Skip to main content
Log in

An integral method for parameter identification of a nonlinear robot subject to quantization error

  • Original Paper
  • Published:
Nonlinear Dynamics Aims and scope Submit manuscript

Abstract

This paper focuses on the parameter identification of a class of nonlinear mechanical dynamic systems subject to feedback quantization errors. A second-order integral method for the parameter identification of a nonlinear two-degrees-of-freedom robotic manipulator is studied. To obtain accurate parameter estimates, an optimal excitation trajectory subject to state constraints is proposed. The designated excitation trajectory is able to minimize estimation uncertainties, providing the controlled system trajectory reaches the desired optimal excitation trajectory. Moreover, for most of the nonlinear robotic systems, the output measurements are acquired from the encoders subject to apparent quantization error. Directly differentiating the position to estimate the velocity and acceleration trajectories is unrealistic and unsuitable due to the amplified quantization errors will cause considerable estimation errors. Instead, the finite Fourier series is introduced to perform optimal estimation of the position, velocity, and acceleration histories to capture the dominant frequencies of the system response. A weighted least-squares solution is further proposed to improve parameter identification precision. The comparative studies reveal that the integral method has superior parameter estimation performance than the direct difference model and the filtered regression model in the existing literature. Experimental simulations illustrate the effectiveness of the presented method.

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

Similar content being viewed by others

Data availability

The manuscript has no associated data.

References

  1. Bahloul, A., Tliba, S., Chitour, Y.: Dynamic parameters identification of an industrial robot with and without payload. Ifac-Papersonline 51(15), 443–448 (2018)

    Article  Google Scholar 

  2. Calafiore, G., Indri, M., Bona, B.: Robot dynamic calibration: optimal excitation trajectories and experimental parameter estimation. J. Robot. Syst. 18(2), 55–68 (2001)

    Article  Google Scholar 

  3. Calanca, A., Capisani, L.M., Ferrara, A., Magnani, L.: Mimo closed loop identification of an industrial robot. IEEE Trans. Control Syst. Technol. 19(5), 1214–1224 (2010)

    Article  Google Scholar 

  4. Chatzilygeroudis, K., Rama, R., Kaushik, R., Goepp, D., Vassiliades, V., Mouret, J.B.: Black-box data-efficient policy search for robotics. In: 2017 IEEE/RSJ International Conference on Intelligent Robots and Systems (IROS), pp. 51–58. IEEE (2017)

  5. Chen, C.L., Peng, C.C.: Control of a perturbed chaotic system by using a trajectory trapping strategy. Nonlinear Dyn. 69(4), 2105–2115 (2012)

    Article  MathSciNet  Google Scholar 

  6. Chen, L.H., Peng, C.C.: Extended backstepping sliding controller design for chattering attenuation and its application for servo motor control. Appl. Sci. 7(3), 220 (2017)

    Article  Google Scholar 

  7. Deng, J., Shang, W., Zhang, B., Zhen, S., Cong, S.: Dynamic model identification of collaborative robots using a three-loop iterative method. In: 2021 6th IEEE International Conference on Advanced Robotics and Mechatronics (ICARM), pp. 937–942. IEEE (2021)

  8. Dolinskỳ, K., Čelikovskỳ, S.: Application of the method of maximum likelihood to identification of bipedal walking robots. IEEE Trans. Control Syst. Technol. 26(4), 1500–1507 (2017)

    Article  Google Scholar 

  9. Gautier, M.: Dynamic identification of robots with power model. In: Proceedings of International Conference on Robotics and Automation, vol. 3, pp. 1922–1927. IEEE (1997)

  10. Gautier, M., Janot, A., Vandanjon, P.: A new closed-loop output error method for parameter identification of robot dynamics. IEEE Trans. Control Syst. Technol. 21(2), 428–444 (2012)

    Article  Google Scholar 

  11. Gautier, M., Janot, A., Vandanjon, P.O.: Didim: A new method for the dynamic identification of robots from only torque data. In: 2008 IEEE International Conference on Robotics and Automation, pp. 2122–2127. IEEE (2008)

  12. Gautier, M., Khalil, W., Restrepo, P.: Identification of the dynamic parameters of a closed loop robot. In: Proceedings of 1995 IEEE International Conference on Robotics and Automation, vol. 3, pp. 3045–3050. IEEE (1995)

  13. Guého, D., Singla, P., Majji, M., Melton, R.G.: Filtered integral formulation of the sparse model identification problem. J. Guid. Control Dyn. 45(2), 232–247 (2022)

