Skip to main content

Backstepping-Based Nonlinear RISE Feedback Control for an Underactuated Quadrotor UAV Without Linear Velocity Measurements

  • Chapter
  • First Online:
New Developments and Advances in Robot Control

Part of the book series: Studies in Systems, Decision and Control ((SSDC,volume 175))

Abstract

In this paper, a robust three dimensional output feedback control problem is proposed for a 6-degrees-of-freedom model of a quadrotor unmanned aerial vehicle (UAV) to track a bounded and sufficiently smooth reference trajectory in the presence of slowly varying force disturbances. Due to the underactuation structure of the UAV, a nonlinear output feedback controller based on the robust integral of the sign error signal (RISE) mechanism is first designed for the translational dynamics to ensure position reference tracking without velocity measurement. The angular velocity is then regarded as intermediate control signal for the rotational dynamics to fulfill the task of attitude angle reference tracking. The torque input is designed taking full advantage of the smooth exact differentiator that circumvents derivatives computation of virtual controls, the backstepping technique is then judiciously modified to allow the use of the RISE control technique to compensate for the external disturbances. The proposed controller yields semi-global asymptotic stability tracking despite the added disturbances in the dynamics. Simulation results are shown to demonstrate the proposed approach.

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

Access this chapter

eBook
USD 16.99
Price excludes VAT (USA)
  • Available as EPUB and PDF
  • Read on any device
  • Instant download
  • Own it forever
Hardcover Book
USD 109.99
Price excludes VAT (USA)
  • Durable hardcover edition
  • Dispatched in 3 to 5 business days
  • Free shipping worldwide - see info

Tax calculation will be finalised at checkout

Purchases are for personal use only

Institutional subscriptions

References

  • Abdessameud, A., & Tayebi, A. (2010). Global trajectory tracking control of VTOL-UAVs without linear velocity measurements. Automatica, 46(6), 1053–1059.

    Article  MathSciNet  MATH  Google Scholar 

  • Benallegue, A., Mokhtari, A., & Fridman, L. (2008). High-order sliding-mode observer for a quadrotor UAV. International Journal of Robust and Nonlinear Control, 18(4–5), 427–440.

    Article  MathSciNet  MATH  Google Scholar 

  • Bertrand, S., Guénard, N., Hamel, T., Piet-Lahanier, H., & Eck, L. (2011). A hierarchical controller for miniature VTOL UAVs: Design and stability analysis using singular perturbation theory. Control Engineering Practice, 19(10), 1099–1108.

    Article  Google Scholar 

  • Cai, Z., de Queiroz, M. S., & Dawson, D. M. (2006). Robust adaptive asymptotic tracking of nonlinear systems with additive disturbance. IEEE Transactions on Automatic Control, 51(3), 524–529.

    Article  MathSciNet  MATH  Google Scholar 

  • Castillo, P., Dzul, A., & Lozano, R. (2004). Real-time stabilization and tracking of a four-rotor mini rotorcraft. IEEE Transactions on Control Systems Technology, 12(4), 510–516.

    Article  Google Scholar 

  • Ciulkin, M., Pawluszewicz, E., Kaparin, V., & Kotta, U. (2015). Input-output linearization by dynamic output feedback on homogeneous time scales. In 20th International Conference on Methods and Models in Automation and Robotics (MMAR), Miȩdzyzdroje, Poland (pp. 477–482).

    Google Scholar 

  • Clarke, F. H. (1990). Optimization and nonsmooth analysis. Philadelphia: SIAM.

    Book  MATH  Google Scholar 

  • Cunha, R., Cabecinhas, D., & Silvestre, C. (2009). Nonlinear trajectory tracking control of a quadrotor vehicle. In Proceedings of ECC 2009 – 10th European Control Conference, Budapest, Hungary.

    MATH  Google Scholar 

  • De Queiroz, M., Hu, J., Dawson, D., Burg, T., & Donepudi, S. (1997). Adaptive position/force control of robot manipulators without velocity measurements: Theory and experimentation. IEEE Transactions on Systems, Man, and Cybernetics, 27-B(5), 796–809.

    Article  Google Scholar 

  • Ding, X., Wang, X., Yu, Y., & Zha, C. (2017). Dynamics modeling and trajectory tracking control of a quadrotor unmanned aerial vehicle. Journal of Dynamic Systems, Measurement, and Control, 139, 1–11.

