 97–107.

    Article  Google Scholar 

  • Ozog, P., Carlevaris-Bianco, N., Kim, A. Y., et al. (2016). Long-term mapping techniques for ship hull inspection and surveillance using an autonomous underwater vehicle. Journal of Field Robotics, 33, 265–289.

    Article  Google Scholar 

  • Pereira, A. A., Binney, J., Hollinger, G. A., et al. (2013). Risk-aware path planning for autonomous underwater vehicles using predictive ocean models. Journal of Field Robotics, 30(5), 741–762. https://doi.org/10.1002/rob.21472.

    Article  Google Scholar 

  • Poole, D. L., & Mackworth, A. K. (2010). Artificial Intelligence: foundations of computational agents. Cambridge University Press.

  • Rhoads, B., Mezic, I., & Poje, A. C. (2012). Minimum time heading control of underpowered vehicles in time-varying ocean currents. Ocean Engineering, 66(1), 12–31.

    Google Scholar 

  • Russell, S., & Norvig, P. (2002). Artificial intelligence: a modern approach.

  • Ser-Giacomi, E., Rossi, V., López, C., et al. (2015). Flow networks: A characterization of geophysical fluid transport. Chaos: An Interdisciplinary Journal of Nonlinear Science, 25(3), 036404.

    Article  Google Scholar 

  • Sethian, J. A. (1999). Level set methods and fast marching methods: Evolving interfaces in geometry, fluid mechanics. Computer Vison and Material Science: Cambridge University Press.

    MATH  Google Scholar 

  • Shi, C., Wei, B., Wei, S., et al. (2021). (2021) A quantitative discriminant method of elbow point for the optimal number of clusters in clustering algorithm. EURASIP Journal on Wireless Communications and Networking, 1, 1–16.

    Google Scholar 

  • Shome, R., Solovey, K., Dobson, A., et al. (2020). dRRT*: Scalable and informed asymptotically-optimal multi-robot motion planning. Autonomous Robots, 44(3), 443–467.

    Article  Google Scholar 

  • Sinha, A., Malo, P., & Deb, K. (2017). A review on bilevel optimization: from classical to evolutionary approaches and applications. IEEE Transactions on Evolutionary Computation, 22(2), 276–295.

    Article  Google Scholar 

  • Smith, R. N., Chao, Y., Li, P. P., et al. (2010). Planning and implementing trajectories for autonomous underwater vehicles to track evolving ocean processes based on predictions from a Regional Ocean Model. The International Journal of Robotics Research, 29(12), 1475–1497.

    Article  Google Scholar 

  • Soulignac, M. (2011). Feasible and optimal path planning in strong current fields. IEEE Transactions on Robotics, 27(1), 89–98. https://doi.org/10.1109/tro.2010.2085790.

    Article  Google Scholar 

Download references

Acknowledgements

The authors would like to thank the support from NSF Grants DMS-1830225, ONR Grant N00014-21-1-2891, ONR Grants N00014-19-1-2556 and N00014-19-1-2266; AFOSR Grant FA9550-19-1-0283; NSF Grants CNS-1828678, S &AS-1849228 and GCR-1934836; NRL Grants N00173-17-1-G001 and N00173-19-P-1412 ; and NOAA Grant NA16NOS0120028.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Haomin Zhou.

Additional information

Publisher's Note

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

Appendix

Appendix

In this appendix, we give proofs for the properties related to Algorithm 1.

First we present proof of Theorem 4.1. The proof leverages the following theorem in Chow et al. (2013).

Theorem 6.1

Let Q be the set of global minimizers, U be a small neighborhood of Q and \(\lambda _{opt}\) the optimal solution obtained by the ID process. Then for any given \(\epsilon >0\), there exists \(\tau >0\), \(\sigma _0>0\) and \(N_0>0\) such that if \(T_i-S_i>\tau \), \(\sigma _i<\sigma _0\) (for \(i=1,\cdots ,N\)) and \(N>N_0\),

$$\begin{aligned} \mathbb {P}(\lambda _{opt}\in U)\ge 1-\epsilon . \end{aligned}$$

Then we provide completeness proof of Algorithm 1.