    Article  Google Scholar 

  14. Han, S.H., Yoo, A., Yoon, S.W., Yoon, Y.D.: A simple method for identifying mechanical parameters based on integral calculation. J. Power Electron. 16(4), 1387–1395 (2016)

    Article  Google Scholar 

  15. Han, Y., Wu, J., Liu, C., Xiong, Z.: An iterative approach for accurate dynamic model identification of industrial robots. IEEE Trans. Robot. 36(5), 1577–1594 (2020)

    Article  Google Scholar 

  16. He, W., Ge, W., Li, Y., Liu, Y.J., Yang, C., Sun, C.: Model identification and control design for a humanoid robot. IEEE Trans. Syst. Man Cybern. Syst. 47(1), 45–57 (2016)

    Article  Google Scholar 

  17. Horvath, S., Neuner, H.: System identification of a robot arm with extended kalman filter and artificial neural networks. J. Appl. Geod. 13(2), 135–150 (2019)

    Article  Google Scholar 

  18. Jia, J., Zhang, M., Zang, X., Zhang, H., Zhao, J.: Dynamic parameter identification for a manipulator with joint torque sensors based on an improved experimental design. Sensors 19(10), 2248 (2019)

    Article  Google Scholar 

  19. Jin, J., Gans, N.: Parameter identification for industrial robots with a fast and robust trajectory design approach. Robot. Comput. Integr. Manuf. 31, 21–29 (2015)

    Article  Google Scholar 

  20. Juang, J.N.: Applied System Identification. Prentice-Hall Inc., Hoboken (1994)

    Google Scholar 

  21. Juang, J.N.: Continuous-time bilinear system identification. Nonlinear Dyn. 39(1), 79–94 (2005)

    Article  MathSciNet  Google Scholar 

  22. Juang, J.N., Pappa, R.S.: An eigensystem realization algorithm for modal parameter identification and model reduction. J. Guid. Control Dyn. 8(5), 620–627 (1985)

    Article  Google Scholar 

  23. Juang, J.N., Phan, M.Q.: Identification and Control of Mechanical Systems. Cambridge University Press, Cambridge (2001)

    Book  Google Scholar 

  24. Lee, T., Wensing, P.M., Park, F.C.: Geometric robot dynamic identification: a convex programming approach. IEEE Trans. Robot. 36(2), 348–365 (2019)

    Article  Google Scholar 

  25. Li, Y., Li, Y., Zhu, M., Xu, Z., Mu, D.: A nonlinear momentum observer for sensorless robot collision detection under model uncertainties. Mechatronics 78, 102603 (2021)

    Article  Google Scholar 

  26. Li, Y.R., Peng, C.C.: Encoder position feedback based indirect integral method for motor parameter identification subject to asymmetric friction. Int. J. Non-Linear Mech. 152, 104386 (2023)

    Article  Google Scholar 

  27. Liu, G., Iagnemma, K., Dubowsky, S., Morel, G.: A base force/torque sensor approach to robot manipulator inertial parameter estimation. In: Proceedings. 1998 IEEE International Conference on Robotics and Automation (Cat. No. 98CH36146), vol. 4, pp. 3316–3321. IEEE (1998)

  28. Liu, X., Yang, X., Yin, S.: Nonlinear system identification with robust multiple model approach. IEEE Trans. Control Syst. Technol. 28(6), 2728–2735 (2019)

    Article  Google Scholar 

  29. Lopez-Sanchez, I., Montoya-Cháirez, J., Pérez-Alcocer, R., Moreno-Valenzuela, J.: Experimental parameter identifications of a quadrotor by using an optimized trajectory. IEEE Access 8, 167355–167370 (2020)

    Article  Google Scholar 

  30. Mablekos-Alexiou, A., Da Cruz, L., Bergeles, C.: Friction-inclusive modeling of sliding contact transmission systems in robotics. IEEE Trans. Robot. 37(4), 1252–1267 (2021)

    Article  Google Scholar 

  31. Majji, M., Juang, J.N., Junkins, J.L.: Observer/kalman-filter time-varying system identification. J. Guid. Control Dyn. 33(3), 887–900 (2010)

    Article  Google Scholar 

  32. Majji, M., Juang, J.N., Junkins, J.L.: Time-varying eigensystem realization algorithm. J. Guid. Control Dyn. 33(1), 13–28 (2010)