    Google Scholar 

  • Do, K. D. (2015). Path-tracking control of stochastic quadrotor aircraft in three-dimensional space. Journal of Dynamic Systems, Measurement, and Control, 137(10), 1–11.

    Article  Google Scholar 

  • Filippov, A. (1964). Differential equations with discontinuous right-hand side. American Mathematical Society Translations, 42(2), 199–231.

    MATH  Google Scholar 

  • Garcia, R., Rubio, F., & Ortega, M. (2012). Robust PID control of the quadrotor helicopter. IFAC Proceedings Volumes, 45(3), 229–234.

    Article  Google Scholar 

  • Ghommam, J., Charland, G., & Saad, M. (2015). Three-dimensional constrained tracking control via exact differentiation estimator of a quadrotor helicopter. Asian Journal of Control, 17(3), 1093–1103.

    Article  MathSciNet  MATH  Google Scholar 

  • Huang, M., Xian, B., Diao, C., Yang, K., & Feng, Y. (2010). Adaptive tracking control of underactuated quadrotor unmanned aerial vehicles via backstepping. In Proceedings of the IEEE American control Conference (ACC), Baltimore, MD, USA (pp. 2076–2081).

    Google Scholar 

  • Ilcev, D. S., & Sibiya, S. S. (2015). Weather observation via stratospheric platform stations. In 2015 IST-Africa Conference, Lilongwe, Malawi (pp. 1–12).

    Google Scholar 

  • Kendoul, F. (2009). Nonlinear hierarchical flight controller for unmanned rotorcraft: Design, stability, and experiments. Journal of Guidance Control and Dynamics, 32(6), 1954–1958.

    Article  Google Scholar 

  • Kim, H. J., & Shim, D. H. (2003). A flight control system for aerial robots: Algorithms and experiments. Control Engineering Practice, 11(2), 1389–1400.

    Article  Google Scholar 

  • Lee, D. B., Burg, T. C., Xian, B., & Dawson, D. M. (2007). Output feedback tracking control of an underactuated quad-rotor UAV. In Proceedings of the 2007 American Control Conference, New York, USA (pp. 1775–1780).

    Google Scholar 

  • Levant, A. (2005). Quasi-continuous high-order sliding-mode controllers. IEEE Transactions on Automatic Control, 50(11), 1812–1816.

    Article  MathSciNet  MATH  Google Scholar 

  • Madani, T., & Benallegue, A. (2007). Sliding mode observer and backstepping control for a quadrotor unmanned aerial vehicles. In 2007 American Control Conference, New York, USA (pp. 5887–5892).

    Google Scholar 

  • Magsino, E. R., Chua, J. R. B., Chua, L. S., de Guzman, C. M., & Gepaya, J. V. L. (2016). A rapid screening algorithm using a quadrotor for crack detection on bridges. In 2016 IEEE Region 10 Conference (TENCON), Singapore (pp. 1829–1833).

    Google Scholar 

  • Ortiz, J. P., Minchala, L. I., & Reinoso, M. J. (2016). Nonlinear robust h-infinity pid controller for the multivariable system quadrotor. IEEE Latin America Transactions, 14(3), 1176–1183.

    Article  Google Scholar 

  • Paden, B., & Sastry, S. (1987). A calculus for computing filippov’s differential inclusion with application to the variable structure control of robot manipulators. IEEE Transactions on Circuits System, 34(1), 73–82.

    Article  MathSciNet  MATH  Google Scholar 

  • Rinaldi, F., Chiesa, S., & Quagliotti, F. (2013). Linear quadratic control for quadrotors UAVs dynamics and formation flight. Journal of Intelligent and Robotic Systems, 70(1–4), 203–220.

    Article  Google Scholar 

  • Salih, A. L., Moghavvemi, M., Mohamed, H. A. F., & Gaeid, K. S. (2010). Modelling and PID controller design for a quadrotor unmanned air vehicle. In 2010 IEEE International Conference on Automation, Quality and Testing, Robotics (AQTR), Cluj-Napoca, Romania (Vol. 1, pp. 1–5).

    Google Scholar 

  • Van Den Eijnden, S. (2017). Cascade based tracking control of quadrotors. Master’s thesis, Eindhoven University of Technology.

    Google Scholar 

  • Wang, L., & Jia, H. (2014). The trajectory tracking problem of quadrotor UAV: Global stability analysis and control design based on the cascade theory. Asian Journal of Control, 16(2), 574–588.

