Skip to main content
Log in

Model predictive control for trajectory-tracking and formation of wheeled mobile robots

  • S.I.: NCAA 2021
  • Published:
Neural Computing and Applications Aims and scope Submit manuscript

Abstract

Model predictive control (MPC) naturally guarantees optimal transient process and constraints satisfaction. Most mature MPC theories concern with linear time-invariant systems, it is not trivial to develop MPC for time-varying or nonlinear systems. A linear time-varying model predictive control strategy is proposed in this paper for single-wheeled mobile robot trajectory tracking subject to the non-holonomic constraint and control constraints. The kinematic equation of the robot is converted into a linear time-varying form after linearizing and discretizing, the time-varying MPC is capable to be applied in consequence. The proposed trajectory tracking task is extended to the formation control among multiple robots solved by means of nonlinear model predictive control directly. The extended formation MPC is in the leader-following framework. Recursive feasibility and stability are proved for the closed-loop system.

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

Access this article

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

Instant access to the full article PDF.

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

Similar content being viewed by others

References

  1. Aguiar AP, Hespanha JP, Kokotovic PV (2005) Path-following for nonminimum phase systems removes performance limitations. IEEE Trans Autom Control 50(2):234–239

    Article  MathSciNet  MATH  Google Scholar 

  2. Yang Q, Sun Z, Cao M et al (2019) Stress-matrix-based formation scaling control. Automatica 101(2):120–127

    Article  MathSciNet  MATH  Google Scholar 

  3. Kanjanawaniskul KF (2012) Motion control of a wheeled mobile robot using model predictive control: a survey. Asia Pac J Sci Technol 17(5):811–837

    Google Scholar 

  4. Kolmanovsky I, Mcclamroch NH (1995) Developments in nonholonomic control problems. IEEE Control Syst 15(6):20–36

    Article  Google Scholar 

  5. Brockett R (1983) Asymptotic stability and feedback stabilization in differential geometric control theory. Progress Math 27:181–208

    Google Scholar 

  6. Godhavn JM, Egeland O (1997) A Lyapunov approach to exponential stabilization of nonholonomic systems in power form. IEEE Trans Autom Control 42(7):1028–1032

    Article  MathSciNet  MATH  Google Scholar 

  7. Sontag ED (2011) Stability and feedback stabilization. Math Complex Dyn Syst 1639–1652

  8. Panagou D, Kyriakopoulos KJ (2014) Dynamic positioning for an underactuated marine vehicle using hybrid control. Int J Control 87(2):264–280

    Article  MathSciNet  MATH  Google Scholar 

  9. Chwa DK, Seo JH, Kim P et al (2004) Sliding-mode tracking control of nonholonomic wheeled mobile robots in polar coordinates. IEEE Trans Control Syst Technol 12(4):637–644

    Article  Google Scholar 

  10. Hou ZG, Zou AM, Cheng L et al (2009) Adaptive control of an electrically driven nonholonomic mobile robot via backstepping and fuzzy approach. IEEE Trans Control Syst Technol 17(4):803–815

    Article  Google Scholar 

  11. Michaek M, Kozowski K (2009) Vector-field-orientation feedback control method for a differentially driven vehicle. IEEE Trans Control Syst Technol 18(1):45–65

    Article  Google Scholar 

  12. Xu D, Zhang X, Zhu Z et al (2014) Behavior-based formation control of swarm robots. Math Probl Eng 2014:1–13

    MATH  Google Scholar 

  13. Wang PK (1991) Navigation strategies for multiple autonomous mobile robots moving in formation. J Robot Syst 8(2):177–195

    Article  MATH  Google Scholar 

  14. Gao XN, Wu LJ (2014) Multi-robot formation control based on the artificial potential field method. Appl Mech Mater 519:1360–1363

    Article  Google Scholar 

  15. Desai JP, Ostrowski JP, Kumar V (2001) Modeling and control of formations of nonholonomic mobile robots. IEEE Trans Robot Autom 17(6):905–908

    Article  Google Scholar 

  16. Indiveri G, Paulus J, Plöger PG (2006) Motion control of Swedish wheeled mobile robots in the presence of actuator saturation. In: Robot soccer world cup Conference 2006. Springer, Heidelberg, pp 35–46

  17. Falcone P, Borrelli F, Asgari J et al (2007) Predictive active steering control for autonomous vehicle systems. IEEE Trans Control Syst Technol 15(3):566–580

