Download as pdf or txt
Download as pdf or txt
You are on page 1of 18

Flute-like musical instruments: a toy model investigated

through numerical continuation


Soizic Terrien, Christophe Vergez, Benoît Fabre

To cite this version:


Soizic Terrien, Christophe Vergez, Benoît Fabre. Flute-like musical instruments: a toy model inves-
tigated through numerical continuation. Journal of Sound and Vibration, Elsevier, 2013, 332 (15),
pp.3833-3848. �10.1016/j.jsv.2013.01.041�. �hal-00721873v2�

HAL Id: hal-00721873


https://hal.archives-ouvertes.fr/hal-00721873v2
Submitted on 6 Feb 2013

HAL is a multi-disciplinary open access L’archive ouverte pluridisciplinaire HAL, est


archive for the deposit and dissemination of sci- destinée au dépôt et à la diffusion de documents
entific research documents, whether they are pub- scientifiques de niveau recherche, publiés ou non,
lished or not. The documents may come from émanant des établissements d’enseignement et de
teaching and research institutions in France or recherche français ou étrangers, des laboratoires
abroad, or from public or private research centers. publics ou privés.

Distributed under a Creative Commons Attribution - NonCommercial - NoDerivatives| 4.0


International License
Flute-like musical instruments: a toy model investigated through
numerical continuation
S. Terriena,1 , C. Vergeza , B. Fabreb
a
Laboratoire de Mécanique et d’Acoustique (LMA-CNRS UPR 7051),
Aix Marseille Université,
31 chemin Joseph Aiguier, 13402 Marseille cedex 20, France
b
LAM, Institut Jean Le Rond d’Alembert (CNRS UMR 7190),
UPMC Univ. Paris 6,
11 rue de Lourmel, 75015 Paris, France

Abstract

Self-sustained musical instruments (bowed string, woodwind and brass instruments) can be mod-
eled by nonlinear lumped dynamical systems. Among these instruments, flutes and flue organ pipes
present the particularity to be modeled as a delay dynamical system. In this paper, such a sys-
tem, a toy model of flute-like instruments, is studied using numerical continuation. Equilibrium
and periodic solutions are explored with respect to the blowing pressure, with focus on amplitude
and frequency evolutions along the different solution branches, as well as "jumps" between periodic
solution branches. The influence of a second model parameter (namely the inharmonicity) on the
behaviour of the system is addressed. It is shown that harmonicity plays a key role in the presence
of hysteresis or quasi-periodic regime. Throughout the paper, experimental results on a real instru-
ment are presented to illustrate various phenomena, and allow some qualitative comparisons with
numerical results.

Keywords: Musical acoustics, Flute-like instruments, Numerical continuation, Periodic solutions,


Nonlinear delay dynamical system

1 Introduction the behaviour of simple models of flute-like in-


struments, no study, to the authors knowledge,
Sound production in self-sustained musical in- has ever investigated the determination of solu-
struments, like bowed string instruments, reed tion branches using numerical continuation, and
instruments or flutes, involves the conversion of the analysis of jumps between branches. This is
a quasi-static energy source (provided by the done in this paper.
musician) into acoustic energy. This generation
A numerical continuation approach has re-
of auto-oscillations necessarily implies nonlinear
cently proved to be useful in the context of mu-
mechanisms. Thus, even the simplest models of
sical instruments, especially for reed instruments
self-sustained musical instruments should include
(like the clarinet) [8]. However, unlike reed in-
nonlinear terms [1].
struments, flute-like instruments are delay dy-
We focus in this paper on flute-like instru- namical systems (see section 2), which compli-
ments. Many works with both experimental and cates the model analysis, and prevents the use of
modeling approaches have highlighted the wide the same numerical tools (as, for example, Auto
variety of their oscillation regimes (for example [9] or Manlab [10]).
[2, 3]), and aim to explore the complexity of their
dynamics [4, 5, 6]. Although different studies Using a numerical continuation software dedi-
[1, 7, 3] has dealt with numerical investigation of cated to delay dynamical systems [11], we study
in this paper a dynamical system inspired by the
1
Corresponding author: terrien@lma.cnrs-mrs.fr physics of flute-like instruments. We call it a

