Skip to main content
Log in

Practical realization of discrete-time Volterra series for high-order nonlinearities

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

Abstract

Full realization of all versions of Volterra series like pure, truncated, and doubly finite Volterra series, and especially, the realization of their high orders is an intractable problem. Hence, practical implementation of Volterra series for high-order nonlinearities is not feasible with reasonable computational cost. For this reason, mathematicians, neuroscientists, and especially, biomedical and electrical engineers are forced to use only the low-order Volterra series. In this paper, we provide a full realization of off-repetitive discrete-time Volterra series (ORDVS) by departure from a traditional approach in favor of choosing a hierarchical structure. The proposed method is named fast full tantamount of off-repetitive discrete-time Volterra series (FFT-ORDVS). We have proven that the proposed off-repetitive discrete-time Volterra series approximates the basic discrete-time Volterra series very well and with much less computational complexity. In a conventional method, if \({M} +1\) is considered as the memory length of the ORDVS, around \(2^{M}\) math operations are needed for the full realization of it. In most cases, M is a large number and consequently, \(2^{M}\) is too large. To solve this problem, we have proposed a simple polynomial time solution and using the proposed method, the same task is done only by 6M math operations. It means that we have found a shortcut to change an intractable problem (\({O}(2^{M}))\) to a simple P problem (O(M)). This achievement enables researchers to use high-order kernels and consequently covers high-order nonlinearities with the lowest possible computational load. We have proven our claims mathematically and validated the performance of the proposed method using two numerical examples and a real problem.

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

Notes

  1. Millions of instructions per second.

References

  1. Hopcroft, J., Ullman, J.: Introduction to Automata Theory, Languages, and Computation, 3rd edn. Addison-Wesley, Boston (2013)

    Google Scholar 

  2. Steane, A.M.: Efficient fault-tolerant quantum computing. Nature 399, 124 (1999)

    Google Scholar 

  3. Peruzzo, A., et al.: A variational eigenvalue solver on a photonic quantum processor. Nat. Commun. 5, 4213 (2014)

    Google Scholar 

  4. Kastner, M.A.: Prospects for quantum dot implementation of adiabatic quantum computers for intractable problems. Proc. IEEE 93(10), 1765–1771 (2005)

    Google Scholar 

  5. Korenberg, M.J., Hunter, I.W.: The identification of nonlinear biological systems: volterra kernel approaches. Ann. Biomed. Eng. 24(2), 250–268 (1996)

    Google Scholar 

  6. Valenza, G., Citi, L., Scilingo, E.P.: Point-process nonlinear models with Laguerre and Volterra expansions: instantaneous assessment of heartbeat dynamics. IEEE Trans. Signal Process. 61(11), 2914–2926 (2013)

    Google Scholar 

  7. Fournier, J., Monier, C., Pananceau, M., Frégnac, Y.: Adaptation of the simple or complex nature of V1 receptive fields to visual statistics. Nat. Neurosci. 14(8), 1053 (2011)

    Google Scholar 

  8. Song, D., Robinson, B.S., Granacki, J.J., Berger, T.W.: Implementing spiking neuron model and spike-timing-dependent plasticity with generalized Laguerre–Volterra models. presented at the 36th Annual international conference of the IEEE, Chicago (2014)

  9. Jing, X., Simpsonb, D.M., Allenb, R., Newlandc, P.L.: Understanding neuronal systems in movement control using Wiener/Volterra kernels: a dominant feature analysis. J. Neurosci. Methods 203, 220–232 (2012)

    Google Scholar 

  10. Song, D., Chan, R.H.M., Marmarelis, V.Z., Hampson, R.E., Deadwyler, S.A., Berger, T.W.: Nonlinear modeling of neural population dynamics for hippocampal prostheses. Neural Netw. 22, 1340–1351 (2009)

    Google Scholar 

  11. Scheibert, J., Leurent, S., Prevost, A., Debrégeas, G.: The role of fingerprints in the coding of tactile information probed with a biomimetic sensor. Science 323, 1503–1506 (2009)