    Article  Google Scholar 

  18. Abdolhosseini M (2012) Model predictive control of an unmanned quadrotor helicopter: theory and flight tests. Doctoral dissertation, Concordia University

  19. Van PR, Pipeleers G (2017) Distributed MPC for multi-vehicle systems moving in formation. Robot Auton Syst 97:144–152

    Article  Google Scholar 

  20. Kuriki Y, Namerikawa T (2015) Formation control with collision avoidance for a multi-UAV system using decentralized MPC and consensus-based control. SICE J Control Meas Syst Integr 8(4):285–294

    Article  Google Scholar 

  21. Samad T (2017) A survey on industry impact and challenges thereof. IEEE Control Syst Mag 37(1):17–18

    Article  Google Scholar 

  22. Lages WF, Alves JAV (2006) Real-time control of a mobile robot using linearized model predictive control. IFAC Proc Vol 39(16):968–973

    Article  Google Scholar 

  23. Klancar G, Skrjanc I (2007) Tracking-error model-based predictive control for mobile robots in real time. Robot Auton Syst 55(6):460–469

    Article  Google Scholar 

  24. Allgower F, Findeisen R, Nagy ZK (2004) Nonlinear model predictive control: from theory to application. J Chin Inst Chem Eng 35(3):299–316

    Google Scholar 

  25. Mayne DQ, Rawlings JB (2000) Constrained model predictive control: stability and optimality. Automatica 36(6):789–814

    Article  MathSciNet  MATH  Google Scholar 

  26. Gu D, Hu H (2006) Receding horizon tracking control of wheeled mobile robots. IEEE Trans Control Syst Technol 14(4):743–749

    Article  Google Scholar 

  27. Yu S, Chen H, Zhang P et al (2008) An LMI optimization approach for enlarging the terminal region of MPC for nonlinear systems. Acta Autom Sin 34(7):798–804

    Article  MathSciNet  MATH  Google Scholar 

  28. Bleris LG, Vouzis PD, Arnold MG et al (2006) A co-processor FPGA platform for the implementation of real-time model predictive control. In: IEEE American control conference Conference 2006. Minneapolis, Minnesota, pp 1912–1917

  29. Nascimento TP, Moreira AP, Conceição A (2013) Multi-robot nonlinear model predictive formation control: moving target and target absence. Robot Auton Syst 61(12):1502–1515

    Article  Google Scholar 

  30. Gu D, Hu H (2009) A model predictive controller for robots to follow a virtual leader. Robotica 27(6):905–913

    Article  Google Scholar 

  31. Dunbar WB, Murray RM (2006) Distributed receding horizon control for multi-vehicle formation stabilization. Automatica 42(4):549–558

    Article  MathSciNet  MATH  Google Scholar 

  32. Matos ACDC (2011) Optimization and control of nonholonomic vehicles and vehicles formations. PhD thesis, University of Porto

  33. Mehrez MW, Mann GKI, Gosine RG (2014) Formation stabilization of nonholonomic robots using nonlinear model predictive control. In: IEEE 27th Canadian conference on electrical and computer engineering Conference 2014, Toronto, Canada, pp 1–6

  34. Guechi EH, Lauber J, DAmbrine M et al (2012) Output feedback controller design of a unicycle-type mobile robot with delayed measurements. IET Control Theory Appl 6(5):726–733

  35. Thuilot B, DAndrea-Novel B, Micaelli A (1996) Modeling and feedback control of mobile robots equipped with several steering wheels. IEEE Trans Robot Autom 12(3):375–390

    Article  Google Scholar 

Download references

Acknowledgements

This work was supported by National Natural Science Foundation of China (NSFC) under Grants 62073015 and 61703018.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Bing Zhu.

Ethics declarations

Conflict of interest

Authors declare no conflict of interest.

Additional information

Publisher's Note

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

Appendix

Appendix

1.1 Proof of Proposition 1

From Proposition 1, optimal problem has feasible solution (24) at sampling time interval k:

$$\begin{aligned} \overline{{\mathbf {U}}}(k)^{*}=\left\{ \tilde{{\mathbf {u}}}^{*}(k\mid k),\tilde{{\mathbf {u}}}^{*}(k+1\mid k),\ldots ,\tilde{{\mathbf {u}}}^{*}(k+N-1\mid k)\right\} . \end{aligned}$$

Corresponding predictive states are

$$\begin{aligned} \overline{{\mathbf {X}}}(k+1)^{*}=\left\{ \tilde{{\mathbf {x}}}^{*}(k+1 \mid k), \tilde{{\mathbf {x}}}^{*}(k+2 \mid k),\ldots ,\tilde{{\mathbf {x}}}^{*}(k+N-1 \mid k), 0\right\} \end{aligned}$$

The terminal constraint (22) results in the last element of predictive state series identically equal to zero.