    Article  Google Scholar 

  33. MATLAB: version 9.6 (R2019a). The MathWorks Inc., Natick, MA, USA (2019)

  34. Moreno-Valenzuela, J., Aguilar-Avelar, C.: Motion Control of Underactuated Mechanical Systems, vol. 1. Springer, Berlin (2018)

    Book  Google Scholar 

  35. Moreno-Valenzuela, J., Miranda-Colorado, R., Aguilar-Avelar, C.: A matlab-based identification procedure applied to a two-degrees-of-freedom robot manipulator for engineering students. Int. J. Electr. Eng. Educ. 54(4), 319–340 (2017)

    Article  Google Scholar 

  36. Naozuka, G.T., Rocha, H.L., Silva, R.S., Almeida, R.C.: Sindy-sa framework: enhancing nonlinear system identification with sensitivity analysis. Nonlinear Dyn. 110, 2589–2609 (2022)

    Article  Google Scholar 

  37. Peng, C.C., Li, Y.R.: Parameters identification of nonlinear Lorenz chaotic system for high-precision model reference synchronization. Nonlinear Dyn. 108(2), 1733–1754 (2022)

    Article  Google Scholar 

  38. Peng, C.C., Lin, Y.I.: Dynamics modeling and parameter identification of a cooling fan system. In: 2018 IEEE International Conference on Advanced Manufacturing (ICAM), pp. 257–260. IEEE (2018)

  39. Reyes, F., Kelly, R.: Experimental evaluation of identification schemes on a direct drive robot. Robotica 15(5), 563–571 (1997)

    Article  Google Scholar 

  40. Swevers, J., Ganseman, C., De Schutter, J., Van Brussel, H.: Experimental robot identification using optimised periodic trajectories. Mech. Syst. Signal Process. 10(5), 561–577 (1996)

    Article  Google Scholar 

  41. Swevers, J., Ganseman, C., Tukel, D.B., De Schutter, J., Van Brussel, H.: Optimal robot excitation and identification. IEEE Trans. Robot. Autom. 13(5), 730–740 (1997)

    Article  Google Scholar 

  42. Swevers, J., Verdonck, W., De Schutter, J.: Dynamic model identification for industrial robots. IEEE Control Syst. Mag. 27(5), 58–71 (2007)

    Article  MathSciNet  Google Scholar 

  43. Vicente, B.A.H., James, S.S., Anderson, S.R.: Linear system identification versus physical modeling of lateral-longitudinal vehicle dynamics. IEEE Trans. Control Syst. Technol. 29(3), 1380–1387 (2020)

    Article  Google Scholar 

  44. Wu, W., Zhu, S., Wang, X., Liu, H.: Closed-loop dynamic parameter identification of robot manipulators using modified Fourier series. Int. J. Adv. Robot. Syst. 9(1), 29 (2012)

    Article  Google Scholar 

Download references

Funding

The authors have not disclosed any funding.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Chao-Chung Peng.

Ethics declarations

Conflict of interest

The authors declare that they have no conflict of interest.

Additional information

Publisher's Note

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

This work was supported by the Ministry of Science and Technology under Grant No. MOST 111-2221-E-006-170.

Appendices

Appendix A Design strategy of integral factors

To illustrate how the proposed integral operator based regression model suppresses the influence of the initial condition, the first element of the matrix differential equation \(\ddot{{\textbf {Y}}}_{2,2}(t) = -(\lambda _1+\lambda _2)\dot{{\textbf {Y}}}_{2,2}(t) - \lambda _1\lambda _2{\textbf {Y}}_{2,2}(t) + \lambda _1^2\varvec{\varphi }_2(t)\) in (42) is considered:

$$\begin{aligned} \ddot{y}_{22,1}(t) = -k_2{\dot{y}}_{22,1}(t) - k_1y_{22,1}(t) + \lambda ^2\theta _1(t) \end{aligned}$$
(95)

where \(k_1=\lambda _1\lambda _2\), \(k_2=\lambda _1+\lambda _2\), and \(y_{22,1}(t)\) is the first element of the matrix \({\textbf {Y}}_{2,2}(t)\). The associated initial conditions are \(y_{22,1}(0)=0\) and \({\dot{y}}_{22,1}(0)=\lambda _1\theta _1(0)-{\dot{\theta }}_1(0)\). Define the auxiliary state variables \(\eta _1(t)=y_{22,1}(t)\) and \(\eta _2(t)={\dot{y}}_{22,1}(t)\), the state-space equation in the controllable form can be expressed by