    Google Scholar 

  12. Cunningham, J., Lee, J., Subramanian, T., Almekkawy, M.: Transcranial enhanced Ultrasound Imaging of induced substantia nigra in brain using adaptive third order volterra filter: in-vivo results. In: 2017 IEEE 14th International symposium on biomedical imaging (ISBI 2017), pp. 684–687 (2017)

  13. Hu, E.Y., Yu, G., Song, D., Bouteiller, C.J.-M., Berger, W.T.: Modeling nonlinear synaptic dynamics: a Laguerre–Volterra network framework for improved computational efficiency in large scale simulations. In: 2018 40th Annual international conference of the IEEE engineering in medicine and biology society (EMBC), pp. 6129–6132 (2018)

  14. Liu, Y.-J., Chen, W., Zhou, J., Zhou, B.-H., Ghannouchi, F.M., Liu, Y.-N.: Modified least squares extraction for Volterra-series digital predistorter in the presence of feedback measurement errors. IEEE Trans. Microw. TheoryTech. 60(11), 3559–3570 (2012)

    Google Scholar 

  15. Borys, A.: Consideration of Volterra series with excitation and/or impulse responses in the form of dirac impulses. IEEE Trans. Circuits Syst. I Express Briefs 57(6), 466–470 (2010)

    Google Scholar 

  16. McRory, J.G., Johnson, R.: Volterra kernel estimation for mildly nonlinear amplifiers. Electron. Lett. 29(23), 2007–2008 (1993)

    Google Scholar 

  17. Crespo-Cadenas, C., Reina-Tosina, J., Madero-Ayora, M.J., Allegue-Martínez, M.: A Volterra-based procedure for multi-port and multi-zone GaN FET amplifier CAD simulation. IEEE Trans. Circuits Syst. I Regul. Pap. 60(11), 3022–3032 (2013)

  18. Yu, H., El-Sankary, K., El-Masry, E.I.: Distortion analysis using Volterra series and linearization technique of nano-scale bulk-driven CMOS RF amplifier. IEEE Trans. Circuits Syst. I Regul. Pap. 62(1), 19–28 (2015)

    Google Scholar 

  19. Chierchie, F., Aase, S.O.: Volterra models for digital PWM and their inverses. IEEE Trans. Circuits Syst. I Regul. Pap. 62(10), 2606–2616 (2015)

    Google Scholar 

  20. Hao, R., Zhang, X., Gao, H., Cheng, J., Li, G.: A Volterra-series-based design technique for the bias circuit of class-AB RF power amplifier. In: 2018 2nd IEEE Advanced information management, communicates, electronic and automation control conference (IMCEC), pp. 54–57 (2018)

  21. Mathews, V.J.: Orthogonalization of correlated Gaussian signals for Volterra system identification. IEEE Signal Process. Lett. 2(10), 188–190 (1995)

    Google Scholar 

  22. Marmarelis, V.Z.: Nonlinear Dynamic Modeling of Physiological Systems. Wiley, New York (2004)

    Google Scholar 

  23. Westwick, DavidT, Kearney, RobertE: Identification of Nonlinear Physiological Systems. IEEE Press, Piscataway (2005)

    Google Scholar 

  24. Wiener, N.: Nonlinear Problems in Random Theory. Wiley, New York (1958)

    Google Scholar 

  25. Marmarelis, P.Z., Marmarelis, V.Z.: The white-noise method in system identification. In: Moore, G.P. (ed.) Analysis of physiological systems (Computers in biology and medicine), pp. 131–180. Springer (1978)

  26. Greblicki, W.: Nonparametric identification of Wiener systems. IEEE Trans. Inf. Theory 38(5), 1487–1493 (1992)

    Google Scholar 

  27. Greblicki, W., Pawlak, M.: Nonparametric identification of Hammerstein systems. IEEE Trans. Inf. Theory 35(2), 409–418 (1989)