Proof of Theorem 4.1

The proof includes two steps. First, we show that the decision tree returns all cell sequences with total cost less than or equal to the lowest-cost path found so far. Then we prove that given a fixed cell sequence, the global minimizer can be found by Algorithm 2.

The DFS algorithm, which avoids repeated states in the graph, is complete in finite state spaces (Russell and Norvig, 2002). In a static flow field divided into convex regions, the optimal path will not visit a cell boundary curve more than one time. Hence, the optimal path connecting the root and the target node in the decision tree does not contain loops. Therefore, the BnBDFS returns all cell sequences with total cost less than or equal to the lowest-cost found so far.

Next we show that the ID algorithm is complete. We combine Theorem 4.4, together with Bellman principle, to show that the global optimal path must be in the structure of constant motion within each flow region. To prove that the proposed algorithm is convergent, we only need to show that there exists a global minimizer \(\lambda ^*=(\lambda _1^*,\cdots ,\lambda _K^*)\), around which there is a closed neighborhood \(U\subset \prod _{i=1}^KD_i\) such that \(vol(U)>0\) (vol is the product Lebesgue measure in \(\prod _{i=1}^KD_i\)) and for all \(\lambda \in U\), the gradient flow \(\dot{\lambda }=-\nabla J(\lambda )\) converges to \(\lambda ^*\). If this condition holds, we can follow the proof of intermittent diffusion and get the desired results.

To this end, if there exists such U that \(vol(U)>0\) and for all \(\lambda \in U\), we have \(J(\lambda )\le J(\mu )\) for arbitrary \(\mu \in S\) for some \(S\subset U\), then the proof is done. Now if the global minimizers are isolated, then given any global minimizer \(\lambda ^*=(\lambda _1^*,\cdots ,\lambda _K^*)\), since J is continuous differentiable, we can have a closed neighborhood \(U\subset \prod _{i=1}^KD_i\) with \(vol(U)>0\) (within the neighborhood, the dimension of the domain does not change) such that \(J(\lambda )>J(\lambda ^*)\) and \(\nabla J(\lambda )\ne 0\) for all \(\lambda \in U\backslash \{\lambda ^*\}\), then the gradient flow starting at \(\lambda \in U\) converges to \(\lambda ^*\).

Therefore, we prove the algorithm is complete. \(\square \)

Proof of Lemma 4.2

We write the value function as

$$\begin{aligned} \psi (x,t)=\frac{\Vert x-x_0\Vert }{\Vert v+u\Vert }. \end{aligned}$$

To make the problem complete, we define \(\psi (x,t)=+\infty \) if the vehicle cannot reach x in time t, which gives the final value function to be

$$\begin{aligned} \psi (x,t)=\left\{ \begin{array}{ll}\frac{\Vert x-x_0\Vert }{\Vert v+u\Vert }&{}\frac{\Vert x-x_0\Vert }{\Vert v+u\Vert }\le t\\ +\infty &{}\text {otherwise}\end{array}\right. . \end{aligned}$$

If only the reachable part is considered, from the above equation, we can calculate \(\psi _t=0\) and

$$\begin{aligned} \begin{aligned} \nabla \psi =\frac{1}{\Vert v+u\Vert ^2}\Big (&\Vert v+u\Vert \frac{x-x_0}{\Vert x-x_0\Vert }\\ {}&-\Vert x-x_0\Vert \nabla \Vert v+u\Vert \Big ). \end{aligned} \end{aligned}$$
(21)