$$\begin{aligned} \begin{bmatrix} {\dot{\eta }}_1(t)\\ {\dot{\eta }}_2(t) \end{bmatrix}= & {} \begin{bmatrix} 0 &{}\quad 1\\ -k_1 &{}\quad -k_2 \end{bmatrix} \begin{bmatrix} {\eta }_1(t)\\ {\eta }_2(t) \end{bmatrix} + \begin{bmatrix} 0\\ \lambda ^2 \end{bmatrix}\theta _1(t)\nonumber \\ y(t)= & {} \eta _1(t) \end{aligned}$$
(96)

where the initial conditions are \(\eta _1(0)=0\) and \(\eta _2(0)=\lambda _1\theta _1(0)-{\dot{\theta }}_1(0)\). The analytic solution of (96) is

$$\begin{aligned} y(t) = {\textbf {C}}e^{{\textbf {A}}t}\varvec{\eta }(0) + {\textbf {C}}\int _{0}^{t}\!e^{{\textbf {A}}t}{} {\textbf {B}}{\theta }_1(t-\tau )\,d\tau \end{aligned}$$
(97)

where \(\varvec{\eta }(0)=[\eta _1(0),\eta _2(0)]^T\), and

$$\begin{aligned} {\textbf {A}} = \begin{bmatrix} 0 &{}\quad 1\\ -k_1 &{}\quad -k_2 \end{bmatrix};\quad {\textbf {B}}=\begin{bmatrix} 0\\ \lambda ^2 \end{bmatrix};\quad {\textbf {C}}=\begin{bmatrix} 1&\quad 0 \end{bmatrix} \end{aligned}$$
(98)

For the analytic solution (97), the term \({\textbf {C}}e^{{\textbf {A}}t}\varvec{\eta }(0)\) represents the transient response of the auxiliary system (96). Another convolution term denotes the force response. If the initial condition of (96) is unknown, applying larger values of \(\lambda _{1}\) and \(\lambda _{2}\) could make the transient response \({\textbf {C}}e^{{\textbf {A}}t}\varvec{\eta }(0)\) decay faster, which is able to attenuate integration bias.

Appendix B Feasibility analysis of robust stability

The feasibility of the positive definite and symmetric matrices \({\textbf {P}}\) and \({\textbf {Q}}\) which satisfy (74) is discussed by introducing the linear matrix inequality (LMI) [6] as follows.

Factorizing \(\Delta {\textbf {A}}\) in (70) yields

$$\begin{aligned} \Delta {\textbf {A}}= & {} \begin{bmatrix} {\textbf {0}} &{}\quad {\textbf {0}} &{}\quad {\textbf {0}}\\ {\textbf {0}} &{}\quad {\textbf {0}} &{}\quad {\textbf {0}}\\ {\textbf {M}}_{0}^{-1} &{}\quad {\textbf {M}}_{0}^{-1} &{}\quad {\textbf {M}}_{0}^{-1} \end{bmatrix}\begin{bmatrix} \tilde{{\textbf {M}}} &{}\quad {\textbf {0}} &{}\quad {\textbf {0}} \\ {\textbf {0}} &{}\quad \tilde{{\textbf {M}}} &{}\quad {\textbf {0}}\\ {\textbf {0}} &{}\quad {\textbf {0}} &{}\quad \tilde{{\textbf {M}}} \end{bmatrix}\nonumber \\{} & {} \begin{bmatrix} {\textbf {K}}_{I} &{}\quad {\textbf {0}} &{}\quad {\textbf {0}}\\ {\textbf {0}} &{}\quad {\textbf {K}}_{P} &{}\quad {\textbf {0}}\\ {\textbf {0}} &{}\quad {\textbf {0}} &{}\quad {\textbf {K}}_{D} \end{bmatrix} \nonumber \\\triangleq & {} {\textbf {E}}\varvec{\nabla }{} {\textbf {F}} \end{aligned}$$
(99)

