Skip to main content
Log in

Analytical prediction of delayed Hopf bifurcations in a simplified stochastic model of reed musical instruments

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

Abstract

This paper investigates the dynamic behavior of a simplified single reed instrument model subject to a stochastic forcing of white noise type when one of its bifurcation parameters (the dimensionless blowing pressure) increases linearly over time and crosses the Hopf bifurcation point of its trivial equilibrium position. The stochastic slow dynamics of the model is first obtained by means of the stochastic averaging method. The resulting averaged system reduces to a non-autonomous one-dimensional Itô stochastic differential equation governing the time evolution of the mouthpiece pressure amplitude. Under relevant approximations the latter is solved analytically treating separately cases where noise can be ignored and cases where it cannot. From that, two analytical expressions of the bifurcation parameter value for which the mouthpiece pressure amplitude gets its initial value back are deduced. These special values of the bifurcation parameter characterize the effective appearance of sound in the instrument and are called deterministic dynamic bifurcation point if the noise can be neglected and stochastic dynamic bifurcation point otherwise. Finally, for illustration and validation purposes, the analytical results are compared with direct numerical integration of the model in both deterministic and stochastic situations. In each considered case, a good agreement is observed between theoretical results and numerical simulations, which validates the proposed analysis.

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

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Fig. 1
Fig. 2
Fig. 3
Fig. 4
Fig. 5
Fig. 6
Fig. 7
Fig. 8
Fig. 9
Fig. 10
Fig. 11

Similar content being viewed by others

Availability of data and material

On request.

Notes

  1. In this case the instrument is modeled by a difference equation similar to the logistic map. This model has been thoroughly studied in musical acoustics for control parameters constant over time [13, 35, 36].

  2. We recall that the ensemble average consists in repeating the same measurement many times, and in calculating the average over them.

  3. Sometimes called exit value in the literature.

  4. Indeed, at the bifurcation one has \(a(0)=\frac{\partial F}{\partial x}(0,0)=0\).

  5. This corresponds to the fact that the solutions of Eq. (53) for both cubic and quadratic approximations become complex for \(\sigma >\sigma _{II/III}\).

  6. All numerical simulations of the Itô stochastic differential equations are performed using the function ItoProcess of the Wolfram Mathematica software.

References

  1. Backus, J.: Small-vibration theory of the clarinet. J. Acoust. Soc. Am. 35(3), 305–313 (1963). https://doi.org/10.1121/1.1918458

    Article  Google Scholar 

  2. Baesens, C.: Noise effect on dynamic bifurcations: The case of a period-doubling cascade. In: Dynamic Bifurcations. Lecture Notes in Mathematics, vol. 1493, pp. 107–130. Springer, Berlin (1991)

    Chapter  Google Scholar 

  3. Baesens, C.: Slow sweep through a period-doubling cascade: delayed bifurcations and renormalisation. Physica D 53, 319–375 (1991)

    Article  MathSciNet  Google Scholar 

  4. Benade, A.H.: Fundamentals of Musical Acoustics. Oxford University Press, Oxford (1976)

    Google Scholar 

  5. Bergeot, B., Almeida, A., Vergez, C., Gazengel, B.: Prediction of the dynamic oscillation threshold in a clarinet model with a linearly increasing blowing pressure. Nonlinear Dyn. 73(1–2), 521–534 (2013). https://doi.org/10.1007/s11071-013-0806-y

    Article  MathSciNet  Google Scholar 

  6. Bergeot, B., Almeida, A., Vergez, C., Gazengel, B.: Prediction of the dynamic oscillation threshold in a clarinet model with a linearly increasing blowing pressure: influence of noise. Nonlinear Dyn. 74(3), 591–605 (2013). https://doi.org/10.1007/s11071-013-0991-8

    Article  MathSciNet  Google Scholar 

  7. Bergeot, B., Almeida, A., Vergez, C., Gazengel, B., Ferrand, D.: Response of an artificially blown clarinet to different blowing pressure profiles. J. Acoust. Soc. Am. 135(1), 479–490 (2014)

    Article  Google Scholar 

  8. Berglund, N., Gentz, B.: Pathwise description of dynamic pitchfork bifurcations with additive noise. Probab. Theory Related Fields 122(3), 341–388 (2002). https://doi.org/10.1007/s004400100174

