Skip to main content
Log in

State-space realization of a describing function

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

Abstract

The describing function is a powerful tool for characterizing nonlinear dynamical systems in the frequency domain. In some cases, it is the only available description of a nonlinear operator characterizing a certain subcomponent of the system. This paper presents a methodology to provide a state-space realization of one given describing function, in order to allow the study of the system in the time domain as well. The realization is based on Hammerstein models and Fourier–Bessel series. It can be embedded in time domain simulations of complex configurations with many nonlinear elements interacting, accurately describing the nonlinear saturation of the system. The technique is applied to an example application in the field of combustion instability, featuring self-excited thermoacoustic oscillations. We benchmark the performance of the tool comparing the results with a frequency domain analysis of the same system, obtaining good agreement between the two formulations.

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.

Institutional subscriptions

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

Similar content being viewed by others

Notes

  1. Many annular combustors are also slightly non-axisymmetric, perturbing this double-Hopf bifurcation.

  2. To be precise the response is defined at discrete values of frequency and amplitude and then interpolated in between.

  3. A phasor is a representation of a sinusoidal function with a certain amplitude, frequency and phase in the complex plane.

  4. Because they are approximately in quadrature.

References

  1. Lieuwen, T., Yang, V.: Combustion Instabilities in Gas Turbine Engines. American Institute of Aeronautics and Astronautics, Reston (2005). doi:10.2514/4.866807

    Google Scholar 

  2. Lieuwen, T.: Unsteady Combustor Physics. Cambridge University Press, Cambridge (2012). doi:10.1017/CBO9781139059961

    Book  MATH  Google Scholar 

  3. Dowling, A.P.: Nonlinear self-excited oscillations of a ducted flame. J. Fluid Mech. 346, 271–290 (1997). doi:10.1017/S0022112097006484

    Article  MATH  MathSciNet  Google Scholar 

  4. Noiray, N., Durox, D., Schuller, T., Candel, S.: A unified framework for nonlinear combustion instability analysis based on the flame describing function. J. Fluid Mech. 615(2008), 139–167 (2008). doi:10.1017/S0022112008003613

    Article  MATH  Google Scholar 

  5. Palies, P., Durox, D., Schuller, T., Candel, S.: Nonlinear combustion instability analysis based on the flame describing function applied to turbulent premixed swirling flames. Combust. Flame 158(10), 1980–1991 (2011). doi:10.1016/j.combustflame.2011.02.012

    Article  Google Scholar 

  6. Schimek, S., Moeck, J.P., Paschereit, C.O.: An experimental investigation of the nonlinear response of an atmospheric swirl-stabilized premixed flame. J. Eng. Gas Turbines Power 133(10), 101502 (2011). doi:10.1115/1.4002946. http://gasturbinespower.asmedigitalcollection.asme.org/article.aspx?articleid=1429530

  7. Ćosić, B., Moeck, J.P., Paschereit, C.O.: Nonlinear instability analysis for partially premixed swirl flames. Combust. Sci. Technol. 186(6), 713–736 (2014). doi:10.1080/00102202.2013.876420

    Article  Google Scholar 

  8. Armitage, C., Balachandran, R., Mastorakos, E., Cant, R.: Investigation of the nonlinear response of turbulent premixed flames to imposed inlet velocity oscillations. Combust. Flame 146(3), 419–436 (2006). doi:10.1016/j.combustflame.2006.06.002

    Article  Google Scholar 

  9. Hemchandra, S.: Premixed flame response to equivalence ratio fluctuations: comparison between reduced order modeling and detailed computations. Combust. Flame 159(12), 3530–3543 (2012). doi:10.1016/j.combustflame.2012.08.003

    Article  Google Scholar 

  10. Krediet, H.J., Beck, C.H., Krebs, W., Schimek, S., Paschereit, C.O., Kok, J.B.W.: Identification of the flame describing function of a premixed swirl flame from LES. Combust. Sci. Technol. 184(7–8), 888–900 (2012). doi:10.1080/00102202.2012.663981

    Article  Google Scholar 

  11. Tay-Wo-Chong, L., Bomberg, S., Ulhaq, A., Komarek, T., Polifke, W.: Comparative validation study on identification of premixed flame transfer function. J. Eng. Gas Turbines Power 134(2), 021502 (2012). doi:10.1115/1.4004183. http://gasturbinespower.asmedigitalcollection.asme.org/article.aspx?articleid=1429715