where

$$\begin{aligned} {\textbf {E}}= & {} \begin{bmatrix} {\textbf {0}} &{}\quad {\textbf {0}} &{}\quad {\textbf {0}}\\ {\textbf {0}} &{}\quad {\textbf {0}} &{}\quad {\textbf {0}}\\ {\textbf {M}}_{0}^{-1} &{}\quad {\textbf {M}}_{0}^{-1} &{}\quad {\textbf {M}}_{0}^{-1} \end{bmatrix};\quad \varvec{\nabla } = \begin{bmatrix} \tilde{{\textbf {M}}} &{}\quad {\textbf {0}} &{}\quad {\textbf {0}} \\ {\textbf {0}} &{}\quad \tilde{{\textbf {M}}} &{}\quad {\textbf {0}}\\ {\textbf {0}} &{}\quad {\textbf {0}} &{}\quad \tilde{{\textbf {M}}} \end{bmatrix};\nonumber \\ {\textbf {F}}= & {} \begin{bmatrix} {\textbf {K}}_{I} &{}\quad {\textbf {0}} &{}\quad {\textbf {0}}\\ {\textbf {0}} &{}\quad {\textbf {K}}_{P} &{}\quad {\textbf {0}}\\ {\textbf {0}} &{}\quad {\textbf {0}} &{}\quad {\textbf {K}}_{D} \end{bmatrix} \end{aligned}$$
(100)

and

$$\begin{aligned} \tilde{{\textbf {M}}}= & {} \left( {\textbf {I}}_{2}+\Delta {\textbf {M}}{} {\textbf {M}}_{0}^{-1}\right) ^{-1}\Delta {\textbf {M}} \end{aligned}$$
(101)

Under the assumption that the uncertain matrix \(\Delta {\textbf {M}}\) is bounded, there exists a known positive constant matrix \(\varvec{\mu }\) such that

$$\begin{aligned}{} & {} \varvec{\nabla }\varvec{\nabla }^{T} = \begin{bmatrix} \tilde{{\textbf {M}}}\tilde{{\textbf {M}}}^T &{}\quad {\textbf {0}} &{}\quad {\textbf {0}} \\ {\textbf {0}} &{}\quad \tilde{{\textbf {M}}}\tilde{{\textbf {M}}}^T &{}\quad {\textbf {0}}\\ {\textbf {0}} &{}\quad {\textbf {0}} &{}\quad \tilde{{\textbf {M}}}\tilde{{\textbf {M}}}^T \end{bmatrix} \nonumber \\ {}{} & {} < \begin{bmatrix} \varvec{\mu }^{+} &{}\quad {\textbf {0}} &{}\quad {\textbf {0}}\\ {\textbf {0}} &{}\quad \varvec{\mu }^{+} &{}\quad {\textbf {0}}\\ {\textbf {0}} &{}\quad {\textbf {0}} &{}\quad \varvec{\mu }^{+} \end{bmatrix} \triangleq \varvec{\mu } \end{aligned}$$
(102)

where \(\tilde{{\textbf {M}}}\tilde{{\textbf {M}}}^{T} < \varvec{\mu }^{+}\). Based on (99), the term \(\Delta {\textbf {A}}{^T}{} {\textbf {P}} + {\textbf {P}}\Delta {\textbf {A}}\) can be factorized by

$$\begin{aligned} \Delta {\textbf {A}}{^T}{} {\textbf {P}} + {\textbf {P}}\Delta {\textbf {A}} = {\textbf {F}}^{T}\varvec{\nabla }^{T}{} {\textbf {E}}^{T}{} {\textbf {P}} + {\textbf {P}}{} {\textbf {E}}\varvec{\nabla }{} {\textbf {F}} \end{aligned}$$
(103)

Consider the well-known inequality

$$\begin{aligned} \textbf{X}^{T}\textbf{Y} + \textbf{Y}^{T}\textbf{X} \le \varepsilon \textbf{X}^{T}\textbf{X} + \frac{1}{\varepsilon }\textbf{Y}^{T}\textbf{Y} \end{aligned}$$
(104)

where \(\varepsilon >0\), and \((\textbf{X},\textbf{Y})\) are square matrices with proper dimensions. On the basis of (102) and (104), (103) satisfies the following inequality [5]

$$\begin{aligned} {\textbf {F}}^{T}\varvec{\nabla }^{T}{} {\textbf {E}}^{T}{} {\textbf {P}} + {\textbf {P}}{} {\textbf {E}}\varvec{\nabla }{} {\textbf {F}}&\le \varepsilon {\textbf {F}}^{T}{} {\textbf {F}} + \frac{1}{\varepsilon }\left( {\textbf {P}}^{T}{} {\textbf {E}}\varvec{\nabla }\right) \left( \varvec{\nabla }^{T}{} {\textbf {E}}^{T}{} {\textbf {P}}\right) \nonumber \\&= \varepsilon {\textbf {F}}^{T}{} {\textbf {F}} + \frac{1}{\varepsilon }({\textbf {P}}{} {\textbf {E}}\varvec{\nabla }\varvec{\nabla }^{T}{} {\textbf {E}}^{T}{} {\textbf {P}})\nonumber \\&\le \varepsilon {\textbf {F}}^{T}{} {\textbf {F}} + \frac{1}{\varepsilon }{} {\textbf {P}}{} {\textbf {E}}\varvec{\mu }{} {\textbf {E}}^{T}{} {\textbf {P}} \end{aligned}$$
(105)

where \(\varepsilon >0\) is a decision variable to be solved.