Considering the MPC is carried on by (25). Then, it is noted that

$$\begin{aligned} \tilde{{\mathbf {x}}}(k+1)={\mathbf {F}}(k) \tilde{{\mathbf {x}}}(k)+{\mathbf {G}}(k) \tilde{{\mathbf {u}}}^{*}(k \mid k)=\tilde{{\mathbf {x}}}^*(k+1|k) \end{aligned}$$

When \(k+1\) comes, the optimization (24) has a feasible (not necessarily optimal) solution allocated by transferring the control sequence at time k:

$$\begin{aligned} \overline{{\mathbf {U}}}(k+1)= & {} \left[ \begin{array}{c}\tilde{{\mathbf {u}}}(k+1\mid k+1)\\ \tilde{{\mathbf {u}}}(k+2\mid k+1)\\ \vdots \\ \tilde{{\mathbf {u}}}(k+N-1\mid k+1)\\ \tilde{{\mathbf {u}}}(k+N\mid k+1)\end{array}\right] \\= & {} \left[ \begin{array}{c}\tilde{{\mathbf {u}}}^{*}(k+1\mid k)\\ \tilde{{\mathbf {u}}}^{*}(k+2\mid k)\\ \vdots \\ \tilde{{\mathbf {u}}}^{*}(k+N-1\mid k)\\ 0\end{array}\right] \end{aligned}$$
(60)

and the relevant predictive states are

$$\begin{aligned} \overline{{\mathbf {X}}}(k+2)= & {} \left[ \begin{array}{c}\tilde{{\mathbf {x}}}(k+2\mid k+1)\\ \tilde{{\mathbf {x}}}(k+3\mid k+1)\\ \vdots \\ \tilde{{\mathbf {x}}}(k+N\mid k+1)\\ \tilde{{\mathbf {x}}}(k+N+1\mid k+1)\end{array}\right] \\= & {} \left[ \begin{array}{c}\tilde{{\mathbf {x}}}^{*}(k+2\mid k)\\ \tilde{{\mathbf {x}}}^{*}(k+3\mid k)\\ \vdots \\ \tilde{{\mathbf {x}}}^{*}(k+N-1\mid k)\\ 0\\ 0\end{array}\right] \end{aligned}$$
(61)

mentioning that there exists at least one solution to (24) satisfied with control constraints (21) and the terminal constraints (22).It implies that, if there is an optimization (24) viable at initial time, then it will be viable at any time in the future, and this achieves the recursive feasibility of the system.

1.2 Proof of Proposition 2

Select the optimal objective function \(J^{*}(k)\) to be the Lyapunov candidate (time-dependent) V(k) under time k condition:

$$\begin{aligned} V(k)= & {} J^{*}(k) \\= & {} \sum _{i=0}^{N-1}\left( \left\| \tilde{{\mathbf {x}}}^{*}(k+i \mid k)\right\| _{Q}^{2}+\left\| \tilde{{\mathbf {u}}}^{*}(k+i \mid k)\right\| _{R}^{2}\right) . \end{aligned}$$
(62)

It is obvious to see \(V(0)=0\), \(V(k)>0\) as to \(\tilde{{\mathbf {x}}}(k)\ne 0\) and \(\tilde{{\mathbf {u}}}(k)\ne 0\). Moreover, V(k) is quadratic positive definite holding inequality:

$$\begin{aligned} V(k)\ge \Vert \tilde{{\mathbf {x}}}(k)\Vert _{Q}^2. \end{aligned}$$
(63)

Owing to the uniform boundness of coefficient matrix \({\mathbf {F}}(k)\), \({\mathbf {G}}(k)\) and the inputs \(\tilde{{\mathbf {u}}}\), it is easy to see from (19) that there exist a constant \(\alpha >0\) satisfying the inequality condition:

$$\begin{aligned} V(k)\le \alpha \Vert \tilde{{\mathbf {x}}}(k)\Vert _{Q}^2 \end{aligned}$$
(64)

indicating V(k) is decrescent.