    Google Scholar 

  28. Lee, Y.W., Schetzen, M.: Measurement of the Wiener kernels of a non-linear system by cross-correlation. Int. J. Control 2, 237–254 (1965)

    Google Scholar 

  29. Goussard, Y., Krenz, W.C., Stark, L.: An improvement of the Lee and Schetzen cross-correlation method. IEEE Trans. Autom. Control AC–30(9), 895–898 (1985)

    Google Scholar 

  30. Gautier, M., Monsion, M., Sagaspe, J.P.: Multidimensional Laguerre transform. IEEE Trans. Autom. Control 23(3), 488–489 (1978)

    Google Scholar 

  31. Korenberg, M.J.: Parallel cascade identification and kernel estimation for nonlinear systems. Ann. Biomed. Eng. 19, 429–45 (1991)

    Google Scholar 

  32. Kibangou, A.Y., Favier, G.: Identification of parallel-cascade wiener systems using joint diagonalization of third-order Volterra kernel slices. IEEE Signal Process. Lett. 16(3), 188–191 (2009)

    Google Scholar 

  33. Panicker, T.M., Mathews, V.J.: Parallel-cascade realizations and approximations of truncated Volterra systems. In: Acoustics, Speech, and Signal Processing, 1996. ICASSP-96, vol. 3, pp. 1589-1592 (1996)

  34. Zoubir, A.M.: Identification of quadratic Volterra systems driven by non-Gaussian processes. IEEE Trans. Signal Process. 43(5), 1302–1306 (1995)

    Google Scholar 

  35. Ralston, J.C., Zoubir, A.M., Boashash, B.: Identification of a class of nonlinear systems under stationary non-Gaussian excitation. IEEE Trans. Signal Process. 45(3), 719–735 (1997)

    Google Scholar 

  36. Marmarelis, V.Z., Zhao, X.: Volterra models and three-layer perceptrons. IEEE Trans. Neural Netw. 8(6), 1421–1433 (1997)

    Google Scholar 

  37. Kun, C., Lixiong, L.: Volterra series-based neural network and its application in tap-water flow forecast. In: 4th International symposium on computational intelligence and design, pp. 97–100 (2011)

  38. Yao, L.: Genetic algorithm based identification of nonlinear systems by sparse Volterra filters. IEEE Trans. Signal Process. 47(12), 3433–3435 (1999)

    Google Scholar 

  39. Chang, W.D.: Volterra filter modeling of nonlinear discrete-time system using improved particle swarm optimization. Digital Signal Process. 22, 1056–1062 (2012)

    Google Scholar 

  40. Korenberg, J.M.: Identification of biological cascades of linear and static nonlinear systems. In: Proceedings 16th midwest symposium circuit Theory, pp. 1–9 (1973)

  41. Korenberg, M.J., Hunter, I.W.: The identification of nonlinear biological systems: LNL cascade models. Biol. Cybern. 55(2–3), 125–134 (1986)

    Google Scholar 

  42. Ogura, H.: Estimation of Wiener kernels of a nonlinear system and a fast algorithm using digital Laguerre filters. presented at the 15th NIBB conference, Okazaki (1985)

  43. Marmarelis, V.: Identification of nonlinear biological systems using Laguerre expansions of kernels. Ann. Biomed. Eng. 21, 573–589 (1993)

    Google Scholar 

  44. Annabestani, M., Naghavi, N., Maymandi-Nejad, M.: Non-autoregressive nonlinear identification of IPMC in large deformation situations using generalized Volterra-based approach. IEEE Trans. Instrum. Meas. 65, 2866–2872 (2016)

    Google Scholar 

  45. Annabestani, M., Naghavi, N.: Nonlinear identification of IPMC actuators based on ANFIS-NARX paradigm. Sens. Actuators A Phys. 209, 140–148 (2014)

    Google Scholar 

  46. Shahinpoor, M., BarCohen, Y., Simpson, J.O., Smith, J.: Ionic polymer-metal composites (IPMC) as biomimetic sensors, actuators & artificial muscles–a review. J. Smart Mater. Struct. 7, R15–R30 (1998)