Thus, the robust stability requirement (74) is achieved if the following inequality is guaranteed:

$$\begin{aligned} {\textbf {A}}^{T}{} {\textbf {P}} + {\textbf {P}}{} {\textbf {A}} + \varepsilon {\textbf {F}}^{T}{} {\textbf {F}} + \frac{1}{\varepsilon }{} {\textbf {P}}{} {\textbf {E}}\varvec{\mu }{} {\textbf {E}}^{T}{} {\textbf {P}} + {\textbf {Q}} < 0 \end{aligned}$$
(106)

By Schur complement, (106) can be reformulated as the feasibility problem of the following LMIs:

$$\begin{aligned}{} & {} \begin{bmatrix} {\textbf {A}}^{T}{} {\textbf {P}} + {\textbf {P}}{} {\textbf {A}} + \varepsilon {\textbf {F}}^{T}{} {\textbf {F}} + {\textbf {Q}} &{}\quad {\textbf {P}}{} {\textbf {E}} \\ {\textbf {E}}^{T}{} {\textbf {P}} &{}\quad -\varepsilon \varvec{\mu }^{-1} \end{bmatrix} < 0,\nonumber \\{} & {} {\textbf {P}}^{T}={\textbf {P}}>0,\quad {\textbf {Q}}^{T}={\textbf {Q}}>0, \quad \varepsilon > 0. \end{aligned}$$
(107)

As a result, given \(({\textbf {K}}_{P},{\textbf {K}}_{I},{\textbf {K}}_{D})\) and \(\varvec{\mu }\), the robust stability of the closed-loop error dynamics (64) is guaranteed if the LMIs in (107) are satisfied.

Appendix C Analysis of quantization error

The quantization error \(e_{Q}(t)\) is the error caused by using the quantized signal \(x_{Q}(t)\) instead of the true signal x(t), defined as

$$\begin{aligned} e_{Q}(t) = x_{Q}(t) - x(t) \end{aligned}$$
(108)

The quantization interval is denoted as Q. The probability density function of the quantization error is given by

$$\begin{aligned} {\mathcal {P}}(e_{Q}) = {\left\{ \begin{array}{ll} \frac{1}{Q}, &{}\quad -\frac{Q}{2} \le e_{Q} \le \frac{Q}{2}\\ 0, &{}\quad \text{ otherwise } \end{array}\right. } \end{aligned}$$
(109)

By definition, the expectation of \(e_{Q}\) is

$$\begin{aligned} {\mathcal {E}}\left[ e_{Q}\right]&= \int _{-\infty }^{\infty }\! e_{Q}{\mathcal {P}}(e_{Q})\,d e_{Q}\nonumber \\&= \int _{-Q/2}^{Q/2}\!e_{Q}\cdot \frac{1}{Q}\,de_{Q}\nonumber \\&= \frac{1}{2Q} \left. e^{2}\right| _{-Q/2}^{Q/2}\nonumber \\&= 0 \end{aligned}$$
(110)

and the variance is

$$\begin{aligned} {{\,\textrm{var}\,}}{(e_{Q})}&= {\mathcal {E}}\left[ e_{Q}^{2}\right] - {\mathcal {E}}\left[ e_{Q}\right] ^{2}\nonumber \\&= \int _{-\infty }^{\infty }\!e_{Q}^{2}{\mathcal {P}}(e_{Q})\,d e_{Q} - 0\nonumber \\&= \frac{1}{Q}\int _{-Q/2}^{Q/2}\!e_{Q}^{2}d e_{Q}\nonumber \\&= \frac{1}{3Q}\left. e_{Q}^{3}\right| _{Q/2}^{Q/2}\nonumber \\&= \frac{Q^{2}}{12} \end{aligned}$$
(111)

Hence, due to the zero-mean property, the root-mean-square error of \(e_{Q}\) can be obtained by taking the square root of \({{\,\textrm{var}\,}}{(e_{Q})}\):

$$\begin{aligned} {{\,\textrm{RMS}\,}}{\{e_{Q}(t)\}} = \sqrt{{{\,\textrm{var}\,}}{(e_{Q})}} = \frac{Q}{\sqrt{12}} \end{aligned}$$
(112)

The property (112) will be used in the cost function of the optimal trajectory estimation presented in (85).