    Article  MathSciNet  MATH  Google Scholar 

  9. Berglund, N., Gentz, B.: Noise-Induced Phenomena in Slow-fast Dynamical Systems. Probability and its Applications (New York). Springer, London (2006). A sample-paths approach

  10. Chaigne, A., Kergomard, J.: Acoustics of Musical Instruments. Modern Acoustics and Signal Processing. Springer, New York (2016). https://doi.org/10.1007/978-1-4939-3679-3

    Book  Google Scholar 

  11. Colinot, T.: Numerical Simulation of Woodwind Dynamics: Investigating Nonlinear Sound Production Behavior in Saxophone-like Instruments. Ph.D. thesis, Aix-Marseille Université (2020)

  12. Colinot, T., Guillemain, P., Vergez, C., Doc, J.B., Sanchez, P.: Multiple two-step oscillation regimes produced by the alto saxophone. J. Acoust. Soc. Am. 147(4), 2406–2413 (2020). https://doi.org/10.1121/10.0001109

    Article  Google Scholar 

  13. Dalmont, J.P., Frappé, C.: Oscillation and extinction thresholds of the clarinet: comparison of analytical results and experiments. J. Acoust. Soc. Am. 122(2), 1173–1179 (2007). https://doi.org/10.1121/1.2747197

    Article  Google Scholar 

  14. Dalmont, J.P., Gilbert, J., Kergomard, J., Ollivier, S.: An analytical prediction of the oscillation and extinction thresholds of a clarinet. J. Acoust. Soc. Am. 118(5), 3294–3305 (2005)

    Article  Google Scholar 

  15. Dalmont, J.P., Gilbert, J., Ollivier, S.: Nonlinear characteristics of single-reed instruments: quasistatic volume flow and reed opening measurements. J. Acoust. Soc. Am. 114(4), 2253–2262 (2003). https://doi.org/10.1121/1.1603235

    Article  Google Scholar 

  16. Fletcher, N.: Nonlinear theory of musical wind instruments. Appl. Acoust. 30(2–3), 85–115 (1990). https://doi.org/10.1016/0003-682X(90)90040-2

    Article  Google Scholar 

  17. Fletcher, N.H., Rossing, T.D.: The Physics of Musical Instruments. Springer, New York (1991)

    Book  Google Scholar 

  18. Fréour, V., Guillot, L., Masuda, H., Usa, S., Tominaga, E., Tohgi, Y., Vergez, C., Cochelin, B.: Numerical continuation of a physical model of brass instruments: application to trumpet comparisons. J. Acoust. Soc. Am. 148(2), 748–758 (2020). https://doi.org/10.1121/10.0001603

    Article  Google Scholar 

  19. Gilbert, J., Maugeais, S., Vergez, C.: Minimal blowing pressure allowing periodic oscillations in a simplified reed musical instrument model: Bouasse–Benade prescription assessed through numerical continuation. Acta Acustica 4(27), 12 (2020). https://doi.org/10.1051/aacus/2020026

    Article  Google Scholar 

  20. Hirschberg, A.: Aero-acoustics of wind instruments. In: Mechanics of musical instruments by A. Hirschberg/ J. Kergomard/ G. Weinreich, vol. 335 of CISM Courses and lectures, chap. 7, pp. 291–361. Springer (1995)

  21. Jansons, K.M., Lythe, G.D.: Stochastic calculus: application to dynamic bifurcations and threshold crossings. J. Statist. Phys. 90(1), 227–251 (1998). https://doi.org/10.1023/A:1023207919293

    Article  MathSciNet  MATH  Google Scholar 

  22. Kergomard, J.: Elementary considerations on reed-instrument oscillations. In: Mechanics of Musical Instruments by A. Hirschberg/ J. Kergomard/ G. Weinreich, vol. 335 of CISM Courses and lectures, chap. 6, pp. 229–290. Springer (1995)

  23. Kergomard, J., Ollivier, S., Gilbert, J.: Calculation of the spectrum of self-sustained oscillators using a variable truncation method: Application to cylindrical reed instruments. Acustica 86(4), 685–703 (2000)