From (60) and (61), \(J(k+1)\) (not necessarily optimal) at \(k+1\) will be acquired by

$$\begin{aligned} J(k+1) =J^{*}(k)-\left\| \tilde{{\mathbf {x}}}^{*}(k \mid k)\right\| _{Q}^{2}-\left\| \tilde{{\mathbf {u}}}^{*}(k \mid k)\right\| _{R}^{2} \end{aligned}$$

and

$$\begin{aligned} {V}(k+1)-V(k)&=J^{*}(k+1)-J^{*}(k) \le J(k+1)-J^{*}(k)\nonumber \\ &\le -\left\| \tilde{{\mathbf {x}}}^{*}(k \mid k)\right\| _{Q}^{2}-\left\| \tilde{{\mathbf {u}}}^{*}(k \mid k)\right\| _{R}^{2}\nonumber \\ &\le -\left\| \tilde{{\mathbf {x}}}^{*}(k \mid k)\right\| _{Q}^{2} \end{aligned}$$
(65)

implying the change of V(k) is negative.

It is concluded from (63), (64) and (65) that the linearized closed-loop system meets the condition of uniform asymptotic stability.

The nonlinear model of WMR can be deemed as a vanishing disturbance (vanishes at \(\tilde{{\mathbf {x}}}=0\)) brought to the linearized model, and it straightforwardly has the conclusion that around the reference trajectory, nonlinear tracking error closed-loop system has the locally uniform asymptotically stability.

1.3 Proof of Proposition 3

The optimal control and state feasible sequences (46) from Proposition 3 at time k are chosen by

$$\begin{aligned} \overline{{\mathbf {U}}}_{{Le}}^{*}(k)= & {} \left\{ {\mathbf {u}}_{{Le}}^{*}(k \mid k), {\mathbf {u}}_{{Le}}^{*}(k+1 \mid k), \ldots , {\mathbf {u}}_{{Le}}^{*}(k+N-1 \mid k)\right\} \nonumber \\ \overline{{\mathbf {U}}}_{f e}^{*}(k)= & {} \left\{ {\mathbf {u}}_{f e}^{*}(k \mid k), {\mathbf {u}}_{f e}^{*}(k+1 \mid k), \ldots , {\mathbf {u}}_{f e}^{*}(k+N-1 \mid k)\right\} \nonumber \\ \overline{{\mathbf {X}}}_{{Le}}^{*}(k)= & {} \left\{ {\mathbf {x}}_{{Le}}^{*}(k \mid k), {\mathbf {x}}_{{Le}}^{*}(k+1 \mid k), \ldots , {\mathbf {x}}_{{Le}}^{*}(k+N-1 \mid k), 0\right\} \nonumber \\ \overline{{\mathbf {X}}}_{f e}^{*}(k)= & {} \left\{ {\mathbf {x}}_{f e}^{*}(k \mid k), {\mathbf {x}}_{f e}^{*}(k+1 \mid k), \ldots , {\mathbf {x}}_{f e}^{*}(k+N-1 \mid k), 0\right\} . \end{aligned}$$

The predictive errors in next sample moment are calculated with (52), (53) by

$$\begin{aligned} {\mathbf {x}}_{{Le}}(k+1)= & {} {\mathbf {x}}_{L e}(k)+T{\mathbf {f}}\left( {\mathbf {x}}_{{Le}}(k), {\mathbf {u}}_{{Le}}^{*}(k)\right) \nonumber \\= & {} {\mathbf {x}}_{{Le}}^{*}(k+1 \mid k), \end{aligned}$$
(66)
$$\begin{aligned} {\mathbf {x}}_{f e}(k+1)= & {} {\mathbf {x}}_{f e}(k)+T{\mathbf {h}}\left( {\mathbf {x}}_{f e}(k), {\mathbf {u}}_{f e}^{*}(k)\right) \nonumber \\= & {} {\mathbf {x}}_{f e}^{*}(k+1 \mid k). \end{aligned}$$
(67)