  12. Zinn, B.: A theoretical study of non-linear damping by Helmholtz resonators. J. Sound Vib. 13(3), 347–356 (1970). doi:10.1016/S0022-460X(70)80023-2

    Article  MATH  Google Scholar 

  13. Bellucci, V., Flohr, P., Paschereit, C.O., Magni, F.: On the use of Helmholtz resonators for damping acoustic pulsations in industrial gas turbines. J. Eng. Gas Turbines Power 126(2), 271–275 (2004). doi:10.1115/1.1473152. http://gasturbinespower.asmedigitalcollection.asme.org/article.aspx?articleid=1421769

  14. Ćosić, B., Reichel, T.G., Paschereit, C.O.: Acoustic response of a Helmholtz resonator exposed to hot-gas penetration and high amplitude oscillations. J. Eng. Gas Turbines Power 134(10) 101503 (2012). doi:10.1115/1.4007024. http://gasturbinespower.asmedigitalcollection.asme.org/article.aspx?articleid=1477072

  15. Gelb, A., Vander Velde, W.: Multiple Input Describing Functions and Nonlinear System Design. McGraw-Hill Book, New York (1968)

    MATH  Google Scholar 

  16. Basso, M., Genesio, R., Tesi, A.: A frequency method for predicting limit cycle bifurcations. Nonlinear Dyn. 13, 339–360 (1997). doi:10.1023/A:1008298205786

    Article  MATH  MathSciNet  Google Scholar 

  17. Moeck, J.P., Paschereit, C.O.: Nonlinear interactions of multiple linearly unstable thermoacoustic modes. Int. J. Spray Combust. Dyn. 4(1), 1–28 (2012). doi:10.1260/1756-8277.4.1.1

    Article  Google Scholar 

  18. Boudy, F., Durox, D., Schuller, T., Candel, S.: Analysis of limit cycles sustained by two modes in the flame describing function framework. C. R. Mécanique 341(1), 181–190 (2013). doi:10.1016/j.crme.2012.10.014

    Article  Google Scholar 

  19. Kabiraj, L., Sujith, R.I., Wahi, P.: Bifurcations of self-excited ducted laminar premixed flames. J. Eng. Gas Turbines Power 134(3), 031502 (2012). doi:10.1115/1.4004402. http://gasturbinespower.asmedigitalcollection.asme.org/article.aspx?articleid=1429767

  20. Kashinath, K., Waugh, I.C., Juniper, M.P.: Nonlinear self-excited thermoacoustic oscillations of a ducted premixed flame: bifurcations and routes to chaos. J. Fluid Mech. 761, 399–430 (2014)

    Article  MathSciNet  Google Scholar 

  21. Lanza, V., Bonnin, M., Gilli, M.: On the application of the describing function technique to the bifurcation analysis of nonlinear systems. IEEE Trans. Circuits Syst. II Express Briefs 54(4), 343–347 (2007). doi:10.1109/TCSII.2006.890406. http://ieeexplore.ieee.org/lpdocs/epic03/wrapper.htm?arnumber=4155062

  22. Schuermans, B.: Non-linear combustion instabilities in annular gas-turbine combustors. 44th AIAA Aerospace Sciences Meeting and Exhibit. American Institute of Aeronautics and Astronautics, Virginia (2006). doi:10.2514/6.2006-549

  23. Sanders, J., Verhulst, F.: Averaging Methods in Nonlinear Dynamical Systems, Applied Mathematical Sciences, vol. 59, 2nd edn. Springer, New York (2007). doi:10.1007/978-0-387-48918-6

    Google Scholar 

  24. Kevorkian, J., Cole, J.: Multiple Scale and Singular Perturbation Methods. Springer, New York (1996). doi:10.1007/978-1-4612-3968-0

    Book  MATH  Google Scholar 

  25. Noiray, N., Bothien, M.R., Schuermans, B.: Investigation of azimuthal staging concepts in annular gas turbines. Combust. Theory Model. 15(5), 585–606 (2011). doi:10.1080/13647830.2011.552636

    Article  MATH  Google Scholar 

  26. Noiray, N., Schuermans, B.: On the dynamic nature of azimuthal thermoacoustic modes in annular gas turbine combustion chambers. Proc. R. Soc. A Math. Phys. Eng. Sci. 469 (2013). doi:10.1098/rspa.2012.0535. http://rspa.royalsocietypublishing.org/content/469/2151/20120535