    Google Scholar 

  47. Shahinpoor, M., Kim, K.J.: Ionic polymer-metal composites: I. Fundamentals. Inst. Phys. Publ. J. Smart Mater. Struct. 10, 819–833 (2001)

    Google Scholar 

  48. Annabestani, M., Naghavi, N.: Nonuniform deformation and curvature identification of ionic polymer metal composite actuators. J. Intell. Mater. Syst. Struct. 26, 582–598 (2014)

    Google Scholar 

  49. Annabestani, M., Naghavi, N., Maymandi-Nejad, M.: From modeling to implementation of a method for restraining back relaxation in ipmc soft actuators. J. Intell. Mater. Syst. Struct. 29(15), 3124–3135 (2018)

    Google Scholar 

  50. Annabestani, M., Maymandi-Nejad, M., Naghavi, N.: Restraining IPMC back relaxation in large bending displacements: applying non-feedback local Gaussian disturbance by patterned electrodes. IEEE Trans. Electron Devices 63(4), 1689–1695 (2016)

    Google Scholar 

  51. Bonomo, C., Brunetto, P., Fortuna, F., Giannone, P., Graziani, S., Strazzeri, S.: A tactile sensor for biomedical applications based on IPMCs. IEEE Sens. J. 8(8), 1486–1493 (2008)

    Google Scholar 

  52. Brunetto, P., Fortuna, L., Giannone, P., Graziani, S., Pagano, F.: A resonant vibrating tactile probe for biomedical applications based on IPMC. IEEE Trans. Instrum. Meas. 59(5), 1453–1462 (2010)

    Google Scholar 

  53. Chang, Y.C., Kim, W.J.: Aquatic Ionic-polymer-metal-composite insectile robot with multi-DOF legs. IEEE/ASME Trans. Mechatron. 18(2), 547–555 (2013)

    Google Scholar 

  54. Chen, Z., Shatara, S., Tan, X.: Modeling of biomimetic robotic fish propelled by an ionic polymer–metal composite caudal fin. IEEE/ASME Trans. Mechatron. 15(3), 448–459 (2010)

    Google Scholar 

  55. Arena, P., Bonomo, C., Fortuna, L., Frasca, M., Graziani, S.: Design and control of an IPMC wormlike robot. IEEE Trans. Syst. Man Cybern. Part B Cybern. 36(5), 1044–1052 (2006)

    Google Scholar 

  56. Ruiz, S., Mead, B., Palmre, V., Kim, K.J., Yim, W.: A cylindrical ionic polymer-metal composite-based robotic catheter platform: modeling, design and control. Smart Mater. Struct. 24, 015007 (2015)

    Google Scholar 

  57. Annabestani, M., Fardmanesh, M.: Ionic electro active polymer-based soft actuators and their applications in microfluidic micropumps, microvalves, and micromixers: a review. arXiv preprint arXiv:1904.07149 (2019)

  58. Annabestani, M., Mohammadzadeh, H., Aghassizadeh, A., Azizmohseni, S., Fardmanesh, M.: Active microfluidic micromixer design using Ionic polymer–metal composites. In: Iranian conference on electrical engineering (ICEE) (2019)

  59. Sigrist, Z., Grivel, E., Alcoverro, B.: Estimating second-order Volterra system parameters from noisy measurements based on an LMS variant or an errors-in-variables method. Signal Process. 92(4), 1010–1020 (2012)

    Google Scholar 

  60. Mathews, V.J.: Adaptive Volterra filters using orthogonal structures. Signal Process. Lett. IEEE 3(12), 307–309 (1996)

    Google Scholar 

  61. Billings, S.A., Fakhouri, S.Y.: Non-linear system identification using the Hammerstein model. Int. J. Syst. Sci. 10(5)567–578, 1979/05/01 (1979)

Download references

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Nadia Naghavi.

