Skip to main content
Log in

Motion Controller Design of Wheeled Inverted Pendulum with an Input Delay Via Optimal Control Theory

  • Published:
Journal of Optimization Theory and Applications Aims and scope Submit manuscript

Abstract

This paper investigates the control of Back-and-Forth motion to fetch object and Lower-Raise-Head motion to avoid obstacle of the wheeled inverted pendulum system with an input delay. In controlling the Back-and-Forth motion, the linear optimal control theory is used because the tilt angle of the pendulum is forced to be small enough by minimizing a quadratic performance criterion with large weight of tilt angle error, and the controller is represented by using predictor-based feedback. In controlling the Lower-Raise-Head motion, linearized models do not work due to the strong nonlinearity caused by big tilt angle for avoiding obstacle. Without considering yaw movement, the wheeled inverted pendulum system is decoupled, the subsystem governing the state of the tilt angle is transformed into a simple linear system by using feedback linearization. With a properly chosen trajectory tracking target, the control of Lower-Raise-Head motion is solved on the basis of optimal trajectory tracking control for linear subsystems with an input delay. Numerical simulations illustrate the effectiveness of the proposed approaches.

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

Similar content being viewed by others

References

  1. Awrejcewicz, J., Kudra, G.: The piston–connecting rod–crankshaft system as a triple physical pendulum with impacts. Int. J. Bifurc. Chaos. 15(7), 2207–2226 (2005)

    Article  Google Scholar 

  2. Awrejcewicz, J., Wasilewski, G., Kudra, G., Reshmin, S.A.: An experiment with swinging up a double pendulum using feedback control. J. Comput. Syst. Sci. Int. 51(2), 176–182 (2012)

    Article  MathSciNet  MATH  Google Scholar 

  3. Jeong, S., Takahashi, T.: Wheeled inverted pendulum type assistant robot: design concept and mobile control. Intell. Serv. Robot. 1(4), 313–320 (2008)

    Article  Google Scholar 

  4. Nawawi, S.W., Ahmad, M.N., Osman, J.: Real-time control of a two-wheeled inverted pendulum mobile robot. World Acad. Sci. Eng. Technol. 29, 214–220 (2008)

    Google Scholar 

  5. Pathak, K., Franch, J., Agrawal, S.K.: Velocity and position control of a wheeled inverted pendulum by partial feedback linearization. IEEE Trans. Robot. 21(3), 505–513 (2005)

    Article  Google Scholar 

  6. Huang, J., Guan, Z.H., Matsuno, K., Fukuda, T., Sekiyama, K.: Sliding-mode velocity control of mobile-wheeled inverted-pendulum systems. IEEE Trans. Robot. 26(4), 750–758 (2010)

    Article  Google Scholar 

  7. Li, Z.J., Yang, C.G., Fan, L.P.: Advanced Control of Wheeled Inverted Pendulum Systems. Springer, London (2013)

    Book  Google Scholar 

  8. Yavin, Y., Frangos, C.: Feasible command strategies for the control of motion of a simple mobile robot. J. Optim. Theory Appl. 90(3), 671–692 (1996)

    Article  MathSciNet  MATH  Google Scholar 

  9. Ambrosino, G., Celentano, G., Garofalo, F.: Tracking control of high-performance robots via stabilizing controllers for uncertain systems. J. Optim. Theory Appl. 50(2), 239–255 (1986)

    Article  MathSciNet  MATH  Google Scholar 

  10. Saramago, S.F.P., Steffen, V.: Trajectory modeling of robot manipulators in the presence of obstacles. J. Optim. Theory Appl. 110(1), 17–34 (2001)

    Article  MathSciNet  MATH  Google Scholar 

  11. Liu, B., Hu, H.Y.: Group delay induced instability and its suppression for a controlled-double pendulum. J. Vib. Eng. 5, 128–135 (2009)

    Google Scholar 

  12. Chen, J.D.: LMI-based robust \(H_{\infty }\) control of uncertain neutral systems with state and input delays. J. Optim. Theory Appl. 126(3), 553–570 (2005)