We can rewrite \(v=v^0+v^\perp \) and \(V^2=\Vert v^0\Vert ^2+\Vert v^\perp \Vert ^2\) if we denote

$$\begin{aligned} v^0= & {} \frac{(x-x_0)}{\Vert x-x_0\Vert }\frac{(x-x_0)^Tv}{\Vert x-x_0\Vert },\\ v^\perp= & {} \left( I-\frac{(x-x_0)}{\Vert x-x_0\Vert }\frac{(x-x_0)^T}{\Vert x-x_0\Vert }\right) v, \end{aligned}$$

where I is the identity matrix. And u can be decomposed in the same manner \(u=u_0+u^\perp \). It is easy to see that \(v^\perp =-u^\perp \) since \((v+u)/\Vert v+u\Vert =(x-x_0)/\Vert x-x_0\Vert \). Then, we see that \(\Vert v+u\Vert =\Vert v_0\Vert +\Vert u_0\Vert \) and

$$\begin{aligned} \begin{aligned}&\sqrt{\left( \frac{(x-x_0)^Tu}{\Vert x-x_0\Vert }\right) ^2+V^2-\Vert u\Vert ^2}\\ =&\Big (\Vert u^0\Vert ^2+\Vert v^0\Vert ^2+\Vert v^\perp \Vert ^2-\Vert u^0\Vert ^2-\Vert u^\perp \Vert ^2\Big )^{1/2}\\ =&\Vert v^0\Vert . \end{aligned} \end{aligned}$$

Hence, we have

$$\begin{aligned}&\nabla \Vert v+u\Vert \\ =&\nabla \left( \frac{(x-x_0)^Tu}{\Vert x-x_0\Vert }+\sqrt{\left( \frac{(x-x_0)^Tu}{\Vert x-x_0\Vert }\right) ^2+V^2-\Vert u\Vert ^2}\right) \\ =&\frac{\Vert v+u\Vert }{\Vert v^0\Vert }\nabla \frac{(x-x_0)^Tu}{\Vert x-x_0\Vert }\\ =&\frac{\Vert v+u\Vert }{\Vert x-x_0\Vert \Vert v^0\Vert }\left( I-\frac{(x-x_0)}{\Vert x-x_0\Vert }\frac{(x-x_0)^T}{\Vert x-x_0\Vert }\right) u\\ =&\frac{\Vert v+u\Vert }{\Vert x-x_0\Vert \Vert v^0\Vert }u^\perp . \end{aligned}$$

Taking \(\nabla \Vert v+u\Vert \) back to (21) and noticing that \(v^\perp =-u^\perp \), we reduce the gradient to be

$$\begin{aligned} \nabla \psi= & {} \left( \frac{x-x_0}{\Vert x-x_0\Vert }-\frac{u^\perp }{\Vert v^0\Vert }\right) \frac{1}{\Vert v+u\Vert }=\frac{v}{\Vert v^0\Vert \Vert v+u\Vert }. \end{aligned}$$

With the above equation, the Hamiltonian is

$$\begin{aligned} H&=\sup _{\hat{v}:\Vert \hat{v}\Vert \le V}\left( \nabla \psi ^T(\hat{v}+u)-1\right) \\&=\sup _{\hat{v}:\Vert \hat{v}\Vert \le V}\left\{ \frac{v^T}{\Vert v^0\Vert }\frac{\hat{v}+u}{\Vert v+u\Vert }\right\} -1\\&=\frac{1}{\Vert v+u\Vert }\left( \sup _{\hat{v}:\Vert \hat{v}\Vert \le V}\left\{ \frac{v^T\hat{v}}{\Vert v^0\Vert }\right\} +\Vert u_0\Vert -\frac{\Vert u^\perp \Vert ^2}{\Vert v^0\Vert }\right) -1\\&=\frac{1}{\Vert v+u\Vert }\left( \frac{V^2}{\Vert v^0\Vert }+\Vert u^0\Vert -\frac{\Vert u^\perp \Vert ^2}{\Vert v^0\Vert }\right) -1\\&=\frac{1}{\Vert v+u\Vert }(\Vert v^0\Vert +\Vert u^0\Vert )-1=0, \end{aligned}$$

which leads to the conclusion that the value function induced by the maximum speed constant velocity motion solves the Hamilton-Jacobi equation, thus is the optimal moving pattern in a constant flow speed region since \(\min _t\psi =\psi \). \(\square \)

Meanwhile, using the same notation and logic, we can give the proof of Proposition 3.1:

Proof of Proposition 3.1

First we show that the objective function is well-defined if there exists a feasible trajectory, and \(\Vert u_{c_i}\Vert \le V\). If \((x_i-x_{i-1})^T u_{c_i}\le 0\), unless \(V>\Vert u_{c_i}\Vert \), there does not exists a feasible path. Therefore,

$$\begin{aligned} \begin{aligned}&(x_i-x_{i-1})^Tu_{c_i}<\\&\sqrt{\left( (x_i-x_{i-1})^Tu_{c_i}\right) ^2+\Vert x_i-x_{i-1}\Vert ^2(V^2-\Vert u_{c_i}\Vert ^2)}. \end{aligned} \end{aligned}$$

Since \(\Vert u_{c_i}\Vert ^2-V^2<0\), we have \(t_i^*>0\). On the other hand, if \((x_i-x_{i-1})^Tu_{c_i}>0\), we can have two cases: \(V>\Vert u_{c_i}\Vert \), which shares the same conclusion as the first case, and \(V<\Vert u_{c_i}\Vert \). In the latter circumstance, since \(\Vert u_{c_i}\Vert ^2-V^2>0\) and

$$\begin{aligned} \begin{aligned}&(x_i-x_{i-1})^Tu_{c_i}>\\ {}&\sqrt{\left( (x_i-x_{i-1})^Tu_{c_i}\right) ^2+\Vert x_i-x_{i-1}\Vert ^2(V^2-\Vert u_{c_i}\Vert ^2)}, \end{aligned} \end{aligned}$$

it is still true that \(t_i^*>0\).

Meanwhile, When \((x_i-x_{i-1})^Tu_{c_i}>0\), \(t_i^*>0\) still holds if \(V=\Vert u_{c_i}\Vert \) and actually

$$\begin{aligned}&t_i^*=\lim _{V^2-\Vert u_{c_i}\Vert ^2\rightarrow 0}\frac{1}{\Vert u_{c_i}\Vert ^2-V^2}\Big ((x_i-x_{i-1})^Tu_{c_i}\\&-\sqrt{\left( (x_i-x_{i-1})^Tu_{c_i}\right) ^2+\Vert x_i-x_{i-1}\Vert ^2(V^2-\Vert u_{c_i}\Vert ^2)}\Big )\\&=\frac{\Vert x_i-x_{i-1}\Vert ^2}{2(x_i-x_{i-1})^Tu_{c_i}}>0. \end{aligned}$$

However, if \((x_i-x_{i-1})^Tu_{c_i}\le 0\), \(V=\Vert u_{c_i}\Vert \) becomes a singular point since there is no feasible path. Thus, in this case, we cannot formally solve the problem.

Since \(g^t_i(x_i,x_{i-1})=g^t_i(x_i-x_{i-1})\) and \(g^e_i(x_i,x_{i-1})=g^e_i(x_i-x_{i-1})\), we only need to consider the differentibility of

$$\begin{aligned} g(a)=\left\{ \begin{array}{ll} g_1(a) &{} \text {if 10 holds} \\ g_2(a) &{} \text {otherwise} \end{array}\right. \end{aligned}$$

where

$$\begin{aligned} g_1(a)&=2\sqrt{\Vert u_{c_i}\Vert ^2+C}\Vert a\Vert -2a^Tu_{c_i}\\ {}&=2\Vert a\Vert \left( \sqrt{\Vert u_{c_i}\Vert ^2+C}-\Vert u^0_{c_i}\Vert \right) ,\\ g_2(a)&=\frac{V^2+C}{\Vert u_{c_i}\Vert ^2-V^2}\Big (a^Tu_{c_i}\\ -&\sqrt{(a^Tu_{c_i})^2+\Vert a\Vert ^2(V^2-\Vert u_{c_i}\Vert ^2)}\Big )\\ {}&=\frac{(V^2+C)\Vert a\Vert }{\Vert u_{c_i}^0\Vert +\Vert v^0\Vert }. \end{aligned}$$

First of all, when equality in (10) holds, we have

$$\begin{aligned} \begin{aligned}&\sqrt{\Vert u_{c_i}\Vert ^2+C}=\Vert u_{c_i}^0\Vert \pm \Vert v^0\Vert \\ \Longrightarrow&\sqrt{\Vert u_{c_i}\Vert ^2+C}=\Vert u_{c_i}^0\Vert +\Vert v^0\Vert . \end{aligned} \end{aligned}$$
(22)

