Skip to main content
Log in

A novel ADRC-based design for a kind of flexible aerocraft

  • Research Article
  • Published:
Control Theory and Technology Aims and scope Submit manuscript

Abstract

The paper presents a novel control design, which is based on the idea of active disturbance rejection control (ADRC), for a kind of flexible aerocraft whose controlled variable cannot be measured directly. Since the original frame of ADRC cannot be directly applied, the paper puts forward a new extended state observer (ESO) and the corresponding ADRC law. In order to assign the poles of the closed-loop system to ideal positions such that the vibration can be quickly suppressed, an elastic damping term is added into feedback law. The advantages of the new ESO for effectively estimating both the rigid mode and elastic mode from the measurements are discussed. Moreover, the analysis on the stability, the relative stability and the steady state of the closed-loop system is given. Finally, the effectiveness and robustness of the proposed ADRC are verified by simulations.

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
Fig. 15

Similar content being viewed by others

References

  1. Pan, C., Rong, J., Xu, T., & Xiang, D. (2020). Novel approach for active vibration control of a flexible missile. Defence Technology, 16(4), 956–967.

    Article  Google Scholar 

  2. Su, C., Wang, X., Xu, Z., Li, J., & Wang, X. (2019). Research on elastic vibration suppression of missile attitude control system. Journal of Projectiles, Rockets, Missiles and Guidance, 39(5), 154–157 (in Chinese).

    Google Scholar 

  3. Oh, C., Bang, H., & Park, C. (2008). Attitude control of a flexible launch vehicle using an adaptive notch filter: ground experiment. Control Engineering Practice, 16(1), 30–42.

    Article  Google Scholar 

  4. Wang, Z., & Liu, Z. (2010). Aero-elastic stability design of spinning ballistic missile with sliding mode control. In International Conference on Computer Design and Applications (pp. 3147–3152). Qinhuangdao, China.

  5. Khoshnood, A. M., Roshanian, J., & Jafari, A. A. (2010). An adjustable model reference adaptive control for a flexible launch vehicle. Journal of Dynamic Systems, Measurement, and Control, 132(4), 1–7.

    Article  Google Scholar 

  6. Piao, M., Zhang, Y., Sun, M., Yang, Z., & Wang, Z. (2019). Adaptive aeroservoelastic mode stabilization of flexible airbreathing hypersonic vehicle. Journal of Vibration and Control, 25(15), 2124–2142.

    Article  MathSciNet  Google Scholar 

  7. George, K. K., & Bhat, M. S. (1998). Two-degree-of-freedom H-infinity robust controller for a flexible missile. Journal of Guidance, Control, and Dynamics, 21(3), 518–520.

    Article  Google Scholar 

  8. Zhang, L., Peng, C., Wang, Y., et al. (2010). Two-degree-of-freedom \({\rm H}_\infty\) controller design for vibration attenuation and tracking of high length-to-diameter ratio missile. Journal of Balls, 22(4), 1–5.

    Google Scholar 

  9. Maki, M.C., & Vegte, J. (1972). Optimal and constrained-optimal control of a flexible launch vehicle. AIAA Journal, 10(6): 796–799.

    Article  Google Scholar 

  10. Liu, X., Huang, W., Wu, Y., & Xiong, S. (2014). Passive, active and compound control for elastic missile and applicability analysis. The Journal of New Industrialization, 4(5), 11–19.

    Google Scholar 

  11. Huang, Y., & Xue, W. (2014). Active disturbance rejection control: methodology and theoretical analysis. ISA transactions, 53(4), 963–976.

    Article  MathSciNet  Google Scholar 

  12. Han, J. (2009). From PID to active disturbance rejection control. IEEE Transactions on Industrial Electronics, 56(3), 900–906.

    Article  Google Scholar 

  13. Huang, Y., Xu, K., Han, J., & Lam, J. (2001). Flight control design using extended state observer and non-smooth feedback. In Proceedings of the 40th IEEE Conference on Decision and Control (pp. 223–228). Orlando, FL, USA.

  14. Zhao, S., & Gao, Z. (2013). An active disturbance rejection based approach to vibration suppression in two-inertia systems. Asian Journal of Control, 15(2), 350–362.

    Article  MathSciNet  Google Scholar 

  15. Xue, W., Madonski, R., Lakomy, K., Gao, Z., & Huang, Y. (2017). Add-on module of active disturbance rejection for set-point tracking of motion control systems. IEEE Transactions on Industry Applications, 53(4), 4028–4040.