    Article  MathSciNet  MATH  Google Scholar 

  13. Mahmoud, M.S., Zribi, M.: \(H_{\infty }\) controllers for time-delay systems using linear matrix inequalities. J. Optim. Theory Appl. 100(1), 89–122 (1999)

    Article  MathSciNet  MATH  Google Scholar 

  14. Zhou, Y.S., Wang, Z.H.: Optimal feedback control for linear systems with input delays revisited. J. Optim. Theory Appl. 163(3), 989–1017 (2014)

    Article  MathSciNet  MATH  Google Scholar 

  15. Arstein, Z.: Linear systems with delayed control: a reduction. IEEE Trans. Autom. Control 27(4), 869–879 (1982)

    Article  Google Scholar 

  16. Zhou, B.: Truncated Predictor Feedback for Time-delay Systems. Springer, Berlin (2014)

    Book  MATH  Google Scholar 

  17. Mondi, S., Michiels, W.: Finite spectrum assignment of unstable time-delay systems with a safe implementation. IEEE Trans. Autom. Control 48(12), 2207–2212 (2003)

    Article  Google Scholar 

  18. Cai, G.P., Huang, J.Z., Yang, S.X.: An optimal control method for linear systems with time delay. Comput. Struct. 81(15), 1539–1546 (2003)

    Article  MathSciNet  Google Scholar 

  19. Sastry, S.: Nonlinear Systems: Analysis, Stability and Control. Springer, New York (1999)

    Book  MATH  Google Scholar 

Download references

Acknowledgments

The authors thank the financial support of NSF of China under Grant 11372354 and Funding of Jiangsu Innovation Program for Graduate Education (CXLX13-129), and thank the anonymous reviewers for their valuable comments that help much in improving the presentation of the manuscript.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Zaihua Wang.

Additional information

Communicated by Felix L. Chernousko.

Appendix: Derivation of Equation (12)

Appendix: Derivation of Equation (12)

Consider the optimal control problem of system (10) that minimizes the performance criterion (11). Firstly, let consider the special case of \(\tau =0\), the augmented quadratic performance criterion can be defined as

$$\begin{aligned} \bar{J}= & {} \frac{1}{2}\mathbf{Y }^\mathrm{T}(t_f)\mathbf{Q }_0\mathbf{Y }(t_f)+\frac{1}{2}\int _0^{t_f}\left[ \mathbf{Y }^\mathrm{T}(t)\mathbf{Q }\mathbf{Y }(t)+u_{2}^\mathrm{T}(t)\mathbf{R }u_2(t)\right] \mathrm{d}t\\&+\int _0^{t_f}\left[ \varvec{\lambda }^\mathrm{T}(t)(\mathbf{A }\mathbf{Y }(t)+\mathbf{B }u_2(t)+\varvec{\omega }(t)-\dot{\mathbf{Y }}(t))\right] \mathrm{d}t, \end{aligned}$$

and the Hamiltonian function is defined by

$$\begin{aligned} H(\mathbf{Y },u_2,\varvec{\lambda },t)=\frac{1}{2}\mathbf{Y }^\mathrm{T}(t)\mathbf{Q }\mathbf{Y }(t)+\frac{1}{2}u_{2}^\mathrm{T}(t)\mathbf{R }u_2(t) +\varvec{\lambda }^\mathrm{T}(t)(\mathbf{A }\mathbf{Y }(t)+\mathbf{B }u_2(t)+\varvec{\omega }(t)). \end{aligned}$$
(20)

Then, \(\bar{J}\) can be rewritten in the form of

$$\begin{aligned} \bar{J}= & {} \frac{1}{2}\mathbf{Y }^\mathrm{T}(t_f)\mathbf{Q }_0\mathbf{Y }(t_f)-\varvec{\lambda }^\mathrm{T}(t_f)\mathbf{Y }(t_f)+\varvec{\lambda }^\mathrm{T}(0)\mathbf{Y }(0)\\&+\int _0^{t_f}\left[ H(\mathbf{Y },u_2,\varvec{\lambda },t)+\dot{\varvec{\lambda }}^\mathrm{T}\mathbf{Y }(t)\right] \mathrm{d}t. \end{aligned}$$