    Google Scholar 

  24. Khas’minskii, R.Z.: A limit theorem for the solutions of differential equations with random right-hand sides. Theory of Probability & Its Applications 11(3), 390–406 (1966). https://doi.org/10.1137/1111038

  25. Klebaner, F.C.: Introduction to Stochastic Calculus with Applications, 2nd edn. Imperial College Press, London (2005)

    Book  Google Scholar 

  26. Øksendal, B.: Stochastic Differential Equations: An Introduction with Applications, 6th edn. Universitext. Springer, Berlin (2003)

    Book  Google Scholar 

  27. Ollivier, S., Dalmont, J.P., Kergomard, J.: Idealized models of reed woodwinds. part 1 : Analogy with bowed string. Acta. Acust. Acust. 90, 1192–1203 (2004)

  28. Roberts, J.B., Spanos, P.D.: Stochastic averaging: an approximate method of solving random vibration problems. Int. J. Non-Linear Mech. 21(2), 111–134 (1986). https://doi.org/10.1016/0020-7462(86)90025-9

    Article  MathSciNet  MATH  Google Scholar 

  29. Silva, F., Kergomard, J., Vergez, C., Gilbert, J.: Interaction of reed and acoustic resonator in clarinet-like systems. J. Acoust. Soc. Am. 124(5), 3284–3295 (2008)

    Article  Google Scholar 

  30. Silva, F., Kergomard, J., Vergez, C., Gilbert, J.: Interaction of reed and acoustic resonator in clarinetlike systems. J Acoust. Soc. Am. 124(5), 3284–3295 (2008). https://doi.org/10.1121/1.2988280

    Article  Google Scholar 

  31. Silva, F., Vergez, C., Guillemain, P., Kergomard, J., Debut, V.: MoReeSC: a framework for the simulation and analysis of sound production in reed and brass instruments. Acta Acustica United Acustica 100(1), 126–138 (2014). https://doi.org/10.3813/AAA.918693

    Article  Google Scholar 

  32. Spiegel, M., Lipschutz, S., Liu, J.: Mathematical Handbook of Formulas and Tables (page 13), 4th edn. McGraw-Hill, New York (2012). https://doi.org/10.1036/9780071795388

  33. Stocks, N.G., Mannella, R., McClintock, P.V.: Influence of random fluctuations on delayed bifurcations: the case of additive white noise. Phys. Rev. A 40(9), 5361–5369 (1989). https://doi.org/10.1103/PhysRevA.40.5361

    Article  Google Scholar 

  34. Stratonovich, R.L.: Topics In the Theory of Random Noise, vol. 1, chap. 4. Taylor & Francis (1963)

  35. Taillard, P.A., Kergomard, J.: An analytical prediction of the bifurcation scheme of a Clarinet-Like Instrument: effects of resonator losses. Acta Acustica United Acustica 101(2), 279–291 (2015). https://doi.org/10.3813/AAA.918826

    Article  Google Scholar 

  36. Taillard, P.A., Kergomard, J., Laloë, F.: Iterated maps for clarinet-like systems. Nonlinear Dyn. 62, 253–271 (2010)

    Article  MathSciNet  Google Scholar 

  37. Terrien, S., Vergez, C., Fabre, B.: Flute-like musical instruments: a toy model investigated through numerical continuation. J. Sound Vibr. 332(15), 3833–3848 (2013). https://doi.org/10.1016/j.jsv.2013.01.041

    Article  Google Scholar 

  38. Wilson, T.A., Beavers, G.S.: Operating modes of the clarinet. J. Acoust. Soc. Am. 56(2), 653–658 (1974)

    Article  Google Scholar 

Download references

Funding

Not applicable.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Baptiste Bergeot.

Ethics declarations

Conflict of interest

The authors declare that they have no conflict of interest concerning the publication of this manuscript.

Code availability

On request.

Additional information

Publisher's Note

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

Appendices

A The 1-dimensional Itô’s formula

Let the following Itô differential equation

$$\begin{aligned} \mathrm {d}x_t=m(x_t,t) \mathrm {d}t+\sigma (x_t,t) \mathrm {d}W_t, \end{aligned}$$
(68)

where m and \(\sigma \) are real functions and \(W_t\) is the so-called Wiener process.