The corresponding state and control sequences are

$$\begin{aligned} \overline{{\mathbf {U}}}_{{Le}}(k+1)= & {} \left[ \begin{array}{c} {\mathbf {u}}_{{Le}}(k+1 \mid k+1) \\ {\mathbf {u}}_{{Le}}(k+2 \mid k+1) \\ \vdots \\ {\mathbf {u}}_{{Le}}(k+N-1 \mid k+1) \\ {\mathbf {u}}_{{Le}}(k+N \mid k+1) \end{array}\right] \nonumber \\= & {} \left[ \begin{array}{c} {\mathbf {u}}_{{Le}}^{*}(k+1 \mid k) \\ {\mathbf {u}}_{{Le}}^{*}(k+2 \mid k) \\ \vdots \\ {\mathbf {u}}_{{Le}}^{*}(k+N-1 \mid k) \\ 0 \end{array}\right] \end{aligned}$$
(68)
$$\begin{aligned} \overline{{\mathbf {U}}}_{f e}(k+1)= & {} \left[ \begin{array}{c} {\mathbf {u}}_{f e}(k+1 \mid k+1) \\ {\mathbf {u}}_{f e}(k+2 \mid k+1) \\ \vdots \\ {\mathbf {u}}_{f e}(k+N-1 \mid k+1) \\ {\mathbf {u}}_{f e}(k+N \mid k+1) \end{array}\right] \nonumber \\= & {} \left[ \begin{array}{c} {\mathbf {u}}_{f e}^{*}(k+1 \mid k) \\ {\mathbf {u}}_{f e}^{*}(k+2 \mid k) \\ \vdots \\ {\mathbf {u}}_{f e}^{*}(k+N-1 \mid k) \\ 0 \end{array}\right] . \end{aligned}$$
(69)

The first \(N-1\) individuals of the feasible given inputs sequences are generated by transforming the optimization solution of time k. The reason of selecting zero as the last element is that predict states \({\mathbf {x}}_{{Le}}(k+N+1|k)\), \({\mathbf {x}}_{f e}(k+N+1|k)\) will be equal to zero under the influence of \({\mathbf {u}}_{{Le}}(k+N|k)=0\), \({\mathbf {u}}_{f e}(k+N|k)=0\).

The corresponding \(k+1\) time interval prediction state sequences are:

$$\begin{aligned} \overline{{\mathbf {X}}}_{{Le}}(k+1)= & {} \left[ \begin{array}{c} {\mathbf {x}}_{{Le}}(k+1 \mid k+1) \\ {\mathbf {x}}_{{Le}}(k+2 \mid k+1) \\ {\mathbf {x}}_{{Le}}(k+3 \mid k+1) \\ \vdots \\ {\mathbf {x}}_{{Le}}(k+N \mid k+1) \\ {\mathbf {x}}_{{Le}}(k+N+1 \mid k+1) \end{array}\right] \nonumber \\= & {} \left[ \begin{array}{c} {\mathbf {x}}_{{Le}}^{*}(k+1 \mid k) \\ {\mathbf {x}}_{{Le}}^{*}(k+2 \mid k) \\ {\mathbf {x}}_{{Le}}^{*}(k+3 \mid k) \\ \vdots \\ {\mathbf {x}}_{{Le}}^{*}(k+N-1 \mid k) \\ 0 \\ 0 \end{array}\right] \end{aligned}$$
(70)
$$\begin{aligned} \overline{{\mathbf {X}}}_{f e}(k+1)= & {} \left[ \begin{array}{c} {\mathbf {x}}_{f e}(k+1 \mid k+1) \\ {\mathbf {x}}_{f e}(k+2 \mid k+1) \\ {\mathbf {x}}_{f e}(k+3 \mid k+1) \\ \vdots \\ {\mathbf {x}}_{f e}(k+N \mid k+1) \\ {\mathbf {x}}_{f e}(k+N+1 \mid k+1) \end{array}\right] \nonumber \\= & {} \left[ \begin{array}{c} {\mathbf {x}}_{f e}^{*}(k+1 \mid k) \\ {\mathbf {x}}_{f e}^{*}(k+2 \mid k) \\ {\mathbf {x}}_{f e}^{*}(k+3 \mid k) \\ \vdots \\ {\mathbf {x}}_{f e}^{*}(k+N-1 \mid k) \\ 0 \\ 0 \end{array}\right] . \end{aligned}$$
(71)

The above derivations indicate that there exists at least one set of solutions satisfying all constraints, and the optimization (46) is feasible at all predictive steps.