  27. Ghirardo, G., Juniper, M.: Azimuthal instabilities in annular combustors: standing and spinning modes. Proc. R. Soc. A Math. Phys. Eng. Sci. 469 (2013). doi:10.1098/rspa.2013.0232. http://rspa.royalsocietypublishing.org/content/469/2157/20130232.short

  28. Taylor, J.H.: A systematic nonlinear controller design approach based on quasilinear system models. In: American Control Conference (1983)

  29. Nassirharand, A.: Matlab software for inversion of describing functions. Adv. Eng. Softw. 40(8), 600–606 (2009). doi:10.1016/j.advengsoft.2008.11.003

    Article  MATH  Google Scholar 

  30. Eskinat, E., Johnson, S.H., Luyben, W.L.: Use of Hammerstein models in identification of nonlinear systems. AIChE J. 37(2), 255–268 (1991). doi:10.1002/aic.690370211

    Article  Google Scholar 

  31. Ćosić, B., Moeck, J.P., Paschereit, C.O.: Prediction of pressure amplitudes of self-excited thermoacoustic instabilities for a partially premixed swirl-flame. In: Proceedings of ASME Turbo Expo 2013, pp. 1–11 (2013). doi:10.1115/GT2013-94160

  32. Dierckx, P.: Curve and Surface Fitting with Splines. Oxford University Press, Oxford (1993)

    MATH  Google Scholar 

  33. Gómez, J.C., Baeyens, E.: Identification of block-oriented nonlinear systems using orthonormal bases. J. Process Control 14(6), 685–697 (2004). doi:10.1016/j.jprocont.2003.09.010

    Article  Google Scholar 

  34. Schuller, T., Durox, D., Candel, S.: A unified model for the prediction of laminar flame transfer functions: comparisons between conical and V-flame dynamics. Combust. Flame 134(1–2), 21–34 (2003). doi:10.1016/S0010-2180(03)00042-7

    Article  Google Scholar 

  35. Gustavsen, B., Semlyen, A.: Rational approximation of frequency domain responses by vector fitting. IEEE Trans. Power Deliv. 14(3), 1052–1061 (1999)

  36. Gustavsen, B.: Improving the pole relocating properties of vector fitting. IEEE Trans. Power Deliv. 21(3), 1587–1592 (2006). doi:10.1109/TPWRD.2005.860281

    Article  Google Scholar 

  37. Kurokawa, S., Yamanaka, F.N.R., Prado, A.J., Pissolato, J.: Inclusion of the frequency effect in the lumped parameters transmission line model: state space formulation. Electr. Power Syst. Res. 79, 1155–1163 (2009). doi:10.1016/j.epsr.2009.02.007

  38. Bothien, M.R., Moeck, J.P., Lacarelle, A., Paschereit, C.O.: Time domain modelling and stability analysis of complex thermoacoustic systems. Proc. Inst. Mech. Eng. Part A J. Power Energy 221(5), 657–668 (2007). doi:10.1243/09576509JPE384

  39. Gustavsen, B.: Fast passivity enforcement for pole-residue models by perturbation of residue matrix eigenvalues. IEEE Trans. Power Deliv. 23(4), 2278–2285 (2008). doi:10.1109/TPWRD.2008.919027

    Article  MathSciNet  Google Scholar 

  40. Gustavsen, B., Heitz, C.: Modal vector fitting: a tool for generating rational models of high accuracy with arbitrary terminal conditions. IEEE Trans. Adv. Packag. 31(4), 664–672 (2008)

    Article  Google Scholar 

  41. Klink, W., Payne, G.: Approximating with nonorthogonal basis functions. J. Comput. Phys. 21, 208–226 (1976)

    Article  MATH  MathSciNet  Google Scholar 

  42. Lieuwen, T.: Modeling premixed combustion-acoustic wave interactions: A review. J. Propuls. Power 19(5), 765–781 (2003). doi:10.2514/2.6193

    Article  Google Scholar 

  43. Schuermans, B., Polifke, W., Paschereit, C.O.: Modeling transfer matrices of premixed flames and comparison with experimental results. ASME, International Gas Turbine and Aeroengine Congress and Exhibition. ASME, Indianapolis (1999)

  44. Choi, A.: Real-time fundamental frequency estimation by least-square fitting. IEEE Transactions on Speech and Audio Processing 5(2), 201–205 (1997). doi:10.1109/89.554783. http://ieeexplore.ieee.org/lpdocs/epic03/wrapper.htm?arnumber=554783