Let \(f(x_t,t)\in \mathcal {C}^{2}(\mathbb {R},\mathbb {R}^+)\) (i.e., f is twice continuously differentiable on \((\mathbb {R},\mathbb {R}^+)\)). Then \(f(x_t,t)\) is also Itô process, and

$$\begin{aligned} \mathrm {d}f(x_t,t)&=\frac{\partial f}{\partial t}\mathrm {d}t + \frac{\partial f}{\partial x}\mathrm {d}x_t + \frac{1}{2}\frac{\partial ^2 f}{\partial x^2}\sigma (x_t,t)^2\mathrm {d}t \end{aligned}$$
(69a)
$$\begin{aligned}&= \left( \frac{\partial f}{\partial t} + m(x_t,t) \frac{\partial f}{\partial x} + \frac{\sigma (x_t,t) ^2}{2}\frac{\partial ^2 f}{\partial x^2}\right) \mathrm{d}t\nonumber \\&\quad + \sigma (x_t,t) \frac{\partial f}{\partial x}\,\mathrm{d}W_t. \end{aligned}$$
(69b)

Equation (69) is the 1-dimensional Itô’s formula (for more details and proof see [26], Chap. 4).

B General formulation of the stochastic averaging method

In this appendix, the stochastic averaging method [24, 34] is briefly described. For that we consider the following system of differential equations in standard form

$$\begin{aligned} \dot{\mathbf{x}}_t=\mathbf{f}(\mathbf{x}_t,t)+\mathbf{g}(\mathbf{x}_t,t)\varvec{\eta }_t \end{aligned}$$
(70)

where \(\mathbf{x}_t \in \mathbb {R}^n\). If the deterministic vector function \(\mathbf{f}(\mathbf{x}_t,t)\in \mathbb {R}^n\) and matrix function \(\mathbf{g}(\mathbf{x}_t,t)\in \mathbb {R}^n \times \mathbb {R}^n\) satisfy certain requirements [24] and the elements of the vector \(\varvec{\eta }_t\) are broadband processes, with zero means, then the slow (or averaged) dynamics of Eq. (70) may be approximated by the following Itô equations

$$\begin{aligned} \mathrm {d} \mathbf{x}_t=\mathbf{m}(\mathbf{x}_t ) \mathrm {d}t+\varvec{\sigma }(\mathbf{x}_t,t) \mathrm {d}{} \mathbf{W}_t, \end{aligned}$$
(71)

where \(\mathbf{W}_t\in \mathbb {R}^n\) is a vector of n Wiener processes. The vector \(\mathbf{m}\) and the matrix \(\varvec{\sigma }\) are called drift vector and diffusion matrix, respectively, and defined by

$$\begin{aligned} \mathbf{m} = T^{\text {av}} \left\{ \mathbf{f}+ \int _{-\infty }^0 \mathbb {E}\left[ \left( \frac{\partial {(\mathbf{g} \varvec{\eta })}}{\partial \mathbf{x} } \right) _t (\mathbf{g} \varvec{\eta })_{t+\tau }\right] \mathrm {d}\tau \right\} , \nonumber \\ \end{aligned}$$
(72)

where \(\frac{\partial {(\mathbf{g} \varvec{\eta })}}{\partial \mathbf{x}}\) is the Jacobian matrix of \(\mathbf{g} \varvec{\eta }\), and

$$\begin{aligned} \varvec{\sigma } \varvec{\sigma }^T= T^{\text {av}} \left\{ \int _{-\infty }^{+\infty } \mathbb {E}\left[ (\mathbf{g} \varvec{\eta })_{t} (\mathbf{g} \varvec{\eta })_{t+\tau }^T\right] \mathrm {d}\tau \right\} \nonumber \\ \end{aligned}$$
(73)

where \(\{.\}^T\) and \(\mathbb {E}\left[ \{.\}\right] \) denotes, respectively, the transpose and the expected value of \(\{.\}\). \( T^{\text {av}}\) is an averaging operator defined as follows

$$\begin{aligned} T^{\text {av}} \left\{ . \right\} = \lim \limits _{T \rightarrow +\infty } \frac{1}{T}\int _{t_0}^{t_0+T}\left\{ .\right\} \mathrm {d}t. \end{aligned}$$
(74)

It should be noted that in the case of a periodic variables with period \(T_0\) (which is the case in this paper), the operator \(T^{\text {av}}\) becomes a classical Krylov–Bogolyubov time averaging over one period \(T_0\), i.e.,

$$\begin{aligned} T^{\text {av}} \left\{ . \right\} = \frac{1}{T_0}\int _{t_0}^{t_0+T_0}\left\{ .\right\} \mathrm {d}t \end{aligned}$$
(75)

and the result is independent of \(t_0\).