We take the plus sign since \(\sqrt{\Vert u_{c_i}\Vert ^2+C}\ge \Vert u_{c_i}\Vert \). Meanwhile from (10) and (22), we can derive the following equation

$$\begin{aligned} V^2+C=2\Vert v^0\Vert (\Vert u_{c_i}^0\Vert +\Vert v^0\Vert ). \end{aligned}$$

Therefore, g(a) is continuous. Similar calculations shows that

$$\begin{aligned} \nabla g_1&=2\left( (\sqrt{\Vert u_{c_i}\Vert ^2+C})\frac{a}{\Vert a\Vert }-u_{c_i}\right) =v^0-u_{c_i}^\perp =2v,\\ \nabla g_2&=\frac{V^2+C}{\Vert v^0\Vert \Vert v+u_{c_i}\Vert }v=2v, \end{aligned}$$

which gives us the desired result. \(\square \)

Proof of Lemma 4.3

In the case of minimum energy planning, \(L(x,v)=\Vert v\Vert ^2+C\) where \(C\ge 0\) is a constant running cost. To calculate the optimal solution for the vehicle running from \(x_0\) to the target x in a constant flow velocity field, we again study the constant speed straight line motion. However in this circumstance, the vehicle may no longer travel with maximum speed, hence we take the travel time in the region into consideration. Suppose that the the vehicle moves from \(x_0\) to x in time t, we set the vehicle velocity to be

$$\begin{aligned} v=\frac{x-x_0}{t}-u, \end{aligned}$$

assuming that

$$\begin{aligned} \Vert v\Vert ^2=\frac{\Vert x-x_0\Vert ^2}{t^2}+\Vert u\Vert ^2-\frac{2(x-x_0)^Tu}{t}\le V^2. \end{aligned}$$
(23)

Then the value function is

$$\begin{aligned} \begin{aligned} \psi (x,t)&=(\Vert v\Vert ^2+C)t\\ {}&=\frac{\Vert x-x_0\Vert ^2}{t}-2(x-x_0)^Tu+(C+\Vert u\Vert ^2)t. \end{aligned} \end{aligned}$$

Further we take

$$\begin{aligned} \psi (x,t)=\left\{ \begin{array}{ll} \begin{aligned} &{}\frac{\Vert x-x_0\Vert ^2}{t}\\ {} &{}-2(x-x_0)^Tu+(C+\Vert u\Vert ^2)t \end{aligned} &{} \Vert v\Vert \le V \\ +\infty &{} \text {otherwise} \end{array}\right. . \end{aligned}$$

Then by direct calculation with the finite part of \(\psi \), we have

$$\begin{aligned} \psi _t&=C+\Vert u\Vert ^2-\frac{\Vert x-x_0\Vert ^2}{t^2}, \end{aligned}$$
(24)
$$\begin{aligned} \nabla \psi&=\frac{2(x-x_0)^Tu}{t}-2u, \end{aligned}$$
(25)
$$\begin{aligned} \Vert \nabla \psi \Vert ^2&=\frac{4\Vert x-x_0\Vert ^2}{t^2}+4\Vert u\Vert ^2-\frac{8(x-x_0)^Tu}{t} \end{aligned}$$
(26)

The Hamilton-Jacobi equation is in the form of

$$\begin{aligned} \psi _t+\sup _{v:\Vert v\Vert \le V}\left\{ \nabla \psi ^T(v+u)-\Vert v\Vert ^2-C\right\} =0. \end{aligned}$$
(27)