  45. Pankiewitz, C., Sattelmayer, T.: Time domain simulation of combustion instabilities in annular combustors. J. Eng. Gas Turbines Power 125(3), 677–685 (2003). doi:10.1115/1.1582496. http://gasturbinespower.asmedigitalcollection.asme.org/article.aspx?articleid=1421647

  46. Schuermans, B., Bellucci, V., Paschereit, C.O.: Thermoacoustic modeling and control of multi burner combustion systems. Proceedings of ASME Turbo Expo 2003 (2003). doi:10.1115/GT2003-38688

  47. Stow, S.R., Dowling, A.P.: A time-domain network model for nonlinear thermoacoustic oscillations. J. Eng. Gas Turbines Power 131(3), 031502 (2009). doi:10.1115/1.2981178. http://link.aip.org/link/JETPEZ/v131/i3/p031502/s1&Agg=doi

  48. Hajimolahoseini, H., Taban, M.R., Soltanian-Zadeh, H.: Extended Kalman filter frequency tracker for nonstationary harmonic signals. Measurement 45(1), 126–132 (2012). doi:10.1016/j.measurement.2011.09.008

    Article  Google Scholar 

Download references

Acknowledgments

This work was supported by the European Research Council through project ALORS N.259620 and by the German Research Association for Combustion Engines (Forschungsvereinigung Verbrennungskraftmaschinen e. V. FVV).

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Giulio Ghirardo.

Appendix: Describing function calculation

Appendix: Describing function calculation

In this appendix, we evaluate the describing function (defined by (1)) of the saturation function \(\mathcal {N}\). We will prove here the more general result for an input with structure

$$\begin{aligned} u(t)=A_1\cos (\omega t +\varphi _1) + A_2\cos (\omega t+\varphi _2) \end{aligned}$$
(34)

to the function \(\mathcal {N}\), as opposed to the case under consideration in this article introduced in (1) where \(u(t)=A\cos (\omega t)\). In particular, the input described by (34) describes the contribution of two modes, oscillating at the same frequency, instead of a single sinusoidal input \(A\cos (\omega t+\varphi )\). The motivation to cover this more general case is to make this framework usable in rotationally symmetric annular combustors featuring azimuthal modes. In that case, each burner is subject to the combined input of two thermoacoustic modes, depending on the amplitudes \(A_1\) and \(A_2\) of the two modes at that location, and on their phases \(\varphi _1,\varphi _2\). Once the result for the input (34) is obtained, it will be sufficient to set \(A_1=A\), \(\varphi _1=\varphi _2=A_2=0\) to obtain the special case of the single-input response used in this article, as presented at the end of the appendix.

We proceed by rewriting \(u\) as

$$\begin{aligned}&u=a\cos (\omega t)+b\sin (\omega t) \end{aligned}$$
(35)

by introducing the constants

$$\begin{aligned}&{\left\{ \begin{array}{ll} &{}a\equiv A_1\cos \varphi _1+A_2\cos \varphi _2\\ &{}b\equiv -A_1\sin \varphi _1-A_2\sin \varphi _2 \end{array}\right. } \end{aligned}$$
(36)

Notice that \(a,b\) do not depend on the time variable \(t\). We study the averaging integral in the definition (1) of describing function for the operator \(\mathcal {N}\), and we will later divide by the amplitude \(A\) to recover the full expression. In other words, for the time being we study the product \(N(A,\omega )A\). We substitute in the product Eq. (35) and change the time variable:

$$\begin{aligned} \frac{1}{\pi }\int _0^{2\pi }\mathcal {N}\left( a\cos t \!+\! b\sin t\right) \left( \cos t \!+\!i\sin t\right) {\hbox {d}}t\!=\!f_c\!+\!if_s \end{aligned}$$
(37)