1.4 Proof of Proposition 4

Due to the terminal equality constraint,

$$\begin{aligned}&{\mathbf {x}}_{{Le}}(k+N \mid k)=0 \\ &{\mathbf {x}}_{f e}(k+N \mid k)=0, \quad f=1,2, \ldots , m-1, \end{aligned}$$

the cost function J(k) becomes

$$\begin{aligned} J(k)= & {} \sum _{j=0}^{N-1}\left( \left\| {\mathbf {x}}_{{Le}}(k+i \mid k) \right\| _{Q_{L}}^{2}+\left\| {\mathbf {u}}_{{Le}}(k+i \mid k)\right\| _{R_{L}}^{2}\right) \nonumber \\ &+\sum _{f=1}^{m-1}\left[ \sum _{j=0}^{N-1}\left( \left\| {\mathbf {x} }_{f e}(k+i \mid k)\right\| _{Q_{f}}^{2}+\left\| {\mathbf {u}}_{f e}(k+i \mid k)\right\| _{R_{f}}^{2}\right) \right] . \end{aligned}$$
(72)

Select the candidate for Lyapunov function V(k) providing by optimal function \(J^{*}(k)\) at time k:

$$\begin{aligned} J^{*}(k)= & {} \sum _{j=0}^{N-1}\left( \left\| {\mathbf {x}}_{{Le}}^{*}(k+i \mid k)\right\| _{Q_{L}}^{2}+\left\| {\mathbf {u}}_{{Le}}^{*}(k+i \mid k) \right\| _{R_{L}}^{2}\right) \nonumber \\ &+\sum _{f=1}^{m-1}\left[ \sum _{j=0}^{N-1}\left( \left\| {\mathbf {x}}_{f e}^{*} (k+i \mid k)\right\| _{Q_{f}}^{2}+\left\| {\mathbf {u}}_{f e}^{*}(k+i \mid k) \right\| _{R_{f}}^{2}\right) \right] . \end{aligned}$$
(73)

The selected Lyapunov function V(k) is positive definite and decrescent. Substituting the predictive state and control sequences (68), (69), (70), and (71) into the objective function \(J(k+1)\) at time interval \(k+1\) yields

$$\begin{aligned} J(k+1)= & {} \sum _{i=0}^{N-1}\left( \left\| {\mathbf {x}}_{{Le}}(k+1+i \mid k+1)\right\| _{Q_{L}}^{2}\right. \\ &\left. +\left\| {\mathbf {u}}_{{Le}}(k+1+i \mid k+1)\right\| _{R_{L}}^{2}\right) \\ &+\sum _{f=1}^{m-1}\left[ \sum _{j=0}^{N-1}\left( \left\| {\mathbf {x}}_{f e}(k+1+i \mid k)\right\| _{{\mathcal {Q}}_{f}}^{2}\right. \right. \\ &\left. \left. +\left\| {\mathbf {u}}_{f e}(k+1+i \mid k)\right\| _{R_{f}}^{2}\right) \right] \\ =\, & {} J^{*}(k)-\left\| {\mathbf {x}}_{{Le}}^{*}(k \mid k)\right\| _{Q_{L}}^{2}-\left\| {\mathbf {u}}_{{Le}}^{*}(k \mid k)\right\| _{R_{L}}^{2} \\ &-\sum _{f=1}^{m-1}\left( \left\| {\mathbf {x}}_{f e}^{*}(k \mid k)\right\| _{Q_{f}}^{2}-\left\| {\mathbf {u}}_{f e}^{*}(k \mid k)\right\| _{R_{f}}^{2}\right) , \end{aligned}$$

and

$$\begin{aligned}V(k+1)-V(k)&=J^{*}(k+1)-J^{*}(k) \le J(k+1)-J^{*}(k) \nonumber \\ &\le -\left\| {\mathbf {x}}_{{Le}}^{*}(k \mid k)\right\| _{Q_{L}}^{2}-\left\| {\mathbf {u}}_{{Le}}^{*}(k \mid k)\right\| _{R_{L}}^{2}\nonumber \\ &\qquad -\sum _{f=1}^{m-1}\left( \left\| {\mathbf {x}}_{f e}^{*}(k \mid k)\right\| _{Q_{f}}^{2}-\left\| {\mathbf {u}}_{f e}^{*}(k \mid k)\right\| _{R_{f}}^{2}\right) , \end{aligned}$$
(74)

indicating that the Lyapunov function is decreasing, and uniform asymptotical stability is confirmed.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Wei, J., Zhu, B. Model predictive control for trajectory-tracking and formation of wheeled mobile robots. Neural Comput & Applic 34, 16351–16365 (2022). https://doi.org/10.1007/s00521-022-07195-4

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00521-022-07195-4

Keywords

Navigation