To solve the optimization part of (27), we denote

$$\begin{aligned} F(v)=\nabla \psi ^T(v+u)-\Vert v\Vert ^2-C \end{aligned}$$

and calculate its critical point as

$$\begin{aligned} v^*=\frac{1}{2}\nabla \psi , \end{aligned}$$

which means that the optimal is

$$\begin{aligned} H=\sup _{v:\Vert v\Vert \le V}F(v)=\frac{1}{4}\Vert \nabla \psi \Vert ^2+\nabla \psi ^Tu-C \end{aligned}$$
(28)

and by (25), we have

$$\begin{aligned} \begin{aligned} \Vert v^*\Vert ^2&=\frac{1}{4}\Vert \nabla \psi \Vert ^2=\frac{\Vert x-x_0\Vert ^2}{t^2}+\Vert u\Vert ^2-\frac{2(x-x_0)^Tu}{t} \\&\le V^2, \end{aligned} \end{aligned}$$
(29)

which leads to the fact that \(F(v^*)=\sup _{v:\Vert v\Vert \le V}F(v)\). Let us take (25),(26) into (28) and the result is

$$\begin{aligned} H=\frac{\Vert x-x_0\Vert ^2}{t^2}-\Vert u\Vert ^2-C. \end{aligned}$$
(30)

Combining (24) and (30) finally results in the constructed \(\psi \) being the solution of (27).

Based on the solution \(\psi \), we further find the minimizer over time t and solve the minimization problem as follow:

$$\begin{aligned} \begin{aligned}&\min _{t\ge 0}\psi =\min _{t\ge 0}(\Vert v\Vert ^2+C)t\\=&\frac{\Vert x-x_0\Vert ^2}{t}-2(x-x_0)^Tu+(C+\Vert u\Vert ^2)t. \end{aligned} \end{aligned}$$

It is easy to see that the global minimizer of \(\psi \) over t is

$$\begin{aligned} t^*=\frac{\Vert x-x_0\Vert }{\sqrt{C+\Vert u\Vert ^2}} \end{aligned}$$

and the corresponding minimum is

$$\begin{aligned} \psi ^*=2\Vert x-x_0\Vert \sqrt{C+\Vert u\Vert ^2}-2(x-x_0)^Tu. \end{aligned}$$
(31)

Thus, if \(t^*\) is reachable, that is, using (23), we have

$$\begin{aligned} \Vert v(t^*)\Vert ^2=C+2\Vert u\Vert ^2-\sqrt{C+\Vert u\Vert ^2}\frac{2(x-x_0)^Tu}{\Vert x-x_0\Vert }\le V^2 \end{aligned}$$

the optimal is given as (31).

On the other hand, if \(\Vert v(t^*)\Vert >V\), the global minimizer \(t^*\) is on longer in the domain of our problem. In this case, we notice that \(\Vert v\Vert ^2\) is decreasing on the interval

$$\begin{aligned} \left[ t^*,\frac{(x-x_0)^Tu}{\Vert x-x_0\Vert ^2}\right] , \end{aligned}$$

and is increasing on

$$\begin{aligned} \left[ \frac{(x-x_0)^Tu}{\Vert x-x_0\Vert ^2},+\infty \right) . \end{aligned}$$

Also by noticing that \(\lim _{t\rightarrow +\infty }\Vert v\Vert ^2=\Vert u\Vert ^2\le V^2\), we conclude that there exists \(t_0>t^*\) when \(t\ge t_0>t^*\), \(\Vert v\Vert \le V\). Meanwhile, when \(t>t^*\), \(\psi \) is monotone increasing with respect to t. Hence, to get the minimum, we should take the time \(t=t_0\), where \(\Vert v(t_0)\Vert =V\). By taking the equality in (23), we have then

$$\begin{aligned} (\Vert u\Vert ^2-V^2)t^2-2(x-x_0)^Tut+\Vert x-x_0\Vert ^2=0, \end{aligned}$$

from which we have

$$\begin{aligned} \begin{aligned} t_0=&\frac{1}{\Vert u\Vert ^2-V^2}\Big ((x-x_0)^Tu\\ {}&-\sqrt{\left( (x-x_0)^Tu\right) ^2+\Vert x-x_0\Vert ^2(V^2-\Vert u\Vert ^2)}\Big ), \end{aligned} \end{aligned}$$

and \(\min _{t}\psi =(V^2+C)t_0\). \(\square \)