C Derivation of the expression of the deterministic dynamic bifurcation point

In this appendix, we give the details of the résolution of Eq. (41). First we state \(y+\hat{\gamma }^{\text {st}}=X^2\) and then it can be shown that, using Eq. (34), Eq. (41) takes the following form

$$\begin{aligned}&\left( X-X_0\right) \Big (\zeta F_1 X^2+ (\zeta F_1X_0-\alpha _1\omega _1)X\nonumber \\&\quad -\zeta F_1+\zeta F_1 X_0^2-\alpha _1 X_0 \omega _1\Big )=0 \end{aligned}$$
(76)

where \(X_0^2=y_0+\hat{\gamma }^{\text {st}}\). Obviously \(X=X_0\) and therefore \(y=y_0\) is a solution of (76). The second term of the product in the left-hand side of (76) is a second order polynomial equations whose roots \(X_1\) and \(X_2\) are

$$\begin{aligned} X_{1}&=\frac{1}{2 \zeta F_1}\Bigg (\alpha _1 \omega _1-\zeta F_1 X_0\nonumber \\&\quad -\sqrt{\alpha _1^2 \omega _1^2+2 \alpha _1 \zeta F_1 X_0 \omega _1+\zeta ^2 F_1^2 \left( 4-3 X_0^2\right) }\Bigg ) \end{aligned}$$
(77)

and

$$\begin{aligned} X_2&=\frac{1}{2 \zeta F_1}\Bigg (\alpha _1 \omega _1-\zeta F_1 X_0\nonumber \\&\quad +\sqrt{\alpha _1^2 \omega _1^2+2 \alpha _1 \zeta F_1 X_0 \omega _1+\zeta ^2 F_1^2 \left( 4-3 X_0^2\right) }\Bigg ). \end{aligned}$$
(78)

The initial value \(y_0\) is always chosen to be larger than \(-\hat{\gamma }^{\text {st}}\) (because the mouth pressure \(\gamma \) must be larger than zero). Therefore, X must be larger than zero and, for a realistic set of parameters, only \(X_2\) is positive. Consequently, the expression of the deterministic dynamic bifurcation point is given by

$$\begin{aligned} {\hat{y}}^{\text {dyn}}_{\text {det}}=X_2^2-\hat{\gamma }^{\text {st}}. \end{aligned}$$
(79)

D Derivation of the expression of the stochastic dynamic bifurcation point

Using Eqs. (34), (53) takes the following explicit form

$$\begin{aligned} K=\frac{\zeta F_1\sqrt{\hat{\gamma }^{\text {st}} +y}}{2 \omega _1}\left( y-1+\hat{\gamma }^{\text {st}}\right) -\frac{\alpha _1 y}{2}. \end{aligned}$$
(80)

To obtain a solvable cubic form (i.e., without square root), Eq. (80) is transformed into

$$\begin{aligned} \left( K+\frac{\alpha _1 y}{2}\right) ^2= \frac{\zeta ^2 F_1^2\left( \hat{\gamma }^{\text {st}} +y\right) }{4 \omega _1^2} \left( y-1+\hat{\gamma }^{\text {st}}\right) ^2 \end{aligned}$$
(81)

which yields Eq.  (55).

In the remaining of this appendix the Cardano’s method (see, e.g., [32]) is used to solve the latter, i.e., \( a_1y^3+a_2y^2+a_3y+a_4=0. \)

First, the following parameters are introduced

$$\begin{aligned} p= & {} -\frac{a_2^2}{3a_1^2}+\frac{c_3}{a_1}\quad \text {and}\quad \\ q= & {} \frac{a_2}{27a_1}\left( \frac{2a_2^2}{a_1^2}-\frac{9a_3}{a_1}\right) +\frac{a_4}{a_1}. \end{aligned}$$

The discriminant \(\varDelta \) is defined as \( \varDelta =-\left( 4 p^3+27 q^2\right) . \) Then:

  1. 1.