    Article  MathSciNet  MATH  Google Scholar 

  • Xian, B., Dawson, D. M., de Queiroz, M. S., & Chen, J. (2004). A continuous asymptotic tracking control strategy for uncertain nonlinear systems. IEEE Transactions on Automatic Control, 49(7), 1206–1211.

    Article  MathSciNet  MATH  Google Scholar 

  • Xian, B., & Zhang, Y. (2017). A new smooth robust control design for uncertain nonlinear systems with non-vanishing disturbances. International Journal of Control, 89(6), 1285–1302.

    Article  MathSciNet  MATH  Google Scholar 

  • Yao, Z. (2017). Nonlinear robust adaptive hierarchical sliding mode control approach for quadrotors. International Journal of Robust and Nonlinear Control, 27(6), 925–941.

    Article  MathSciNet  MATH  Google Scholar 

  • Zou, Y. (2017). Nonlinear hierarchical control for quad-rotors with rotation matrix. International Journal of Control, 90(7), 1308–1318.

    Article  MathSciNet  MATH  Google Scholar 

Download references

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Jawhar Ghommam .

Editor information

Editors and Affiliations

Appendix

Appendix

For the clarity of the stability proof, we first write the closed loop-system by substituting the control input (15.35) into the dynamics (15.34), yields

$$\displaystyle \begin{aligned} \begin{array}{rcl} \left[ \begin{array}{c} \dot{\boldsymbol{\eta}} \\ \dot{\mathbf{e}}_\psi \\ \end{array} \right]&=&\left[ \begin{array}{c} (k_2+1)({\mathbf{r}}_f-\boldsymbol{\eta})-\mathbf{S}(\omega)\boldsymbol{\eta}+\widetilde{\mathbf{M}}+{\mathbf{M}}_d-K_1\text{Sgn}({\mathbf{e}}_p+{\mathbf{e}}_f)-{\mathbf{e}}_p \\ 0 \\ \end{array} \right]\notag\\ &&+\overline{\mathbf{B}}_C\boldsymbol{z} {} \end{array} \end{aligned} $$
(15.40)
$$\displaystyle \begin{aligned} \begin{array}{rcl} {\mathbf{I}}_f\dot{\mathbf{r}}_a&=&\frac{1}{k_3}\Big(\mathbf{D}-K_2\text{Sng}(\boldsymbol{z})-k_3(k_3{\mathbf{I}}_f+{\mathbf{I}}_3){\mathbf{r}}_a\Big)+{\mathbf{I}}_f({\mathbf{r}}_a-\boldsymbol{z}){} \end{array} \end{aligned} $$
(15.41)

where we have used the following relationship:

$$\displaystyle \begin{aligned} {\mathbf{B}}_C\left[ \begin{array}{c} \boldsymbol{z} \\ 0\\ \end{array} \right]=\left[ \begin{array}{c} -\mathbf{S}(\boldsymbol{\delta}) \boldsymbol{z} \\ T_{\boldsymbol{z}}(\boldsymbol{\varTheta})\boldsymbol{z}\\ \end{array} \right]=\underbrace{\left[ \begin{array}{c} -\mathbf{S}(\boldsymbol{\delta}) \\ T_{\boldsymbol{z}}(\boldsymbol{\varTheta})\\ \end{array} \right]}_{\overline{\mathbf{B}}_C}\boldsymbol{z} \end{aligned}$$

For the closed-loop system (15.41) and (15.40), we consider the Lyapunov function candidate function V  which is Lipschitz positive definite function defined as:

$$\displaystyle \begin{aligned} V(\boldsymbol{y},t)=V_1+V_2+P+Q \end{aligned} $$
(15.42)

where we define the functions V 1 and V 2 as follows

$$\displaystyle \begin{aligned} \begin{array}{rcl} V_1 &\displaystyle =&\displaystyle \frac{m}{2}{\mathbf{e}}_p^\top {\mathbf{e}}_p+\frac{1}{2}{\mathbf{e}}_f^\top {\mathbf{e}}_f+\frac{1}{2} {\mathbf{r}}_f^\top {\mathbf{r}}_f+\frac{1}{2}\boldsymbol{\chi}^\top \boldsymbol{\chi}{} \end{array} \end{aligned} $$
(15.43)
$$\displaystyle \begin{aligned} \begin{array}{rcl} V_2 &\displaystyle =&\displaystyle \frac{1}{2}\boldsymbol{z}^\top {\mathbf{I}}_f \boldsymbol{z}+\frac{1}{2}{\mathbf{r}}_a^\top {\mathbf{I}}_f {\mathbf{r}}_a{}\vspace{-3pt} \end{array} \end{aligned} $$
(15.44)