Consider the variation of \(\bar{J}\) due to the variables \(u_2\) and \(\mathbf{Y }\), we have

$$\begin{aligned} \delta \bar{J}=\delta \mathbf{Y }^\mathrm{T}(t_f)\mathbf{Q }_0\mathbf{Y }(t_f)-\delta \mathbf{Y }^\mathrm{T}(t_f)\varvec{\lambda }(t_f) +\int _0^{t_f}\left[ \delta \mathbf{Y }^\mathrm{T}\left( \frac{\partial H}{\partial \mathbf{Y }}+\dot{\varvec{\lambda }}\right) +\delta u_2^\mathrm{T}\frac{\partial H}{\partial u_2}\right] \mathrm{d}t \end{aligned}$$

By setting \(\delta \bar{J}=0\), the necessary conditions for the minimum value problem are

$$\begin{aligned} \dot{\varvec{\lambda }}=-\frac{\partial H}{\partial \mathbf{Y }}, ~~ \dot{\mathbf{Y }}=\frac{\partial H}{\partial \varvec{\lambda }}, ~~ \frac{\partial H}{\partial u_2}=0 , ~~ \varvec{\lambda }(t_f)=\mathbf{Q }_0\mathbf{Y }(t_f). \end{aligned}$$
(21)

From (20) and (21), the optimal control of linear system (10) minimizes performance criterion (11) is given by

$$\begin{aligned} u_2^{*}(t)=-\mathbf{R }^{-1}\mathbf{B }^\mathrm{T}\varvec{\lambda }(t), \end{aligned}$$
(22)

where \(\varvec{\lambda }(t)\) is determined by the following differential equations

$$\begin{aligned} -\dot{\varvec{\lambda }}= & {} \mathbf{Q }\mathbf{Y }+\mathbf{A }^\mathrm{T}\varvec{\lambda }, ~~~~~~\varvec{\lambda }(t_f)=\mathbf{Q }_0\mathbf{Y }(t_f) ,\end{aligned}$$
(23)
$$\begin{aligned} \dot{\mathbf{Y }}= & {} \mathbf{A }\mathbf{Y }-\mathbf{S }\varvec{\lambda }+\varvec{\omega },~~~~ \mathbf{Y }(0)=\mathbf X (0)-\bar{\mathbf{X }}(0), \end{aligned}$$
(24)

where \(\mathbf{S }=\mathbf{B }\mathbf{R }^{-1}\mathbf{B }^\mathrm{T}\). According to (22), (23), and (24), we have

Lemma 5.1

When \(\tau =0\), the optimal control of the linear system (10) that minimizes (11) is given by

(25)

where \(\mathbf{P }_{0}(t)\in {\mathbb {R}}^{n\times n}\) and \(\mathbf b _{0}(t)\in {\mathbb {R}}^{n}\) are the solutions of the Riccati differential equations

(26)
(27)

Proof

Firstly, let

$$\begin{aligned} \varvec{\lambda }(t)=\mathbf{P }_{0}\mathbf{Y }(t)+\mathbf b _{0}(t). \end{aligned}$$
(28)

Differentiating at the both sides of (28), combining (23) and (24) lead to

$$\begin{aligned} \dot{\varvec{\lambda }}=\dot{\mathbf{P }}_0\mathbf{Y }+\mathbf{P }_0 \dot{\mathbf{Y }}+\dot{\mathbf{b }}_0 =\dot{\mathbf{P }}_0\mathbf{Y }+\mathbf{P }_0(\mathbf{A }\mathbf{Y }-\mathbf{S }\varvec{\lambda }+\varvec{\omega })+\dot{\mathbf{b }}_0 =-\mathbf{Q }\mathbf{Y }-\mathbf{A }^\mathrm{T}\varvec{\lambda }. \end{aligned}$$
(29)