If the quantized signal is used in the traditional velocity estimation, it will induce significant estimation errors. For example, estimate the velocity \({\dot{x}}(t)\) at \(t=kT\) by applying the backward difference formula:

$$\begin{aligned} {\dot{x}}(t)|_{t=kT}&= \frac{x_{Q}(kT)-x_{Q}(kT-T)}{T}\nonumber \\&= \frac{x(kT)-x(kT-T)}{T} + {\dot{e}}(kT) \end{aligned}$$
(113)

where T is the sample interval and

$$\begin{aligned} {\dot{e}}(kT) = \frac{e_{Q}(kT)-e_{Q}(kT-T)}{T} \end{aligned}$$
(114)

It can be seen that velocity estimation error \({\dot{e}}(kT)\) is amplified by a factor of 1/T due to the numerical differentiation. Hence, if the quantized signal is used in the acceleration estimation, the estimation error will be affected by a factor of \(1/T^{2}\).

Appendix D Derivation of low-pass filter-based filtered regression model

The studies [27, 29, 34, 35] introduce the first-order low-pass filter to establish a filtered regression model to avoid differentiating the acceleration. The derivation is discussed as follows.

Rewrite (20) for the regression model as

$$\begin{aligned} \varvec{\uptau }(t) = \left[ \frac{d}{dt}\varvec{\varphi }_1(t) + \varvec{\varphi }_0(t)\right] \textbf{X} \end{aligned}$$
(115)

where

$$\begin{aligned}{} & {} \varvec{\varphi }_1(t) = \begin{bmatrix} {\dot{\theta }}_1 &{}\quad {\dot{\theta }}_1+{\dot{\theta }}_2 &{}\quad 0 &{}\quad \phi _{1,14} &{}\quad 0 &{}\quad \theta _1 &{}\quad 0 &{}\quad 0 &{}\quad 0\\ 0 &{}\quad {\dot{\theta }}_1+{\dot{\theta }}_2 &{}\quad 0 &{}\quad \phi _{1,24} &{}\quad 0 &{}\quad 0 &{}\quad \theta _2 &{}\quad 0 &{}\quad 0 \end{bmatrix}\nonumber \\{} & {} \varvec{\varphi }_0(t)\nonumber \\{} & {} = \begin{bmatrix} 0 &{}\quad 0 &{}\quad gc_{\theta _1} &{}\quad \phi _{0,14} &{}\quad gc_{\theta _{12}} &{}\quad 0 &{}\quad 0 &{}\quad \phi _{0,18} &{}\quad 0\\ 0 &{}\quad 0 &{}\quad 0 &{}\quad \phi _{0,24} &{}\quad gc_{\theta _{12}} &{}\quad 0 &{}\quad 0 &{}\quad 0 &{}\quad \phi _{0,29} \end{bmatrix}\nonumber \\ \end{aligned}$$
(116)

and \(\phi _{1,14}\), \(\phi _{0,14}\), \(\phi _{0,24}\), \(\phi _{0,18}\), \(\phi _{0,29}\) are defined in (28). Taking the Laplace transform of (115) yields

$$\begin{aligned} \left[ s\varvec{\varphi }_1(s) - \varvec{\varphi }_1(0) + \varvec{\varphi }_0(s)\right] \textbf{X} = \varvec{\uptau }(s) \end{aligned}$$
(117)

where s is the Laplace operator and

$$\begin{aligned} \varvec{\varphi }_1(s)= & {} {\mathscr {L}}\left\{ \varvec{\varphi }_1(t)\right\} \nonumber \\ \varvec{\varphi }_0(s)= & {} {\mathscr {L}}\left\{ \varvec{\varphi }_0(t)\right\} \nonumber \\ \varvec{\uptau }(s)= & {} {\mathscr {L}}\left\{ \varvec{\uptau }(t)\right\} \end{aligned}$$
(118)