Ethics declarations

Conflicts of interest

The authors declared no potential conflicts of interest with respect to the research, authorship, and/or publication of this article.

Financial support

The authors received no financial support for the research, authorship, and/or publication of this article.

Additional information

Publisher's Note

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

Appendices

Appendices

Appendix A

By mathematical simplification of the rth-order Volterra functional, it is straightforward to show that there is no noticeable difference in the existence or absence of the repetitive terms in the realization of Volterra series. Therefore, in order to estimate the rth-order Volterra functional in Eqs. (1) and (2), \(N_{{\mathrm{BDVS}}} \left( r \right) \) and \(N_{{\mathrm{ORDVS}}} \left( r \right) \) parameters should be estimated for the basic discrete-time Volterra series (Volterra series with repetitive and non-repetitive components) and for the off-repetitive discrete-time Volterra series (Volterra series with non-repetitive components), respectively. Consequently, we reach Eqs. (A.1) and (A.2) as follows.

$$\begin{aligned} N_{{\mathrm{BDVS}}} (r)= & {} \frac{\left( {M+r} \right) !}{r!M!} \end{aligned}$$
(A.1)
$$\begin{aligned} N_{{\mathrm{ORDVS}}} (r)= & {} \frac{\left( {M+1} \right) !}{r!\left( {M+1-r} \right) !} \end{aligned}$$
(A.2)

where \(M+1\) is the memory length.

We define the ratio of the number of the non-repetitive components to the number of all components (repetitive and non-repetitive) as a criterion for comparison between exploited information in the basic Volterra series (BDVS) and the ORDVS.

$$\begin{aligned} \% \psi (r)=\left( {\frac{N_{{\mathrm{ORDVS}}} (r)}{N_{{\mathrm{BDVS}}} (r)}} \right) 100 \end{aligned}$$
(A.3)

By substituting Eqs. (A.1) and (A.2) into Eq. (A.3) and after some mathematical simplification, we obtain the following equation.

$$\begin{aligned} \% \psi (r)=\left( {\frac{\Gamma \left( {M+1} \right) \Gamma \left( {\frac{M+2-r}{1+r}} \right) }{\Gamma \left( {\frac{M+r}{1+r}} \right) \Gamma \left( {M+2-r} \right) }} \right) 100 \end{aligned}$$
(A.4)

where \(\Gamma \) is the Gamma function. In an ideal state, if this ratio equals to 100, we can conclude that all exploited information in the basic discrete-time Volterra series (BDVS) have been used exactly in the off-repetitive discrete-time Volterra series (ORDVS). Since the value of M is much greater than the value of r (\(M>>r)\), it is obvious the value of \(\% \psi \left( r \right) \) tends to reach 100. For example, if the input signal discretized by a 10 MS/s A/D and using just 5 ms memory length, the results that are shown in Table 4 are obtained. Results shown in Table 4 indicate that almost all information of the BDVS (Volterra series with repetitive and non-repetitive components) have appeared in different orders of the ORDVS (Volterra series with non-repetitive components). This means that there is no noticeable difference in the existence or absence of the repetitive terms in Volterra series.

Appendix B

Equation (48) was calculated using Stirling’s approximation and Laplace’s method by following procedures.

We know

$$\begin{aligned} \max (N\,(r)) = N\,(r) \quad \forall \,r=\left( \frac{M}{2}+1\right) . \end{aligned}$$
(B.1)

Then by substituting \(r=\left( \frac{M}{2}+1\right) \) into Eq. (43) we have

$$\begin{aligned} \max (N\,(r)) = N\,\left( \frac{M}{2}+1\right) = \frac{(M+1)!}{\left( \frac{M}{2}\right) !\left( \frac{M}{2}\right) !}. \end{aligned}$$
(B.2)

An appropriate alternative for \(\theta !\) is Gamma function.