1
toy model, since it is known in musical acoustics
that more accurate (much more complex) mod-
els should be considered [12, 13]. We investigate
throughout the paper the diversity of oscillation
regimes of the toy model, as well as bifurcations
and jumps between branches. We particularly
focus on the nature of the solutions (static or
oscillating), and in the later case, on the evolu-
tion of frequency and amplitude along periodic
solution branches. Indeed, in a musical context,
periodic solutions correspond to notes, and oscil- Figure 1: Recorder section, and simplified repre-
lating frequency and amplitude are respectively sentation of the jet oscillation around the labium,
related to the pitch and the intensity of the note in the mouth of the resonator.
played. Comparisons between the obtained bi-
furcation diagrams, experimental data and time-
system can be interpreted as an extremely sim-
domain simulations highlight the valuable con-
plified model of this kind of musical instruments.
tribution of numerical continuation. Moreover,
surprisingly enough, the qualitative behaviour of
the toy model displays close similarities to that
2.1 Sound production in flute-like in-
of a real instrument.
struments
The paper is structured as follows. The dy-
namical system under study is presented in sec- Since the work of Helmholtz [14], a classical ap-
tion 2. In section 3, we present the results of proach consists in modeling a musical instrument
numerical continuation of the branches of static by a nonlinear coupling between an exciter and
and periodic solutions and their stability analy- a resonator. In other wind instruments, the ex-
sis, with focus on amplitude and frequency evo- citer involves the vibration of a solid element:
lutions. The study of "jumps" between different a cane reed in the clarinet or the saxophone,
periodic solution branches is presented in section the musician’s lips in the trumpet or the trom-
4. Finally, we discuss, in section 5, the influence bone... In flute-like instruments (which includes
of a parameter related to the instrument makers’ recorders, transverse flutes, flue organ pipes...),
work on the characteristics of the different oscil- the exciter consists in the oscillation of the blown
lating regimes. Qualitative comparisons with the air jet around an edge called labium (see figure
sound produced by real instruments and with the 1) [15, 13].
experience of an instrument maker are presented. The jet-labium interaction generates the
pressure source which excites the resonator
formed by the pipe, and thus creates an acoustic
2 System studied (toy model) field in the instrument. In turn, the acoustic
field disturbs the jet at the channel exit (see for
In this paper, we study the following delay dy-
example [16]). As the jet is naturally unstable,
namical system:
this perturbation is convected and amplified
 along the jet [15, 17], from the channel exit to
 z(t) = v(t − τ )
the labium, which sustains the oscillations of
p(t) = α · tanh [z(t)] (1)
the jet around the labium, and thus the sound
V (ω) = Y (ω) · P (ω)

production [12, 13].
where lowercase variables are written in the time
domain, and uppercase variables in the frequency The convection time of perturbations along
domain. The unknowns are z, v, and p, while α the jet introduces a delay in the system, whose
and τ are scalar constant parameters. Y is a value is related to the convection velocity cv of
known function of the frequency, and will be de- these hydrodynamic perturbations. Theoretical
tailed later. We first briefly describe the mech- [15, 17] and experimental [18] studies have shown
anism of sound production in flute-like instru- that cv is related to the jet velocity Uj (itself
ments, and then we precise to which extent this related to the blowing pressure applied by the

2
of the perturbations along the jet, and to the
convection delay defined by equation (3).

Using a modal decomposition of the resonator


acoustical response, the transfer function is writ-
ten in the frequency domain as a sum of m modes
(as it is done, for example, in the work of Silva
[19]):
m
X ak · jω
Figure 2: Basic modeling of sound production Y (ω) = ωk (4)
ω2 − ω 2 + jω ·
k=1 k Qk
mechanism in flute-like instruments, as a system
with a feedback loop [12, 13]. where ak , ωk and Qk are respectively the modal
amplitude, the resonance pulsation and the
quality factor of the k-th mode.
musician) through:

cv ≈ 0.5 · Uj . (2) This representation is very simplified com-


pared to the most recent models of flute-like in-
Noting W the mouth height (i.e. the length of struments [12, 13], which take into account the
the jet, between the channel exit and the labium, following elements:
highlighted in figure 1), an approximation of the
delay value is given by: • a precise modeling of the aeroacoustic
source, which includes, in the latest models,
W W
τ= ≈ . (3) a time derivative of the delayed term [13].
cv 0.5 · Uj
• nonlinear losses, due to vortex shedding at
2.2 System studied: a toy model of the labium [20].
flute-like instruments
• an accurate description of the resonator: in
This mechanism of sound production can be
theory, the modal decomposition in equation
modeled by a nonlinear oscillator, such as the
(4) can include a large number m of modes.
one represented in figure 2 [12, 13]. This very
For sake of simplicity, we only retain in this
basic modeling includes the three following ele-
paper at most the first two modes of the
ments:
pipe.
• a delay, related to the convection time of the
hydrodynamic perturbations along the jet. Although this "toy model" may seem similar
to those studied by McIntyre et al. [1] and Colt-
• a nonlinear function, representing the jet- man [7], an important difference relies on the de-
labium interaction. scription of the resonator: we use here a modal
• a linear transfer function representing the decomposition, whereas the models of McIntyre
acoustic behaviour of the resonator. et al. [1] and Coltman [7] are both based on
a time-domain description of the resonator be-
Therefore, system (1) can be seen as a very basic haviour.
model of flute-like instruments, and thus can be Here, the interest of using a modal descrip-
related to the nonlinear oscillator represented tion is the ability to change independently of each
in figure 2. Variables z, p and v are then other the different resonator parameters (such as,
respectively related to the transversal deflection for example, the resonance frequencies - see sec-
of the jet at the labium, to the pressure source tion 5).
created by the jet-labium interaction, and to the
acoustic velocity induced by the waves at the
2.3 Rewritting as a first-order system
pipe inlet, so-called "mouth of the instrument"
(see figure 1). The scalar parameters α and τ To be analysed with classical numerical continua-
are respectively associated to the amplification tion methods for delay differential equations [21],