We substitute the expression for \(\mathcal {N}\) from (17):

$$\begin{aligned} f_{c}&=\frac{1}{\pi }\int _0^{2\pi }q^{\text {erf}}_{\mu ,\kappa }(a\cos t+b\sin t)\cos t {\hbox {d}}t\nonumber \\&+\sum _{n=1}^N \frac{c_n}{\pi }\int _0^{2\pi }J_1(\hat{u}_n (a\cos t+b\sin t))\cos t {\hbox {d}}t\nonumber \\&\equiv f^{\text {erf}}_{c}+\sum _{n=1}^Nc_n f^{b,n}_{c} \end{aligned}$$
(38a)
$$\begin{aligned} f_{s}&=\frac{1}{\pi }\int _0^{2\pi }q^{\text {erf}}_{\mu ,\kappa }(a\cos t+b\sin t)\sin t {\hbox {d}}t\nonumber \\&+\sum _{n=1}^N \frac{c_n}{\pi }\int _0^{2\pi }J_1(\hat{u}_n (a\cos t+b\sin t))\sin t {\hbox {d}}t\nonumber \\&\equiv f^{\text {erf}}_{s}+\sum _{n=1}^Nc_n f^{b,n}_{s} \end{aligned}$$
(38b)

We study first the integrals \(f^{\text {erf}}_{c}\) and \(f^{\text {erf}}_{s}\) due to the error function in Sect. 4.1 and then each of the \(N\) integrals \(f^{b,n}_{c}\) and \(f^{b,n}_{s}\) of the Fourier–Bessel series in Sect. 4.2. We put together the expressions and discuss them in Sect. 4.3.

1.1 Averaging the error function

We substitute the definition of \(q_{\mu ,\kappa }^{\text {erf}}\) from (18) into the expression of \(f_{c}^{\text {erf}}\) and \(f_{s}^{\text {erf}}\) in (38). We obtain

$$\begin{aligned} f_{c}^{\text {erf}}&=\frac{\kappa }{\pi }\int _0^{2\pi }\text {erf}\left( \frac{\sqrt{\pi } \mu }{2\kappa }(a\cos t+b\sin t)\right) \cos t d t \end{aligned}$$
(39a)
$$\begin{aligned} f_{s}^{\text {erf}}&=\frac{\kappa }{\pi }\int _0^{2\pi }\text {erf}\left( \frac{\sqrt{\pi } \mu }{2\kappa }(a\cos t+b\sin t)\right) \sin t d t \end{aligned}$$
(39b)

For conciseness, we introduce the constant

$$\begin{aligned} k=\sqrt{\frac{\pi }{8}}\frac{\mu }{\kappa }, \end{aligned}$$
(40)

so that the argument of the \(\text {erf}\) function is \(\sqrt{2}k(a\cos t+b\sin t)\). This leads to neater expressions in the following. We proceed by expressing the argument of the exponential function as

$$\begin{aligned} a\cos t+b\sin t=R\sin (t+\psi ), \end{aligned}$$
(41)

where \(R\) and \(\psi \) are defined as

$$\begin{aligned} R\equiv&\sqrt{a^2+b^2} \end{aligned}$$
(42a)
$$\begin{aligned} \psi \equiv&\arg (b+ia) \end{aligned}$$
(42b)

The two integrals (39) become

$$\begin{aligned} f_{c}^{\text {erf}}&=\frac{\kappa }{\pi }\int _0^{2\pi }\text {erf}\left( \sqrt{2}kR\sin ( t+\psi )\right) \cos t d t \end{aligned}$$
(43a)
$$\begin{aligned} f_{s}^{\text {erf}}&=\frac{\kappa }{\pi }\int _0^{2\pi }\text {erf}\left( \sqrt{2}kR\sin ( t+\psi )\right) \sin t d t \end{aligned}$$
(43b)

We exploit the fact that the \(\text {erf}\) function is defined as an integral itself and apply integration by parts to (43a):

$$\begin{aligned} f_{c}&^{\text {erf}}=\frac{\kappa }{\pi }\left[ \text {erf}\left( \sqrt{2} kR\sin ( t+\psi )\right) \sin t\right] _0^{2\pi }\nonumber \\&\qquad -\frac{\kappa }{\pi }\frac{2}{\sqrt{\pi }}\sqrt{2}k\int _0^{2\pi }e^{-2k^2R^2\sin ^2( t+\psi )}\nonumber \\ {}&\qquad \times R\cos ( t+\psi )\sin t d t \end{aligned}$$
(44)

The first term trivially vanishes. In the second, \(2/\sqrt{\pi }\) is the factor present in the definition (18) of the \(\text {erf}\) function, and \(\sqrt{2}k\) comes from the chain rule of the derivative of \(\text {erf}\) with respect to \( t\), together with the term \(R\cos ( t+\psi )\) inside the integral. Substituting the definition of \(k\) from (40) only at its first occurrence into the second term, we can simplify:

$$\begin{aligned} \frac{\kappa }{\pi }\frac{2}{\sqrt{\pi }}\sqrt{2}k=\frac{\mu }{\pi } \end{aligned}$$
(45)