    Article  Google Scholar 

  16. Xiang, G., Huang, Y., Yu, J., Zhu, M., & Su, J. (2018). Intelligence evolution for service robot: An ADRC perspective. Control Theory and Technology, 16(4), 324–335.

    Article  MathSciNet  Google Scholar 

  17. Zheng, Q., & Gao, Z. (2018). Active disturbance rejection control: some recent experimental and industrial case studies. Control Theory and Technology, 16(4), 301–313.

    Article  MathSciNet  Google Scholar 

  18. Madonski, R., Ramirez-Neria, M., Stankovic, M., Shao, S., Gao, Z., Yang, J., & Li. S. (2019). On vibration suppression and trajectory tracking in largely uncertain torsional system: An error-based ADRC approach. Mechanical Systems and Signal Processing. https://doi.org/10.1016/j.ymssp.2019.106300.

    Article  Google Scholar 

  19. Fareh, R., Al-Shabi, M., Bettayeb, M., & Ghommam, J. (2020). Robust active disturbance rejection control for flexible link manipulator. Robotica, 38(1), 118–135.

    Article  Google Scholar 

  20. Chen, S., Bai, W., Hu, Y., et al. (2020) On the conceptualization of total disturbance and its profound implications. Science China Information Sciences. https://doi.org/10.1007/s11432-018-9644-3.

    Article  MathSciNet  Google Scholar 

Download references

Acknowledgements

This work was supported by the National Key R&D Program of China (No. 2018YFA0703800) and the National Center for Mathematics and Interdisciplinary Sciences, Chinese Academy of Sciences.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Yi Huang.

Appendix

Appendix

Proof of Lemma 1

Since the pair \(({\bar{A}},{\bar{C}}_1)\) is observable, the matrix \(M_O\) is nonsingular. Denote

$$\begin{aligned} v(t) = M_O (X(t) - z(t) - M(t)), \end{aligned}$$
(a1)

the dynamics of v(t) is obtained as follows:

$$\begin{aligned} {\dot{v}}(t) = ({\bar{A}}_{e}-L_{e}{\bar{c}}_{e}) {v}(t) + H f_v, \end{aligned}$$
(a2)

where \(L_{e}=M_O L,\) \({\bar{c}}_{e}={\bar{C}}_1 M_O^{-1},\)

$$\begin{aligned} {\bar{A}}_{e} =M_O {\bar{A}} M_O^{-1}= \left[ \begin{array}{c@{~~~}c@{~~~}c@{~~}cc} 0&{}1&{}0 &{}0 &{} 0 \\ 0&{}0&{}1 &{}0 &{} 0 \\ 0&{}0&{}0 &{}1 &{} 0 \\ 0&{}0&{}0 &{}0 &{} 1\\ 0 &{} 0 &{} 0&{}-{\bar{\omega }}_1^2 &{} -2\xi {\bar{\omega }}_1 \end{array}\right] , \end{aligned}$$
(a3)

and

$$\begin{aligned} f_v = {\bar{c}}_{e} {\bar{A}}_{\mathrm{e}}^{5} M(t) +\textstyle \sum \limits _{j=1}^{3} C_1 A^{4-j} B_2 f^{(j)}+\textstyle \sum \limits _{j=1}^{3} C_1 A^{4-j} B_1 f_q^{(j)}. \end{aligned}$$
(a4)

Due to the assumption that the disturbance \(f_q \in \mathbb{R}\) is bounded, \(f\in \mathbb{R}\) and \(f_q \in \mathbb{R}\) have the bounded first 3rd-order derivatives; \(f_v\) is bounded.

Since the pair \(({\bar{A}}_{e},~{\bar{c}}_{e})\) is observable, there exist \({\bar{L}}_{e,1}\), \(\bar{L}_{e,2}\) and the invertible matrices \(P_{L1}\) and \({P}_{L2}\) such that

$$\left\{\begin{aligned} & {\bar{A}}_{e}-{\bar{L}}_{e,1}{\bar{c}}_{e}= \beta P_{L1}^{-1} M_\varLambda P_{L1}, \\ &{\bar{A}}_{e}-\bar{L}_{e,2} {\bar{c}}_{e}= {P}_{L2}^{-1} M_\varLambda {P}_{L2}, \end{aligned}\right.$$
(a5)

where

$$\begin{aligned} M_\varLambda = \begin{bmatrix} {\lambda }_{1}&{} 0 &{} \cdots &{}0\\ 0&{}{\lambda }_{2}&{}\cdots &{} 0 \\ \vdots &{}\vdots &{} &{}\vdots \\ 0&{} 0 &{}\cdots &{}{\lambda }_{{5}} \end{bmatrix}. \end{aligned}$$
(a6)