Let y ∈ R 20 be defined as

$$\displaystyle \begin{aligned} \begin{array}{rcl} \boldsymbol{y}&\displaystyle =&\displaystyle [\boldsymbol{X}^\top,\sqrt{P(t)},\sqrt{Q(t)}]^\top{}\\ \boldsymbol{X}&\displaystyle =&\displaystyle [{\mathbf{e}}_p^\top,{\mathbf{e}}_f^\top,{\mathbf{r}}_f^\top,\boldsymbol{\eta}^\top,e_\psi,\boldsymbol{z}^\top,{\mathbf{r}}_a^\top]\notag \end{array} \end{aligned} $$
(15.45)

and the auxiliary function \(P\in \mathbb {R}\) and \(Q \in \mathbb {R}\) are defined as the Filippov solutions (Filippov 1964) to the following differential equations respectively:

$$\displaystyle \begin{aligned} \begin{array}{rcl} \dot{P} &\displaystyle =&\displaystyle -\boldsymbol{\eta}^\top \Big({\mathbf{M}}_d-K_1\text{Sng}({\mathbf{e}}_p+{\mathbf{e}}_f)\Big) \end{array} \end{aligned} $$
(15.46)
$$\displaystyle \begin{aligned} \begin{array}{rcl} \dot{Q} &\displaystyle =&\displaystyle \frac{{\mathbf{r}}_a^\top}{k_3}\Big(\mathbf{D}-K_2\text{Sgn}(\boldsymbol{z})\Big) \end{array} \end{aligned} $$
(15.47)

with initial conditions P(t 0) and Q(t 0) are given as follows

$$\displaystyle \begin{aligned} \begin{array}{rcl} P(t_0) &\displaystyle =&\displaystyle \sum_{i=1}^3 K_1|e_{pi}(0)| -{\mathbf{e}}_p^\top(0)M_d(0) {} \end{array} \end{aligned} $$
(15.48)
$$\displaystyle \begin{aligned} \begin{array}{rcl} Q(t_0) &\displaystyle =&\displaystyle \frac{1}{k_3}\Big(\sum_{i=1}^3 K_2|z_{i}(0)|-\boldsymbol{z}(0)^\top \mathbf{D}(0) \Big){} \end{array} \end{aligned} $$
(15.49)

with the subscript i = 1, …, n denotes the i-th element of the vectors e p(t) and z(t). The gain control K 1 and K 2 are chosen according to the sufficient condition (15.3) underlined in Lemma 2.

It can be verified that the Lyapunov function candidate defined in (15.42) can be bounded as :

$$\displaystyle \begin{aligned} U_1(\boldsymbol{y})\leq V(\boldsymbol{y},t) \leq U_2(\boldsymbol{y}) \end{aligned} $$
(15.50)

where \(U_1 : \mathbb {R}^{20} \rightarrow \mathbb {R}\) and \(U_2 : \mathbb {R}^{20} \rightarrow R\) are positive definite functions defined as \(U_1(\boldsymbol {y})=\min \{\frac {1}{2},\frac {m}{2}\}\| \boldsymbol {y}\|{ }^2\) and \(U_2(\boldsymbol {y})=\max \{1,\frac {m}{2}\}\| \boldsymbol {y}\|{ }^2\). Under Filippov’s framework, strong stability of the closed-loop system will be established in the following.

The time derivative of (15.42) along the Filippov trajectories exists almost everywhere (a.e), for almost all t ∈ [t 0, t], one can get:

(15.51)

where ∂V  is the generalized gradient of V  (Clarke 1990). Since V  is continuously differentiable with respect to y, then one has

$$\displaystyle \begin{aligned} \dot{\widetilde{V}}\subset \nabla V^\top \varPsi[\dot{\mathbf{e}}_p^\top, \dot{\mathbf{e}}_f^\top, \dot{\mathbf{r}}_f^\top, \dot{\boldsymbol{\eta}}^\top, \dot{e}_\psi, \frac{1}{2}P^{\frac{1}{2}} \dot{P}, \frac{1}{2}Q^{\frac{1}{2}} \dot{Q}, 1] \end{aligned} $$
(15.52)