We now apply a change of integration variable \( t\rightarrow \chi -\psi \) to the integral (44), and because the integrand is periodic in \( t\) and then in \(\chi \), we keep the same limits of integration in the new variable.

$$\begin{aligned} f_{c}^{\text {erf}}=-\frac{\mu }{\pi }R\int _0^{2\pi }e^{-2k^2R^2\sin ^2\chi }\cos \chi \sin (\chi -\psi ) {\text{ d }}\chi \end{aligned}$$
(46)

We then expand the trigonometric term \(\sin (\chi -\psi )\) and take the sum out of the integral. We obtain

$$\begin{aligned} f_{c}^{\text {erf}}=&-\mu R\left[ +E_{sc}\cos \psi -E_{cc}\sin \psi \right] \end{aligned}$$
(47a)
$$\begin{aligned} f_{s}^{\text {erf}}=&-\mu R\left[ -E_{sc}\sin \psi -E_{cc}\cos \psi \right] \end{aligned}$$
(47b)

where we introduced the integrals

$$\begin{aligned} E_{sc}\equiv & {} \frac{1}{\pi }\int _0^{2\pi }e^{-2 k^2 R^2 \sin ^2\chi }\cos \chi \sin \chi d\chi =0 \end{aligned}$$
(48a)
$$\begin{aligned} E_{cc}\equiv & {} \frac{1}{\pi }\int _0^{2\pi }e^{-2 k^2 R^2 \sin ^2\chi }\cos ^2\chi {\text{ d }}\chi \nonumber \\= & {} \frac{2}{\pi }\int _0^{\pi }e^{-2 k^2 R^2 \sin ^2\chi }\cos ^2\chi {\text{ d }}\chi \end{aligned}$$
(48b)

The first integral (48a) is zero since its integrand is odd. In (48b), we exploited the fact that the integrand has period \(\pi \). We then use power reduction formulas on the terms \(\sin ^2\chi \) and \(\cos ^2\chi \):

$$\begin{aligned} E_{cc}= & {} \frac{2}{\pi }\int _0^{\pi }e^{-k^2 R^2 (1-\cos 2\chi ) }\frac{1+\cos 2\chi }{2} {\text{ d }}\chi \nonumber \\= & {} \frac{1}{2\pi }\int _0^{2\pi }e^{-k^2 R^2 (1-\cos t) }(1+\cos t){\hbox {d}}t \end{aligned}$$
(49)

The integrand in (49) has period \(2\pi \) and is an even function of \(t\), so:

$$\begin{aligned} E_{cc}= & {} \frac{1}{\pi }\int _0^{\pi }e^{-k^2 R^2+k^2 R^2 \cos t }(1+\cos t) {\hbox {d}}t\nonumber \\= & {} e^{-k^2 R^2}\left( \frac{1}{\pi }\int _0^{\pi }e^{k^2 R^2 \cos t } {\hbox {d}}t\right. \nonumber \\&\left. +\frac{1}{\pi }\int _0^{\pi }e^{k^2 R^2 \cos t }\cos t {\hbox {d}}t\right) \nonumber \\= & {} e^{-k^2 R^2} \left( I_0(k^2 R^2)+I_1(k^2 R^2)\right) \end{aligned}$$
(50)

In (50), \(I_0(x)\) and \(I_1(x)\) are the modified Bessel functions of the first kind of the zero order and first order, respectively. We first substitute (50,48a) in (47), and then we substitute \(R\sin \psi =a\) and \(R\cos \psi =b\). We obtain:

$$\begin{aligned} f_{c}^{\text {erf}}= & {} af_{nl}^{\text {erf}}(kR) \end{aligned}$$
(51a)
$$\begin{aligned} f_{2}^{\text {erf}}= & {} bf_{nl}^{\text {erf}}(kR) \end{aligned}$$
(51b)

with

$$\begin{aligned} f_{nl}^{\text {erf}}(kR)\equiv \mu e^{-k^2R^2} \left( I_0(k^2R^2)+I_1(k^2R^2)\right) \end{aligned}$$
(52)

acting as a gain, as it multiplies the linear term in (51) and depends on the amplitude of oscillation. The two analytical expressions (51) have been compared with the numerical integration of (38) and its counterpart for a few values of \(\mu ,\kappa ,a,b\), and lead to relative errors of the order of machine precision, thus confirming their validity.