Proof of Theorem 3.1

Denoting g(w) to be the inverse of f such that if \(w=f(u+v)\) then \(u+v=g(w)\), we will show that

$$\begin{aligned} \psi (x,t)=\left\{ \begin{array}{ll} tL\left( g\left( \frac{x-x_0}{t}\right) -u\right) &{}\Vert g\left( \frac{x-x_0}{t}\right) -u\Vert \le V\\ +\infty &{}\text {otherwise} \end{array}\right. \end{aligned}$$

satisfies the HJB equation (20). First of all, the Hessian matrix \(\mathcal {H}(v)\) is positive definite for all \(\Vert v\Vert \le V\) since L is convex. Therefore, for any \(v_1,v_2\) in the domain, there exists \(\xi \) such that

$$\begin{aligned} \nabla _vL(v_1)=\nabla _vL(v_2)+\mathcal {H}(\xi )(v_2-v_1). \end{aligned}$$

Further if \(\nabla _vL(v_1)=\nabla _vL(v_2)\), then \(\mathcal {H}(\xi )(v_2-v_1)=0\). Because of the positive definite property for \(\mathcal {H}\), we have \(v_2=v_1\), which implies that \(\nabla _vL(v)\) is one-to-one.

Then we do the following calculation on the non-infinity part of \(\psi \)

$$\begin{aligned}&\begin{aligned}&\psi _t=L\left( g\left( \frac{x-x_0}{t}\right) -u\right) \\ {}&-\left[ \nabla _vL\left( g\left( \frac{x-x_0}{t}\right) -u\right) \right] ^T\nabla _wg\left( \frac{x-x_0}{t}\right) \frac{x-x_0}{t}, \end{aligned} \end{aligned}$$
(32)
$$\begin{aligned}&\nabla \psi =\left[ \nabla _wg\left( \frac{x-x_0}{t}\right) \right] ^T\nabla _vL\left( g\left( \frac{x-x_0}{t}\right) -u\right) . \end{aligned}$$
(33)

Since L is convex, we further have the relaxed optimization

$$\begin{aligned} \max _v\{\nabla \psi ^T(v+u)-L(v)\} \end{aligned}$$

is a convex problem and get the condition for the optimal \(v^*\) to be

$$\begin{aligned} \begin{aligned}&\left[ \nabla _vf(u+v^*)\right] ^T\nabla \psi =\nabla _vL(v^*)\\ \Longrightarrow&\nabla \psi =\left[ \nabla _wg(f(u+v^*))\right] ^T\nabla _vL(v^*). \end{aligned} \end{aligned}$$

Combining this with (33), we have

$$\begin{aligned} v^*=g\left( \frac{x-x_0}{t}\right) -u, \end{aligned}$$
(34)

and \(\Vert v^*\Vert \le V\) holds. Thus, \(v^*\) is the maximizer of \(H(x,\nabla \psi )\). Taking (32), (33) and (34), we have

$$\begin{aligned} \psi _t+\nabla \psi ^T(v^*+u)-L(v^*)=0, \end{aligned}$$

implying that (20) holds. At last notice that

$$\begin{aligned} \lim _{t\rightarrow \infty }L\left( g\left( \frac{x-x_0}{t}\right) -u\right) =L(g(0)-u)<\infty , \end{aligned}$$

since 0 is in the range of f. We have that

$$\begin{aligned} \lim _{t\rightarrow \infty }tL\left( g\left( \frac{x-x_0}{t}\right) -u\right) =\infty . \end{aligned}$$

Thus, we have \(t^*>0\) such that given x,

$$\begin{aligned} t^*={{\,\mathrm{arg\,min}\,}}_{t\ge 0}\psi (x,t). \end{aligned}$$

Thus, \(v^*=g((x-x_0)/t^*)-u\) gives us a constant velocity motion. \(\square \)

Rights and permissions

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

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Zhai, H., Hou, M., Zhang, F. et al. Method of evolving junction on optimal path planning in flows fields. Auton Robot 46, 929–947 (2022). https://doi.org/10.1007/s10514-022-10058-5

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s10514-022-10058-5

Keywords

Navigation