The relationship between \(P_{L1}\) and \({P}_{L2}\) is obtained as follows:

$$\begin{aligned} P_{L1}= {P}_{L2} T_P, \end{aligned}$$
(a7)

where \(T_P = M_P^{-1}M_\beta M_P\) and

$$\begin{aligned} M_P=\begin{bmatrix} 1&{}0&{} 0&{}0 &{}0\\ 2\xi {\bar{\omega }}_1&{} 1&{}0&{}0 &{}0 \\ {\bar{\omega }}_1^2&{}2\xi {\bar{\omega }}_1&{} 1&{}0&{}0\\ 0&{}{\bar{\omega }}_1^2&{}2\xi {\bar{\omega }}_1&{} 1&{}0\\ a&{}0&{}{\bar{\omega }}_1^2&{}2\xi {\bar{\omega }}_1&{} 1 \end{bmatrix}, ~~M_{\beta } = \begin{bmatrix} \beta ^{4} &{}0&{} \cdots &{}0\\ 0&{}\beta ^{3}&{}\cdots &{}0\\ \vdots &{}\vdots &{} &{}\vdots \\ 0&{}0&{}\cdots &{}1 \end{bmatrix}. \end{aligned}$$
(a8)

a is a constant. Denote \(\tilde{v}(t) = T_P v(t)\), from (a2), (a5) and (a7), there is

$$\begin{aligned} \dot{\tilde{v}}(t) = \beta ({\bar{A}}_{e}-\bar{L}_{e,2}{\bar{c}}_{e}){\tilde{v}}(t) - H f_v. \end{aligned}$$
(a9)

Since the matrix \(({\bar{A}}_{e}-\bar{L}_{e,2}{\bar{c}}_{e})\) is Hurwitz, there exists a positive definite matrix \(Q_L\) such that

$$\begin{aligned} ({\bar{A}}_{e}-\bar{L}_{e,2}{\bar{c}}_{e})^{\mathrm{T}} Q_L+Q_L({\bar{A}}_{e}-\bar{L}_{e,2}{\bar{c}}_{e}) =-I. \end{aligned}$$
(a10)

Consider the Lyapunov function \({V}(t)={\tilde{v}}(t)^{\mathrm{T}} Q_L {\tilde{v}}(t)\). Then, we have

$$\begin{aligned} \lambda _{\min }(Q_L) \left\| {\tilde{v}}(t) \right\| ^2 \leqslant {V}(t) \leqslant \lambda _{\max }(Q_L) \left\| {\tilde{v}}(t) \right\| ^2 \end{aligned}$$
(a11)

and

$$\begin{aligned} \dfrac{{\rm d} \sqrt{ {V}(t)}}{{\rm d}t}&= \dfrac{ -\beta \left\| {\tilde{v}}(t) \right\| ^2-2 {\tilde{v}}(t)^{\mathrm{T}} Q_L H f_v}{2 \sqrt{ {V}(t)}}\nonumber \\&\leqslant -\dfrac{\beta }{2\lambda _{\max }(Q_L)} \sqrt{{V}(t)}+ \dfrac{\Vert Q_L \Vert | f_v |}{\sqrt{\lambda _{\min }(Q_L)}}. \end{aligned}$$
(a12)

Since \(f_v\) is bounded, i.e., \(|f_v |^2<\eta _{F}\) for some positive \(\eta _{F}\), with the help of Gronwall lemma, it is indicated from (a12) that

$$\begin{aligned} \sqrt{{V}(t)} \leqslant \mathrm{e}^{-\frac{\beta t}{2\lambda _{\max }(Q_L)} } \sqrt{ {V}(0)} + \dfrac{2 \Vert Q_L\Vert \eta _{F} \lambda _{\max }(Q_L)}{\beta \sqrt{\lambda _{\min }(Q_L)}}. \end{aligned}$$
(a13)

Due to (a11), (a13) and

$$\begin{aligned} \Vert X(t)-z(t) - M(t)\Vert \leqslant \Vert T_Q^{-1} \Vert \Vert T_P^{-1}\Vert \Vert \tilde{v}(t)\Vert , \end{aligned}$$
(a14)

(15) is obtained. \(\square\)

Proof of Theorem 1

First, it will be proved that there exist the controller parameters \(k_p,\) \(k_d,\) \(k_q\) and the observer parameter L, such that the closed-loop system is stable despite the uncertainties given in Table 1. If the controller parameters are chosen as \(\omega _c=2,\) \(\xi _c=2,\) \(\omega _q=3,\) \(\lambda (\bar{A}-L\bar{C})=10\{-0.9,-1,-0.95,-1.05,-1\},\) then it can be proved that the characteristic roots of the closed-loop system lie in the following regions, as shown in Table a1, i.e., the closed-loop system is stable, despite various uncertainties of the system parameters in the scopes of Table 1.