1.2 Averaging Bessel functions

The \(n\)-th term of \(f_{c}^{b,n}\) and of \(f_{s}^{b,n}\) in (38) are, respectively,

$$\begin{aligned} f_{c}^{b,n}\equiv & {} \frac{1}{\pi }\int _0^{2\pi }J_1(2u_n a\cos t+2u_nb\sin t)\cos t d t\nonumber \\ \end{aligned}$$
(53a)
$$\begin{aligned} f_{s}^{b,n}\equiv & {} \frac{1}{\pi }\int _0^{2\pi }J_1(2u_n a\cos t+2u_nb\sin t)\sin t d t\nonumber \\ \end{aligned}$$
(53b)

where we introduced \(u_n\equiv \hat{u}_n/2\). We define \(\hat{f}_j^{b,n}=f_{c}^{b,n}+if_{s}^{b,n}\) and apply the substitution

$$\begin{aligned} z=e^{i t},\,\sin t=\frac{1}{2i}\left( 1-\frac{1}{z}\right) ,\,\cos t=\frac{1}{2}\left( 1+\frac{1}{z}\right) \end{aligned}$$
(54)

We obtain

$$\begin{aligned} \hat{f}_j^{b,n}=\frac{1}{\pi }\int _0^{2\pi }J_1\left( u_n (a-ib)z+u_n(a+ib)\frac{1}{z}\right) z d t \end{aligned}$$
(55)

We now change the line integral into a contour integral in the complex plane on the circle \(|z|=1\). From (54), we have that \(d t=dz/iz\), and

$$\begin{aligned} \hat{f}_j^{b,n}=\frac{1}{\pi i}\oint _{|z|=1} J_1\left( u_n (a-ib)z+u_n(a+ib)\frac{1}{z}\right) dz \end{aligned}$$
(56)

The Bessel function \(J_1(z)\) is an entire function, so the only singularity of \(\hat{f}_j^{b,n}(z)\) is at the origin and is of the essential type. We can then apply the residue theorem,

$$\begin{aligned} \hat{f}_j^{b,n}&=\frac{1}{\pi i}\,2\pi i\text{ Res }\left[ J_1\left( u_n (a-ib)z+u_n(a+ib)\frac{1}{z}\right) \right] _{z=0}\nonumber \\&=2\text{ Res }\left[ G(z)\right] _{z=0} \end{aligned}$$
(57)

We expand the Bessel function in \(G(z)\) with its Laurent series:

$$\begin{aligned} G(z)=\sum _{m=0}\frac{(-1)^mu_n^{1+2m}}{m!(m+1)!}\left( \frac{(a-ib)z+(a+ib)/z}{2}\right) ^{1+2m} \end{aligned}$$
(58)

We substitute the binomial expansion of the power of the sum

$$\begin{aligned}&\left( \frac{(a-ib)z+(a+ib)/z}{2}\right) ^{1+2m}\nonumber \\&\quad =\frac{1}{2^{1+2m}}\sum _{k=0}^{1+2m}z^{k-(2m+1-k)}\left( {\begin{array}{c}1+2m\\ k\end{array}}\right) \nonumber \\ {}&\qquad \times (a-ib)^k(a+ib)^{1+2m-k} \end{aligned}$$
(59)

The residue in (58) is the sum of the coefficients of the term \(1/z\). Therefore, in the sum (59), we retain only the term with \(k-(2m+1-k)=-1\), from which follows \(k=m\). This term of (59) is:

$$\begin{aligned}&\frac{1}{2^{1+2m}}z^{-1}\left( {\begin{array}{c}1+2m\\ m\end{array}}\right) (a-ib)^{m}(a+ib)^{m+1}\nonumber \\&=\frac{a+ib}{2}\frac{1}{2^{2m}}z^{-1}\frac{(2m+1)!}{m!(m+1)!}(a^2+b^2)^{m} \end{aligned}$$
(60)