where \(\nabla V=\varPsi [{\mathbf {e}}_p^\top , {\mathbf {e}}_f^\top , {\mathbf {r}}_f^\top , \boldsymbol {\eta }^\top , e_\psi ,\boldsymbol {z}^{\top {\mathbf {I}}}_f,{\mathbf {r}}_a^\top {\mathbf {I}}_f, 2P^{-\frac {1}{2}},2Q^{-\frac {1}{2}}]^\top \). Substituting (15.9), (15.15), (15.24), (15.40) and (15.41) into (15.52) and using the calculus for Ψ[.] from Paden and Sastry (1987), yields:

$$\displaystyle \begin{aligned} \begin{array}{rcl} \dot{\widetilde{V}} &\displaystyle \subset&\displaystyle {\mathbf{e}}_p^{\top} (\boldsymbol{\eta}-{\mathbf{e}}_p-{\mathbf{r}}_f-\boldsymbol{\delta})+(1-m){\mathbf{e}}_p^{\top}{\mathbf{R}}^{\top} \dot{\mathbf{p}}_d-{\mathbf{e}}_f^{\top} {\mathbf{e}}_f+{\mathbf{e}}_f^{\top} {\mathbf{r}}_f-{\mathbf{r}}_f^{\top} {\mathbf{r}}_f \\ &\displaystyle &\displaystyle -(k_2+1){\mathbf{r}}^{\top} \boldsymbol{\eta} +{\mathbf{r}}_f^{\top} {\mathbf{e}}_p-{\mathbf{r}}_f^{\top} {\mathbf{e}}_f-(k_2+1)\boldsymbol{\eta}^{\top} \boldsymbol{\eta}+\boldsymbol{\eta}^{\top} \widetilde{\mathbf{M}}\\ &\displaystyle &\displaystyle +\boldsymbol{\eta}^{\top} \Big({\mathbf{M}}_d-K_1\varPsi[\text{Sng}({\mathbf{e}}_p+{\mathbf{e}}_f)]\Big)-\boldsymbol{\eta}^{\top} {\mathbf{e}}_p +(k_2+1)\boldsymbol{\eta}^{\top} {\mathbf{r}}_f\\ &\displaystyle &\displaystyle +\boldsymbol{\eta}^{\top} \overline{\mathbf{B}}_C\boldsymbol{z}-\Big({\mathbf{M}}_d-K_1\varPsi[\text{Sng}({\mathbf{e}}_p+{\mathbf{e}}_f)]\Big)-k_3\boldsymbol{z}^{\top} {\mathbf{I}}_f \boldsymbol{z}-{\mathbf{r}}_a^{\top} {\mathbf{r}}_a\\ &\displaystyle &\displaystyle +\frac{1}{k_3}{\mathbf{r}}_a^{\top}\Big(\mathbf{D}-K_2\varPsi[\text{Sng}(\boldsymbol{z})]\Big)-\frac{1}{k_3}{\mathbf{r}}_a^{\top}\Big(\mathbf{D}-K_2\varPsi[\text{Sng}(\boldsymbol{z})]\Big){} \end{array} \end{aligned} $$
(15.53)

where Ψ[Sng(x)] = SGN(x) such that ∀x = [x 1, x 2, x 3], the set valued map SNG(x i) = 1 if x i > 0, [−1, 1] if x i = 0 and − 1 if x i < 0 for i = 1, 2, 3. Using the fact that the set in (15.53) reduces to a scalar equality since the right hand side is continuous a.e, the following upper bound can be obtained on \(\dot {V}\):

(15.54)

Denote by \(\gamma =\| \boldsymbol {\delta } \|+(m+1)\| \dot {\mathbf {p}}_d\|\), using the Mean Value Theorem (De Queiroz et al. 1997), the function \(\widetilde {\mathbf {M}}\) can be upper bounded as \(\| \widetilde {\mathbf {M}} \|\leq \| \varphi (\| \boldsymbol {X} \|)\| \| \boldsymbol {X} \|\), then using the young’s inequality yields:

$$\displaystyle \begin{aligned} \begin{array}{rcl} \| -{\mathbf{e}}_p^\top\big(\boldsymbol{\delta}+(m-1){\mathbf{R}}^\top \dot{\mathbf{p}}_d\big)\| &\displaystyle \leq&\displaystyle \gamma \| {\mathbf{e}}_p\|\leq \frac{\gamma}{\| \boldsymbol{X} \|}\| \boldsymbol{X} \|{}^2 {} \end{array} \end{aligned} $$
(15.55)
$$\displaystyle \begin{aligned} \begin{array}{rcl} \| \widetilde{\mathbf{M}} \| \| \boldsymbol{\eta} \| &\displaystyle \leq&\displaystyle \frac{\| \varphi (\| \boldsymbol{X} \|)\|{}^2}{4k_2}+k_2\| \boldsymbol{\eta}\|{}^2 {}\vspace{-3pt} \end{array} \end{aligned} $$
(15.56)

substituting (15.55) and (15.56) into (15.54), we obtain

(15.57)