    If \(\varDelta <0\), one root is real and two are complex conjugate.

  2. 2.

    If \(\varDelta =0\), all roots are real and at least two are equal.

  3. 3.

    If \(\varDelta >0\), all roots are real and unequal.

A typical example of the discriminant \(\varDelta \), plotted as a function of the noise level \(\sigma \), is depicted in Fig. 12 for a typical set of parameters. It can be shown that \(\varDelta \) can be expressed as a fourth-order polynomial equation with respect to \(\ln \sigma \) which has two distinct roots

$$\begin{aligned} \ln \left( \sigma _1\right)&=\frac{2 \alpha _1^3 \omega _1^3-27 \alpha _1 \hat{\gamma }^{\text {st}} \zeta ^2 F_1^2 \omega _1}{54 \zeta ^2 F_1^2 \omega _1 \epsilon }\nonumber \\&\quad -\frac{2 \left( \alpha _1^2 \omega _1^2+3 \zeta ^2 F_1^2\right) ^{3/2}-9 \alpha _1 \zeta ^2 F_1^2 \omega _1}{54 \zeta ^2 F_1^2 \omega _1 \epsilon }\nonumber \\&\quad -\frac{1}{2} \ln \left( \frac{2 \sqrt{2 \pi } \hat{\gamma }^{\text {st}^{3/4}} \omega _1 e^{-\frac{(\hat{\gamma }^{\text {st}} -1) \sqrt{\hat{\gamma }^{\text {st}} } \zeta F_1}{\omega _1 \epsilon }}}{\sqrt{(3 \hat{\gamma }^{\text {st}} +1) \zeta F_1 \omega _1 \epsilon }}\right) \nonumber \\&\quad +\ln \left( x_{y_0}\right) \end{aligned}$$
(82)

and

$$\begin{aligned} \ln \left( \sigma _2\right)&=\frac{9 \alpha _1 (1-3 \hat{\gamma }^{\text {st}} ) +\frac{2 \left( \alpha _1^3 \omega _1^3+\left( \alpha _1^2 \omega _1^2+3 \zeta ^2 F_1^2\right) {}^{3/2}\right) }{\zeta ^2 F_1^2 \omega _1}}{54 \epsilon }\nonumber \\&\quad -\frac{1}{2} \ln \left( \frac{2 \sqrt{2 \pi } \hat{\gamma }^{\text {st}^{3/4}} \omega _1 e^{-\frac{(\hat{\gamma }^{\text {st}} -1) \sqrt{\hat{\gamma }^{\text {st}} } \zeta F_1}{\omega _1 \epsilon }}}{\sqrt{(3 \hat{\gamma }^{\text {st}} +1) \zeta F_1 \omega _1 \epsilon }}\right) \nonumber \\&\quad +\ln \left( x_{y_0}\right) \end{aligned}$$
(83)

and a double root

$$\begin{aligned} \ln \left( \sigma _3\right)&=\frac{\alpha _1 (\hat{\gamma }^{\text {st}} -1 )}{2 \epsilon }\nonumber \\&\quad -\frac{\log \left( \frac{2 \sqrt{2 \pi } \hat{\gamma }^{\text {st}^{3/4} }\omega _1 e^{-\frac{(\hat{\gamma }^{\text {st}} -1) \sqrt{\hat{\gamma }^{\text {st}} } \zeta F_1}{\omega _1 \epsilon }}}{\sqrt{(3 \hat{\gamma }^{\text {st}} +1) \zeta F_1 \omega _1 \epsilon }}\right) }{2}\nonumber \\&\quad + \ln \left( x_{y_0}\right) . \end{aligned}$$
(84)

In the example shown in Fig. 12 one has: \(\ln \left( \sigma _1\right) =-53.1\), \(\ln \left( \sigma _2\right) =-6.8\) and \(\ln \left( \sigma _3\right) =-26.6\).