Equating the corresponding coefficients of both sides of (29), one has the Riccati equations (26) and (27).\(\square \)

From Lemma 5.1, the optimal control \(u_2^{*}(t)\) consists of two parts, \(-\mathbf{R }^{-1}\mathbf{B }^\mathrm{T}\mathbf{P }_0(t)\mathbf{Y }(t)\) is a feedback control that is determined by the current state, and the rest \(-\mathbf{R }^{-1}\mathbf{B }^\mathrm{T}\mathbf b _0(t)\) is a feedforward that is determined by the tracking target.

When \(\tau \ne 0\), let us introduce a new integral state transformation as follows:

$$\begin{aligned} \mathbf Z (t)=\mathbf{Y }(t)+\int _{t-\tau }^t\mathrm{e}^{-\mathbf{A }(s-t+\tau )}[\mathbf{B }u_2(s)+\varvec{\omega }(s+\tau )]\mathrm{d}s. \end{aligned}$$
(30)

Then, with \(\mathbf{B }_0=\mathrm{e}^{-\mathbf{A }\tau }\mathbf{B }, \mathbf G (t)=\mathrm{e}^{-A\tau }\varvec{\omega }(t+\tau )\), the system (10) is changed to a delay-free form

$$\begin{aligned} \dot{\mathbf{Z }}(t)=\mathbf{A }\mathbf Z (t)+\mathbf{B }_0u_2(t)+\mathbf G (t). \end{aligned}$$
(31)

Equation (10) has an equivalent form as follows:

$$\begin{aligned} \mathbf{Y }(t)=\mathrm{e}^{\mathbf{A }t}\mathbf{Y }(0)+\int _{0}^t\mathrm{e}^{\mathbf{A }(t-s)}[\mathbf{B }u_2(s-\tau )+\varvec{\omega }(s)]\mathrm{d}s, \end{aligned}$$

and the solution \(\mathbf{Y }(t)\) satisfies

$$\begin{aligned} \begin{aligned} \mathbf{Y }(t+\tau )&=\mathrm{e}^{\mathbf{A }\tau }\left( \mathbf{Y }(t)+\int _{t}^{t+\tau }\mathrm{e}^{\mathbf{A }(t-s)}[\mathbf{B }u_2(s-\tau )+\varvec{\omega }(s)]\mathrm{d}s\right) \\&=\mathrm{e}^{\mathbf{A }\tau }\left( \mathbf{Y }(t)+\int _{t-\tau }^t\mathrm{e}^{-\mathbf{A }(s-t+\tau )}[\mathbf{B }u_2(s)+\varvec{\omega }(s+\tau )]\mathrm{d}s\right) =\mathrm{e}^{\mathbf{A }\tau }\mathbf Z (t). \end{aligned} \end{aligned}$$
(32)

Thus, the initial condition \(\mathbf{Y }(0)=\mathbf{Y }_{0}\) for system (10) is changed into \(\mathbf Z (0)=\mathrm{e}^{-\mathbf{A }\tau }\mathbf{Y }(\tau )\).

The quadratic performance criterion \(J\) for system (11) can be decomposed by \(J=J_1+J_2,\) where \(J_1=\frac{1}{2}\int _0^{\tau }\mathbf{Y }^\mathrm{T}\mathbf{Q }\mathbf{Y }\mathrm{d}t\) is fixed because the control does not take effect when \(t\in [0,\tau [\), and

$$\begin{aligned} J_2=\frac{1}{2}\mathbf{Y }^\mathrm{T}(t_f)\mathbf{Q }_0\mathbf{Y }(t_f)+\frac{1}{2}\int _0^{t_{f}-\tau }\left[ \mathbf{Y }^\mathrm{T}(t+\tau )\mathbf{Q }\mathbf{Y }(t+\tau )+u_2^\mathrm{T}(t)\mathbf{R }u_2(t)\right] \mathrm{d}t. \end{aligned}$$
(33)