Equation (57) evaluates to

$$\begin{aligned} \hat{f}_j^{b,n}=&(a+ib)\sum _{m=0}^\infty \frac{(-1)^m(2m+1)!u_n^{1+2m}}{(m!(m+1)!)^2}\left( \frac{R}{2}\right) ^{2m} \end{aligned}$$

with \(R=\sqrt{a^2+b^2}\). This series converges to

$$\begin{aligned} \hat{f}_j^{b,n}=&(a+i b)\frac{2 J_0\left( u_n\sqrt{a^2+b^2}\right) J_1\left( u_n\sqrt{a^2+b^2}\right) }{\sqrt{a^2+b^2}} \end{aligned}$$
(61)

The two forcing terms (53) can be evaluated as the real and imaginary parts of (61):

$$\begin{aligned} f_{c}^{b,n}= & {} a f_{nl}^{b,n}(R) \end{aligned}$$
(62a)
$$\begin{aligned} f_{s}^{b,n}= & {} b f_{nl}^{b,n}(R) \end{aligned}$$
(62b)

where we introduced

$$\begin{aligned} f_{nl}^{b,n}(R)\equiv 2\frac{J_0\left( u_nR\right) J_1\left( u_nR\right) }{R} \end{aligned}$$
(63)

1.3 Final expression

The final expression of \(f_{c}\) and of \(f_{s}\) is obtained by substituting (51) and (62) into (38):

$$\begin{aligned} f_{c}(a,b)= & {} a\left( f_{nl}^{\text {erf}}(kR)+\sum _{n=1}^{N}c_nf_{nl}^{b,n}(R)\right) = a f_{nl}(R)\nonumber \\ \end{aligned}$$
(64a)
$$\begin{aligned} f_{s}(a,b)= & {} b\left( f_{nl}^{\text {erf}}(kR)+\sum _{n=1}^{N}c_nf_{nl}^{b,n}(R)\right) = b f_{nl}(R)\nonumber \\ \end{aligned}$$
(64b)

with \(u_n=\hat{u}_n/2\), the constant \(k=\mu /\kappa \sqrt{\pi /8}\) as defined in (40), the value of \(R\) is defined in (42a) and

$$\begin{aligned} f_{nl}(R)\equiv f_{nl}^{\text {erf}}(kR)+\sum _{n=1}^{N}c_nf_{nl}^{b,n}(R) \end{aligned}$$
(65)

where \(f_{nl}^{\text {erf}}\) and \(f_{nl}^{b,n}\) have been defined, respectively, in (52) and (63). The two terms \(f_{c}\) and \(f_{s}\) in (64) are symmetric with respect to \(a,b\), since we have \(f_c(a,b)=f_s(b,a)\).

By exploiting the fact that \(\lim _{R\rightarrow 0}J_1(R)/R=1/2\), and then substituting (21), we observe that

$$\begin{aligned} \lim _{R\rightarrow 0}f_{nl}(R)=\mu +\sum _{j=0}^Nc_n u_n=\mu +\frac{1}{2}\sum _{j=0}^Nc_n\hat{u}_n\equiv \beta \end{aligned}$$
(66)

where we substituted the property (21) in the last passage.

It can be proved that the first derivative at zero is

$$\begin{aligned} \lim _{R\rightarrow 0}\frac{\partial f_{nl}}{\partial R}(R)=0, \end{aligned}$$
(67)

meaning that \(f_{nl}\) is constant at first order in \(R\).

In the case of the input described by a single sinusoid \(A\cos (\omega t)\), it is sufficient to set \(A_1=A\), \(\varphi _1=\varphi _2=A_2=0\). In the definitions (36), this leads to \(a=A,b=0\), and the substitution of these in (64) leads to

$$\begin{aligned} f_{c}(a,b)= & {} A\left( f_{nl}^{\text {erf}}(kA)+\sum _{n=1}^{N}c_nf_{nl}^{b,n}(A)\right) \nonumber \\= & {} A f_{nl}(A) \end{aligned}$$
(68a)
$$\begin{aligned} f_{s}(a,b)= & {} 0 \end{aligned}$$
(68b)

The component in quadrature with the input signal \(u\) is zero, and \(N(A)\) is real valued. Substituting (68) in (37), because (37) is equal to \(N(A,\omega )A\), we obtain

$$\begin{aligned} N(A,\omega )=&f_{nl}(A) \end{aligned}$$
(69)

with \(f_{nl}\) matching the RHS of (22).

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Ghirardo, G., Ćosić, B., Juniper, M.P. et al. State-space realization of a describing function. Nonlinear Dyn 82, 9–28 (2015). https://doi.org/10.1007/s11071-015-2134-x

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s11071-015-2134-x

Keywords

Navigation