3
system (1) should be rewritten as a classical first- Regardless of the parameter values, it is ob-
order delay system ẋ(t) = f (x(t), x(t − τ ), γ). vious that system (6) has a unique equilibrium
In such a formulation, x is the vector of state solution defined by:
variables and γ the set of parameters. To im- 
prove numerical conditioning of the problem, a  ∀i = [1, 2, ..., m] :
v (t̃) = 0 (7)
dimensionless time variable and a dimensionless  i
yi (t̃) = 0.
convection delay are defined :
 A stability analysis of this static solution
t̃ = ω1 t is performed both through computation of the
(5)
τ̃ = ω1 τ eigenvalues of the linearized problem and the
analysis of the open-loop gain (see B for more
with ω1 the first resonance pulsation (see equa- details). Moreover, this linear analysis allows to
tion (4)). Rewritting system (1) finally leads to distinguish two kinds of emerging periodic solu-
the following system of 2m + 2 equations, where tions:
m is the number of modes in the transfer func-
tion Y (ω) (see A for more details): • those resulting from the coupling between
∀ i = [1, ..., m] (where i is an integer), an acoustic mode of the resonator (a partic-
ular term in the sum defining the transfer
function Y (ω) in equation (4)) and the first
 m
X hydrodynamic mode of the jet (correspond-
v(t̃) = vk (t̃)


ing to n = 0 in equation (24) of B).



k=1



 m
• those resulting from the coupling between

 X
y(t̃) = yk (t̃)


an acoustic mode of the resonator and an



k=1
higher-ranked hydrodynamic mode of the jet
v̇i (t̃) =yi (t̃)
(corresponding to n > 0 in equation (24)).


αai 


1 − tanh2 v(t̃ − τ̃ ) · y(t̃ − τ̃ )
 
ẏi (t̃) =





 ω1 Indeed, for flute-like instruments, an auto-
  2

 ωi ωi oscillation results from the coupling between
− vi (t̃) − yi (t̃).


an acoustic mode of the resonator and an

ω1 ω 1 Qi
(6) hydrodynamic mode of the jet [2, 12]. As also
System (6) is studied throughout this paper. Nu- observed experimentally by Verge [2] in the case
merical results are computed using the software of a flue organ pipe, the first case n = 0 corre-
DDE Biftool [11], which performs numerical con- sponds to the "standard" regimes in flute-like
tinuation of delay differential equations using a instruments. On the contrary, the second case
prediction/correction approach [21]. Periodic so- n > 0 corresponds to the so-called "aeolian"
lutions are computed using orthogonal colloca- regimes, for which the time delay τ is larger
tion [22]. Typically, we use about 100 points per than the oscillation period 2π
ω of the system (see
period, and the degree of the Lagrange interpo- equation 24 in B). In the context of musical
lation polynomial is 3 or 4. instruments, it corresponds to particular sounds,
like for example those produced when the wind
machine of an organ is turned off leaving a key
3 Periodic regimes emerging pressed.
from the equilibrium solution
We focus here on these two kinds of peri-
3.1 Branch of equilibrium solutions odic solutions, and we study particularly the
amplitude and frequency evolutions along the
In flute-like instruments, the delay τ is related branches.
to the jet velocity (see equation (3)), and there-
fore to the pressure into the musician’s mouth
3.2 Branches of periodic solutions:
through the Bernoulli equation for stationary
aeolian and standard regimes
flows. To analyze the system (6), it is thus par-
ticularly interesting to choose this variable as the For sake of clarity, we consider in this section
continuation parameter. a transfer function containing a single acoustic

4
30

j
static solution

dimensionless oscillating amplitude v/U


static solution
standard regime (n=0) standard regime (n=0)
oscillating amplitude of variable v
60
25 1st aeolian regime (n=1) n=2
1st aeolian regime (n=1)
2nd aeolian regime (n=2) 2nd aeolian regime (n=2)
50 n = 0 n=1 n=2
stable solutions 20 stable solutions