Hence, \(J_d=J_1+J_2=\min \) if and only if \(J_2=\min \). By substituting (32) into (33), the criterion \(J_2\) takes the form

$$\begin{aligned} J_2=\frac{1}{2}\mathbf Z ^\mathrm{T}(t_{f}-\tau )\tilde{\mathbf{Q }}_0\mathbf Z (t_{f}-\tau ) +\frac{1}{2}\int _0^{t_{f}-\tau }\left[ \mathbf Z ^\mathrm{T}(t)\tilde{\mathbf{Q }}\mathbf Z (t)+u_2^\mathrm{T}(t)\mathbf{R }u_2(t)\right] \mathrm{d}t, \end{aligned}$$

where \(\tilde{\mathbf{Q }}=\left( \mathrm{e}^{\mathbf{A }\tau }\right) ^\mathrm{T}\mathbf{Q }\mathrm{e}^{\mathbf{A }\tau },\tilde{\mathbf{Q }}_0=\left( \mathrm{e}^{\mathbf{A }\tau }\right) ^\mathrm{T}\mathbf{Q }_0\mathrm{e}^{\mathbf{A }\tau }\). Thus the original optimal control problem can be converted into the optimal control problem for linear delay-free system (31) associated with the performance criterion \(J_2\).

According to Lemma 5.1, the optimal control for linear system (31) that minimizes \(J_2\) is given by

$$\begin{aligned} u_2^*(t)=-\mathbf{R }^{-1}\mathbf{B }_0^\mathrm{T}[\mathbf{P }_d(t)\mathbf Z (t)+\mathbf b _d(t)], \end{aligned}$$
(34)

where \(\mathbf{P }_d(t)\in {\mathbb {R}}^{n\times n}\) and \(\mathbf b _{d}(t)\in {\mathbb {R}}^{n}\) are the solutions of the Riccati differential equations

$$\begin{aligned} \dot{\mathbf{P }_d}= & {} -\mathbf{P }_d\mathbf{A }-\mathbf{A }^\mathrm{T}\mathbf{P }_d+\mathbf{P }_d\mathbf{B }_0\mathbf{R }^{-1}\mathbf{B }_0^\mathrm{T}\mathbf{P }_d-\tilde{\mathbf{Q }}, ~~~~\mathbf{P }_{d}(t_f-\tau )=\tilde{\mathbf{Q }}_0 ,\end{aligned}$$
(35)
$$\begin{aligned} \dot{\mathbf{b }}_{d}= & {} -[\mathbf{A }-\mathbf{B }_0\mathbf{R }^{-1}\mathbf{B }_{0}^\mathrm{T}\mathbf{P }_{d}]^\mathrm{T}\mathbf b _{d}-\mathbf{P }_{d}\mathbf G (t), ~~~~\mathbf b _{d}(t_f-\tau )=0. \end{aligned}$$
(36)

By substituting (32) into (34), the delayed optimal control of system (10) can be expressed in terms of the state \(\mathbf{Y }(t)\), rather than \(\mathbf Z (t)\), in the form of

$$\begin{aligned} u_2^{*}(t-\tau )=-\mathbf{R }^{-1}(\mathrm{e}^{-\mathbf{A }\tau }\mathbf{B })^\mathrm{T}\left[ \mathbf{P }_d(t-\tau )\mathrm{e}^{-\mathbf{A }\tau }\mathbf{Y }(t)+\mathbf b _d(t-\tau )\right] . \end{aligned}$$
(37)

Lemma 5.2

Let \(u_2^*(t-\tau )\) be the delayed optimal control of system (10), \(u_{20}^*(t)\) be the optimal control of system (10) with \(\tau =0\), then

$$\begin{aligned} u_2^{*}(t-\tau )=u_{20}^{*}(t)=-\mathbf{R }^{-1}\mathbf{B }^\mathrm{T}[\mathbf{P }_0(t)\mathbf{Y }(t)+\mathbf b _0(t)],~~(t\in [\tau ,t_f]), \end{aligned}$$

where \(\mathbf{P }_{0}(t)\in {\mathbb {R}}^{n\times n}\) and \(\mathbf b _{0}(t)\in {\mathbb {R}}^{n}\) are the solutions of the Riccati differential equations (26) and (27).