Introduce the first-order low-pass filter as

$$\begin{aligned} f(s)= \frac{\lambda }{s+\lambda } \end{aligned}$$
(119)

where \(\lambda >0\) denotes the cut-off frequency of the filter. Multiplying both sides of (117) by (119) gives

$$\begin{aligned} \varvec{\uptau }^{1f}(s) = \varvec{\varphi }^{1f}(s)\textbf{X} \end{aligned}$$
(120)

where

$$\begin{aligned} \varvec{\uptau }^{1f}(s)= & {} \frac{\lambda }{s+\lambda }\varvec{\uptau }(s)\nonumber \\ \varvec{\varphi }^{1f}(s)= & {} \varvec{\varphi }_{1}^{1f}(s) + \varvec{\varphi }_0^{1f}(s)\nonumber \\ \varvec{\varphi }_{1}^{1f}(s)= & {} \frac{\lambda s}{s+\lambda }\varvec{\varphi }_1(s) - \frac{\lambda }{s+\lambda } \varvec{\varphi }_1(0)\nonumber \\ \varvec{\varphi }_{0}^{1f}(s)= & {} \frac{\lambda }{s+\lambda }\varvec{\varphi }_0(s) \end{aligned}$$
(121)

Reformulating \(\varvec{\varphi }_1^{1f}(s)\) to eliminate \(s\varvec{\varphi }_1(s)\) produces

$$\begin{aligned} \varvec{\varphi }_{1}^{1f}(s) ={}&\frac{\lambda s}{s+\lambda }\varvec{\varphi }_1(s) - \frac{\lambda }{s+\lambda } \varvec{\varphi }_1(0)\nonumber \\ ={}&\frac{\lambda (s+\lambda )}{{s+\lambda }}\varvec{\varphi }_1(s)\nonumber \\&+ \frac{1 }{s+\lambda }\left( -\lambda ^{2}\varvec{\varphi }_1(s)-\lambda \varvec{\varphi }_1(0)\right) \nonumber \\ ={}&\lambda \varvec{\varphi }_{1}(s) + \textbf{y}_1(s) \end{aligned}$$
(122)

where

$$\begin{aligned} {\textbf {y}}_1(s) = \frac{1 }{s+\lambda }\left( -\lambda ^{2}\varvec{\varphi }_1(s)-\lambda \varvec{\varphi }_1(0)\right) \end{aligned}$$
(123)

Taking the inverse Laplace transform of (120) gives the following filtered regression model:

$$\begin{aligned} \varvec{\uptau }^{1f}(t) = \varvec{\varphi }^{1f}(t)\textbf{X} \end{aligned}$$
(124)

where

$$\begin{aligned} \varvec{\uptau }^{1f}(t)&= {\mathscr {L}}^{-1}\left\{ \varvec{\uptau }^{1f}(s)\right\} \end{aligned}$$
(125)
$$\begin{aligned} \varvec{\varphi }^{1f}(t)&= {\mathscr {L}}^{-1}\left\{ \varvec{\varphi }^{1f}(s)\right\} \nonumber \\&=\varvec{\varphi }^{1f}_1(t)+\varvec{\varphi }^{1f}_0(t)\nonumber \\&= \lambda \varvec{\varphi }_1(t)+{\textbf {y}}_1(t) +\varvec{\varphi }^{1f}_0(t) \end{aligned}$$
(126)

The integral quantities \(\varvec{\uptau }^{1f}(t)\), \({\textbf {y}}_1(t)\), and \(\varvec{\varphi }_0^{1f}(t)\) are obtained by numerically integrating the following matrix differential equations:

$$\begin{aligned} \dot{\varvec{\uptau }}^{1f}(t)= & {} -\lambda \varvec{\uptau }^{1f}(t) + \lambda \varvec{\uptau }(t),\quad \varvec{\uptau }^{1f}(0)={\textbf {0}}\nonumber \\ \dot{\textbf{y}}_1(t)= & {} -\lambda \textbf{y}_1(t) - \lambda ^2 \varvec{\varphi }_1(t),\quad \textbf{y}_1(0)=-\lambda \varvec{\varphi }_1(0)\nonumber \\ \dot{\varvec{\varphi }}^{1f}_0(t)= & {} -\lambda \varvec{\varphi }_{0}^{1f}(t) + \lambda \varvec{\varphi }_0(t),\quad \varvec{\varphi }^{1f}_0(0)={\textbf {0}} \end{aligned}$$
(127)

The parameter estimation can be conducted by using (20) shown in Sect. 3.

Rights and permissions

Springer Nature or its licensor (e.g. a society or other partner) 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

Li, YR., Peng, CC. & Juang, JN. An integral method for parameter identification of a nonlinear robot subject to quantization error. Nonlinear Dyn 111, 22419–22441 (2023). https://doi.org/10.1007/s11071-023-09027-z

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s11071-023-09027-z

Keywords

Navigation