Fig. 12
figure 12

Discriminant \(\varDelta \) of (55) as a function the natural logarithm of the noise level \(\ln \sigma \). The set of parameters (45) is used

Fig. 13
figure 13

The roots \(r_k\) (\(k=0,1,2\)) of Eq. (55) and the the stochastic dynamic bifurcation \({\hat{y}}^{\text {dyn}}_{\text {stoch,a}}\) as functions of the natural logarithm of the noise level \(\sigma \). The parameters (45) are used and \(x_{y_0}=0.01\)

If \(\sigma _1<\sigma <\sigma _2\) we have \(\varDelta >0\) except at \(\sigma =\sigma _3\) for which \(\varDelta =0\). In general, when \(\varDelta >0\) the three real roots are written using trigonometric functions as follows

$$\begin{aligned} r_k= & {} 2 \sqrt{\frac{-p}{3}} \cos \left( \frac{1}{3}\arccos {\left( \frac{3q}{2p}\sqrt{\frac{3}{-p}}\right) }\right. \nonumber \\&+ \left. \frac{2k\pi }{3}\right) -\frac{a_2}{3a_1}\quad \text {with}\quad k=0,1,2. \end{aligned}$$
(85)

Fig. 13 shows the roots \(r_k\) (\(k=0,1,2\)) as functions of the noise level \(\sigma \) using again the parameters (45) and \(x_{y_0}=0.01\). The stochastic dynamic bifurcation point, denoted \({\hat{y}}^{\text {dyn}}_{\text {stoch,a}}\), is equal to \(r_0\) if \(\sigma <\sigma _3\) and to \(r_2\) if \(\sigma _3<\sigma <\sigma _2\). This choice is justified by means of a numerical resolution which shows that this is the unique positive solution of Eq. (80).

E Static bifurcation diagram

The static bifurcation diagram is the result of the bifurcation analysis of a deterministic dynamical system with constant bifurcation parameter, here Eq. (6) with a constant value of y. It plots, as a function of the considered bifurcation parameter (here y), the possible steady-state regimes (fixed points and periodic motions) indicating their stability.

We use here the averaging procedure to obtain the approximated analytical bifurcation diagram of the one-mode model. For that Eq. (27) is considered without noise and with a constant bifurcation parameter y, i.e.,

$$\begin{aligned} \frac{\mathrm {d}x_t}{\mathrm {d}t}=F(x_t,y), \end{aligned}$$
(86)

where the function \(F(x_t,y_t)\) is given by Eq. (20).

The fixed points \(x^e\) of (86) are obtained by solving \( F(x,y)=0. \) We obtain three solutions: the trivial solution \(x^e_1=0\) and two non-trivial solutions, one is negative and one is positive. Only the positive non-trivial solution is retained and denoted \(x^e_2\), its expression is

$$\begin{aligned} x^e_2=4\sqrt{\frac{\frac{2 \zeta F_1 (3 \hat{\gamma }^{\text {st}} +3 y-1)}{3 \omega _1 \sqrt{\hat{\gamma }^{\text {st}} +y}}-\frac{4 \alpha _1}{3}}{\frac{\zeta F_1 (\hat{\gamma }^{\text {st}} +y+1)}{\omega _1 (\hat{\gamma }^{\text {st}} +y)^{5/2}}}} \end{aligned}$$
(87)

where the expression of \({\hat{y}}^{\text {st}}\) is given by Eq. (7). In the lossless case with \(\alpha _1=0\) and \(\hat{\gamma }^{\text {st}}=\frac{1}{3}\), Eq. (87) reduces to \(x^e_2=4 \sqrt{2} (3 y+1) \sqrt{\frac{y}{9 y+12}}\).

The trivial solution corresponds to zero equilibrium position of (6), whereas the non-trivial solution characterizes its periodic steady-state regimes.

As we know, the trivial fixed point \(x^e_1\) is stable if \(y<0\) and unstable if \(y>0\). One can be shown that the non-trivial fixed point \(x^e_2\) exists only for \(y>0\) and is stable.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Bergeot, B., Vergez, C. Analytical prediction of delayed Hopf bifurcations in a simplified stochastic model of reed musical instruments. Nonlinear Dyn 107, 3291–3312 (2022). https://doi.org/10.1007/s11071-021-07104-9

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s11071-021-07104-9

Keywords

Navigation