Proof

According to Lemma 5.1, the optimal control of the linear system (10) with \(\tau =0\) can be given by (25). Multiplying (26) by \((\mathrm{e}^{\mathbf{A }\tau })^\mathrm{T}\) from the left and by \(\mathrm{e}^{\mathbf{A }\tau }\) from the right, respectively, one see that \(\mathbf W (t)=(\mathrm{e}^{\mathbf{A }\tau })^\mathrm{T}{\mathbf{P }_{0}}(t)\mathrm{e}^{\mathbf{A }\tau }\) is the solution of the following differential equation

$$\begin{aligned} \dot{\mathbf{W }}(t)=-\mathbf W (t)\mathbf{A }-\mathbf{A }^\mathrm{T}\mathbf W (t)+\mathbf W (t)\mathbf{B }_0\mathbf{R }^{-1}\mathbf{B }_0^\mathrm{T}\mathbf W (t)-\tilde{\mathbf{Q }}, \end{aligned}$$

subject to \(\mathbf W (t_f)=\tilde{\mathbf{Q }}_0\). Replacing \(t\) with \(t-\tau \) in (35) leads to

$$\begin{aligned} \dot{\mathbf{P }_d}(t-\tau )=-\mathbf{P }_d(t-\tau )\mathbf{A }-\mathbf{A }^\mathrm{T}\mathbf{P }_d(t-\tau ) +\mathbf{P }_d(t-\tau )\mathbf{B }_0\mathbf{R }^{-1}\mathbf{B }_0^\mathrm{T}\mathbf{P }_d(t-\tau )-\tilde{\mathbf{Q }}, \end{aligned}$$

subject to \(\mathbf{P }_{d}(t_f-\tau )=\tilde{\mathbf{Q }}_0\). It follows that both \(\mathbf W (t)\) and \(\mathbf{P }_d(t-\tau )\) satisfy the same differential equation under the same terminal condition. Thus, \(\mathbf W (t)\) and \(\mathbf{P }_d(t-\tau )\) must be the same, namely

$$\begin{aligned} \mathbf{P }_{d}(t-\tau )=\left( \mathrm{e}^{\mathbf{A }\tau }\right) ^\mathrm{T}\mathbf{P }_{0}(t)\mathrm{e}^{\mathbf{A }\tau }. \end{aligned}$$
(38)

Similarly, multiplying (27) by \((\mathrm{e}^{A\tau })^\mathrm{T}\) from the left, \(\mathbf V (t)=(\mathrm{e}^{\mathbf{A }\tau })^\mathrm{T}\mathbf{b _{0}}\) is the solution of

$$\begin{aligned} \dot{\mathbf{V }}(t)=-\left[ \mathbf{A }-\mathrm{e}^{-\mathbf{A }\tau }\mathbf{S }\mathbf{P }_{0}\mathrm{e}^{\mathbf{A }\tau }\right] ^\mathrm{T}\mathbf V (t)-\left( \mathrm{e}^{\mathbf{A }\tau }\right) ^\mathrm{T}\mathbf{P }_{0}\varvec{\omega }, ~~\mathbf V (t_f)=0. \end{aligned}$$

By substituting (38) into (36) and replacing \(t\) with \(t-\tau \) leads to

$$\begin{aligned} \dot{\mathbf{b }}_d(t-\tau )=-\left[ \mathbf{A }-\mathrm{e}^{-\mathbf{A }\tau }\mathbf{S }\mathbf{P }_{0}\mathrm{e}^{\mathbf{A }\tau }\right] ^\mathrm{T}\mathbf b _{d}(t-\tau )-\left( \mathrm{e}^{\mathbf{A }\tau }\right) ^\mathrm{T}\mathbf{P }_{0}\varvec{\omega },~~ \mathbf b _{d}(t_f-\tau )=0. \end{aligned}$$