Table a1 Ranges of the characteristic roots

Since the closed-loop system is stable, the disturbances f and \(f_q\) are bounded. Denote the transfer function from the total disturbance f to the estimation \(z_i\) as \(\dfrac{D_{fi}(s)}{N(s)}\) and the transfer function from the disturbance \(f_q\) to the estimation \(z_i\) as \(\dfrac{D_{f_{qi}}(s)}{N(s)},\) \(i=1,\ldots ,5.\) It can be obtained that

$$\begin{aligned} \left\{ \begin{array}{l} Z_1(s)=\Delta \varPhi (s)+\dfrac{D_{f1}(s)}{N(s)}F(s)+\dfrac{D_{f_{q1}}(s)}{N(s)}F_q(s),\\ D_{f1}(0)=0,~~\dfrac{D_{f_{q1}}(0)}{N(0)} =\dfrac{c_1}{{\bar{\omega }}_1^2},\\ Z_2(s)=Q_1(s)+\dfrac{D_{f2}(s)}{N(s)}F(s)+\dfrac{D_{f_{q2}}(s)}{N(s)}F_q(s),\\ D_{f2}(0)=0,~~\dfrac{D_{f_{q2}}(0)}{N(0)} =\dfrac{1}{{\bar{\omega }}_1^2},\\ Z_3(s)=s \Delta \varPhi (s)+\dfrac{D_{f3}(s)}{N(s)}F(s)+\dfrac{D_{f_{q3}}(s)}{N(s)}F_q(s),\\ D_{f3}(0)=0,~~{D_{f_{q3}}(0)} =0,\\ Z_4(s)=s Q_1(s)+\dfrac{D_{f4}(s)}{N(s)}F(s)+\dfrac{D_{f_{q4}}(s)}{N(s)}F_q(s),\\ D_{f4}(0)=0,~~{D_{f_{q4}}(0)} =0,\\ Z_5(s)=\dfrac{D_{f5}(s)}{N(s)}F(s)+\dfrac{D_{f_{q5}}(s)}{N(s)}F_q(s),\\ \dfrac{D_{f5}(0)}{N(0)}=1, ~~{D_{f_{q5}}(0)}=0,\\ \end{array}\right. \end{aligned}$$
(a15)

where \(\Delta \varPhi (s),Q_1(s),F(s)\) and \(F_q(s)\) are the Laplace transforms of \(\Delta \varphi (t),~q_1(t),~f(t)\) and \(f_q(t)\), respectively.

Denote the estimation error as \(e_z=z-X.\) From (a15), it can be obtained that, when the closed-loop system reaches steady state, there is

$$\begin{aligned} e_{z_1}=\mathrm{O}(\dfrac{c_1}{{\bar{\omega }}_1^2}),~~e_{z_2}=\mathrm{O}(\dfrac{1}{{\bar{\omega }}_1^2}),~~e_{z_3}=e_{z_4}=e_{z_5}=0. \end{aligned}$$

The dynamic equation of the rigid mode can be rewritten as

$$\begin{aligned} \Delta \ddot{\varphi }= -k_p \Delta \varphi -k_d \Delta {\dot{\varphi }} +k_q \dot{q}_1 -b_{30}Ke_z. \end{aligned}$$
(a16)

Therefore, the steady-state error of the rigid mode \(\Delta \varphi\) depends on \(-e_{z_1}\).

Since \({\lim \limits _{t \rightarrow \infty }}\alpha _w,{\lim \limits _{t \rightarrow \infty }}M_{b_Z}\) and \({\lim \limits _{t \rightarrow \infty }}M_{q1}\) exist, (20) is obtained. \(\square\)

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Zhong, S., Huang, Y., Chen, S. et al. A novel ADRC-based design for a kind of flexible aerocraft. Control Theory Technol. 19, 35–48 (2021). https://doi.org/10.1007/s11768-020-00030-y

Download citation

  • Received:

  • Revised:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s11768-020-00030-y

Keywords

Navigation