$$\begin{aligned} \theta != & {} \Gamma (\theta +1) \end{aligned}$$
(B.3)
$$\begin{aligned} \theta != & {} \int \nolimits ^{\infty }_{0} x^{\theta }\hbox {e}^{-x} \hbox {d}x \Rightarrow \theta ! = \int \nolimits ^{\infty }_{0} \hbox {e}^{\theta \ln x-x} \hbox {d}x \end{aligned}$$
(B.4)

By putting \(x-\theta y\), we have:

$$\begin{aligned} \theta != & {} \hbox {e}^{\theta \ln \theta } \theta \int \nolimits ^{\infty }_{0} \hbox {e}^{\theta (\ln y-y)} \hbox {d}y \,\underrightarrow{\psi (y)=\ln \,y-y}\,\, \theta !\nonumber \\= & {} \hbox {e}^{\theta \ln \,\theta } \theta \int \nolimits ^{\infty }_{0} \hbox {e}^{\theta \psi (y)} \hbox {d}y. \end{aligned}$$
(B.5)

The function \(\psi (y)\) has a unique global maximum at \(y_{0}\) and by applying Laplace’s method, we can solve the integral term in (B.5) as follows:

$$\begin{aligned} \int \nolimits ^{\infty }_{0} \hbox {e}^{\theta \psi (y)} dy\approx \sqrt{\frac{2\pi }{\theta |\psi ''(y_{0})|}}\hbox {e}^{\theta \psi (y_{0})}. \end{aligned}$$
(B.6)

It is obvious that:

$$\begin{aligned} \psi (y)= & {} \ln \,y-y\Rightarrow \psi '(y) =\frac{1}{y}-1\Rightarrow y_{0}=1\nonumber \\ \psi ''(y)= & {} -\frac{1}{y^{2}}\Rightarrow |\psi ''(y_{0})|=1. \end{aligned}$$
(B.7)

And then

$$\begin{aligned} \int \nolimits ^{\infty }_{0} \hbox {e}^{\theta \psi (y)}dy \approx \sqrt{\frac{2\pi }{\theta }}\hbox {e}^{\theta }. \end{aligned}$$
(B.8)

By substituting (B.8) into (B.5) and after simplification, we get (B.9).

$$\begin{aligned} \theta ! \approx \hbox {e}^{\theta (\ln \,\theta -1)}\sqrt{2\pi \theta } \end{aligned}$$
(B.9)

Now called the Stirling’s series

$$\begin{aligned} \theta !\approx & {} \sqrt{2\pi \theta }\left( \frac{\theta }{e}\right) ^{\theta } \underbrace{\left( 1+\frac{1}{12\theta }+\frac{1}{288\theta ^{2}}+\cdots \right) }_{\approx 1} \rightarrow \theta !\nonumber \\\approx & {} \sqrt{2\pi \theta }\left( \frac{\theta }{e}\right) ^{\theta }. \end{aligned}$$
(B.10)

By substituting (B.10) into (B.2), we obtain

$$\begin{aligned}&\max (N(r))\nonumber \\&\quad = \frac{\sqrt{2\pi (M+1)}\left( \frac{M+1}{e}\right) ^{M+1}}{\sqrt{2\pi \left( \frac{M}{2}\right) }\left( \frac{\left( \frac{M}{2}\right) }{e}\right) ^{\frac{M}{2}}\,\,\sqrt{2\pi \left( \frac{M}{2}\right) }\left( \frac{\left( \frac{M}{2}\right) }{e}\right) ^{\frac{M}{2}}}.\nonumber \\ \end{aligned}$$
(B.11)

With simplification and a little approximation, we have

$$\begin{aligned} \max (N(r)) \approx \frac{1}{\sqrt{\pi M}}2^{M+\frac{1}{2}}. \end{aligned}$$
(B.12)

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Annabestani, M., Naghavi, N. Practical realization of discrete-time Volterra series for high-order nonlinearities. Nonlinear Dyn 98, 2309–2325 (2019). https://doi.org/10.1007/s11071-019-05330-w

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s11071-019-05330-w

Keywords

Navigation