Hence, \(\mathbf V (t)\) and \(\mathbf b _d(t-\tau )\) must be the same, namely

$$\begin{aligned} \mathbf b _{d}(t-\tau )=\left( \mathrm{e}^{\mathbf{A }\tau }\right) ^\mathrm{T}\mathbf b _{0}(t). \end{aligned}$$
(39)

By substituting (38) and (39) into (37), the following simple relationship holds for \(t\in [\tau ,t_f]\)

$$\begin{aligned} u_2^{*}(t-\tau )=-\mathbf{R }^{-1}\mathbf{B }^\mathrm{T}[\mathbf{P }_{0}(t)\mathbf{Y }(t)+\mathbf b _0(t)]=u_{20}^{*}(t). \end{aligned}$$

This completes the proof. \(\square \)

Lemma 5.2 implies that the input delay does not affect the optimal control, when the system has no state delay, the optimal control of the delay-free system fully determines the optimal control of the corresponding system with an input delay, no matter how large the delay is.

When \(t_f\rightarrow +\infty \), because the control takes effect only when \(t\ge \tau \), the \(J\) can be simplified as follows:

$$\begin{aligned} J=&\,\frac{1}{2}\int _0^{+\infty }\left[ \mathbf{Y }^\mathrm{T}(t)\mathbf{Q }\mathbf{Y }(t)+\nu ^\mathrm{T}(t-\tau )\mathbf{R }\nu (t-\tau )\right] \mathrm{d}t \\ =&\,\frac{1}{2}\int _0^{+\infty }\mathbf{Y }^\mathrm{T}(t)\mathbf{Q }\mathbf{Y }(t)\mathrm{d}t\\&+\frac{1}{2}\int _0^{\tau }\nu ^\mathrm{T}(t-\tau )\mathbf{R }\nu (t-\tau )\mathrm{d}t +\frac{1}{2}\int _{\tau }^{+\infty }\nu ^\mathrm{T}(t-\tau )\mathbf{R }\nu (t-\tau )\mathrm{d}t \\ =&\,\frac{1}{2}\int _0^{+\infty }\left[ \mathbf{Y }^\mathrm{T}(t)\mathbf{Q }\mathbf{Y }(t)+\nu ^\mathrm{T}(t)\mathbf{R }\nu (t)\right] \mathrm{d}t, \end{aligned}$$

and (26) becomes algebraic equation, then, we have

$$\begin{aligned}&-\mathbf{P }_0\mathbf{A }-\mathbf{A }^\mathrm{T}\mathbf{P }_0+\mathbf{P }_0\mathbf{B }\mathbf{R }^{-1}\mathbf{B }^\mathrm{T}\mathbf{P }_0-\mathbf{Q }=\mathbf 0 ,\end{aligned}$$
(40)
$$\begin{aligned}&\dot{\mathbf{b }}_{0}=-\left[ \mathbf{A }-\mathbf{B }\mathbf{R }^{-1}\mathbf{B }^\mathrm{T}\mathbf{P }_{0}\right] ^\mathrm{T}\mathbf b _{0}-\mathbf{P }_{0}\varvec{\omega },~~~~\mathbf b _{0}(t_f)=0. \end{aligned}$$
(41)

In this case, the matrix \(\mathbf{P }_0\) is time independent and the Riccati equation becomes easy to solve. That is why the case of infinite time horizon is more often studied in practice.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Zhou, Y., Wang, Z. Motion Controller Design of Wheeled Inverted Pendulum with an Input Delay Via Optimal Control Theory. J Optim Theory Appl 168, 625–645 (2016). https://doi.org/10.1007/s10957-015-0759-z

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s10957-015-0759-z

Keywords

Mathematics Subject Classification

Navigation