It follows through the choice of ς as any given positive constant such that ∥X2 ≥ ς 2, the expression in (15.57) can be upper bounded as:

(15.58)

From (15.58), it can be concluded that for some positive constant C 0, we have:

(15.59)

It can be concluded from inequalities (15.50) and (15.59) that V (y, t) ∈ L , therefore e p, e f, r f, η, z, r a, P and \(Q \in \mathcal {L}_\infty \). From (15.13), (15.14) and (15.15), a straightforward conclusion can be drawn to show that \(\dot {\mathbf {e}}_f, \mathbf {q}, \dot {\mathbf {q}} \in \mathcal {L}_\infty \). From the fact that the desired trajectory is sufficiently smooth then using (15.8) we can conclude that p and \(\dot {\mathbf {p}} \in \mathcal {L}_\infty \). From (15.28) and (15.36) we know that \(\overline {\mathbf {u}}_1\) and \({\mathbf {u}}_2 \in \mathcal {L}_\infty \). Given these boundedness statements, it is clear from that the time derivative of U(X) = γX2 is such that \(\dot {U}(\boldsymbol {X})\in \mathcal {L}_\infty \), which also implies that U(X) is uniformly continuous. Define the region \(\mathcal {S}_{\mathcal {D}}\) as follows:

$$\displaystyle \begin{aligned} \mathcal{S}_{\mathcal{D}}=\left\{\boldsymbol{y} \in \mathcal{D}~|\quad U_2(\boldsymbol{y})< \min\{\frac{1}{2},\frac{m}{2}\} \varphi^{-1}\Bigg(2\sqrt{\frac{k_2(\varsigma-\gamma)}{\varsigma}}\Bigg)^2 \right\} \end{aligned} $$
(15.60)

The region of attraction in \(\mathcal {S}_{\mathcal {D}}\) can be made arbitrary large by appropriately tuning the control gain k 2. As a matter of fact, the region of attraction is calculated based on the set given in (15.60) as follows:

$$\displaystyle \begin{aligned} \| \boldsymbol{y}(t_0) \| \leq \sqrt{\frac{\min\{\frac{1}{2},\frac{m}{2}\}}{\max\{1,\frac{m}{2}\}}}\varphi^{-1}\Bigg(2\sqrt{\frac{k_2(\varsigma-\gamma)}{\varsigma}}\Bigg) \end{aligned} $$
(15.61)

which can be rearranged as in (15.38). Based on the definition of the vector state y in (15.45) and the initial conditions defined in (15.16) and (15.17) and (15.48) and (15.49), an explicit expression for ∥y(t 0)∥ can be derived as in (15.39).

From (15.59) it can be concluded that C 0X2 → 0 as time goes to infinity, \(\forall \boldsymbol {y}\in \mathcal {S}_{\mathcal {D}}\), it is then obvious to show form the definition of X that e p, e f, r f, e ψ, η, z, r a → 0 as t →, \(\forall \boldsymbol {y}\in \mathcal {S}_{\mathcal {D}}\). This further implies that e v → [0, 0, −δ] as t →, consequently the velocity tracking error \(m\boldsymbol {\upsilon }-{\mathbf {R}}^\top \dot {\boldsymbol {p}}_d \rightarrow 0\) as t →. This completes the proof.

Rights and permissions

Reprints and permissions

Copyright information

© 2019 Springer Nature Singapore Pte Ltd.

About this chapter

Check for updates. Verify currency and authenticity via CrossMark

Cite this chapter

Ghommam, J., Luque-Vega, L.F., Saad, M. (2019). Backstepping-Based Nonlinear RISE Feedback Control for an Underactuated Quadrotor UAV Without Linear Velocity Measurements. In: Derbel, N., Ghommam, J., Zhu, Q. (eds) New Developments and Advances in Robot Control. Studies in Systems, Decision and Control, vol 175. Springer, Singapore. https://doi.org/10.1007/978-981-13-2212-9_15

Download citation

Publish with us

Policies and ethics