40 n=1
15
30
10
20
n=0
10 τ*
5 τ*
0 0
0 1 2 3 4 5 6 7 8 0 2 4 6 8 10 12 14 16 18
delay τ (s) −3 dimensionless delay τ ⋅ ω (ω = 2260)
x 10 1 1

Figure 3: Bifurcation diagram of system (6). Figure 4: Bifurcation diagram of system (6).
Abscissa: delay τ (in seconds). Ordinate: am- Abscissa: dimensionless delay τ̃ = ω1 τ . Ordi-
plitude of the oscillating variable v(t). Symbols nate: oscillation amplitude of the dimensionless
× indicate the stable parts of the branches, and variable v(t̃)/Uj . Symbols × indicate the sta-
the dashed line indicate the separation between ble parts of the branches, and the dashed line
the standard regime and the aeolians ones. Pa- indicate the separation between the standard
rameters value: m = 1, α = 10 ; a1 = 70 ; regime and the aeolians ones.. Parameters value:
ω1 = 2260 ; Q1 = 50. Values of n are computed m = 1, α = 10 ; a1 = 70 ; ω1 = 2260 ; Q1 = 50.
using equation (24) in B. Values of n are computed using equation (24) in
B.

mode (i.e. m = 1 in the expression of Y (ω),


rameter τ̃ (see equations 5). Throughout the pa-
defined by equation (4)). The addition of a sec-
per, for sake of consistency, we thus present the
ond acoustic mode will be discussed in the next
results using τ̃ and the dimensionless amplitude
section.
v(t̃)/Uj of the variable v(t̃) (with Uj the jet ve-
Numerical continuation of the different peri-
locity). Thereby, figure 4 represents the same bi-
odic solution branches allows to display the bi-
furcation diagram as figure 3. It is important to
furcation diagram showed in figure 3, represent-
keep in mind that the jet velocity Uj is not a con-
ing the amplitude of the oscillating variable v(t̃)
stant (see equation (3) for the relation between
defined in system (6), with respect to the delay τ .
Uj and the continuation parameter τ ): although
It is useful to keep in mind that blowing harder
the order of magnitude of the dimensionless am-
makes the jet velocity Uj increase and thus τ de-
plitude of the different regimes seems to be dif-
crease (according to equation (3)).
ferent (figure 4), it corresponds to a same value
The stability of each branch is addressed by
of the dimensioned amplitude (figure 3).
computing the Floquet multipliers [21]: since all
of them lie within the unit circle, it can be con-
3.2.1 Amplitude evolutions
cluded that all the branches are stable [23].
As shown in figure 3, the use of equation (24) To study flute-like instruments, it is useful to
(B) allows to distinguish the different kind of pe- define the dimensionless jet velocity θ [24, 2]:
riodic regimes, and highlights that the only stan-
Uj
dard regime (related to the first hydrodynamic θ= (8)
mode of the jet) is located in the part of the W ·f
graph where τ is lower than τ ∗ = 0.9·10−3 s. The where W is the jet length (see figure 1), and f
other branches correspond to aeolian regimes re- the oscillation frequency.
lated to the second and the third hydrodynamic Indeed, the representation, in figure 5, of the
modes of the jet (respectively n = 1 and n = 2). oscillation amplitude of the dimensionless vari-
A larger range in τ in figure 3 would reveal able v(t̃)/Uj (related, in flute-like instruments,
other branches of aeolian sounds. They corre- to the acoustic velocity in the mouth of the res-
spond to the infinite series of aeolian instabilities onator) as a function of θ for each branch of pe-
highlighted for each acoustic mode of the transfer riodic solutions plotted in figure 4 shows a clear
function Y (ω) in B. separation between aeolian and standard regimes
As highlighted previously, numerical calcula- in two different zones of the dimensionless jet ve-
tions are performed using the dimensionless pa- locity θ. This is interesting since for flute-like in-

5
30
j 1st standard regime (n = 0) standard regime (n = 0)
dimensionless oscillating amplitude v/U
1st aeolian regime (n=1) 1.15 1st aeolian regime (n=1)

1
25 2nd aeolian regime (n=2) 2nd aeolian regime (n=2)

Dimensionless frequency f/f


stable solutions 1.1 stable solutions
20
1.05

15 1

10 0.95

0.9
5

0.85
0
0 1 2 3 5 10 20 30 40 50 60 0 1 2 3 5 10 20 30 40 50 60
θ = Uj/(W ⋅ f) θ = Uj/(W ⋅ f)

Figure 5: Bifurcation diagram of the system (6) Figure 6: Bifurcation diagram of the system (6),
with same parameter values as in figure 4. Ab- with same parameter values as in figure 4. Ab-
scissa: dimensionless jet velocity θ. Ordinate: scissa: dimensionless jet velocity θ. Ordinate:
oscillation amplitude of the dimensionless vari- dimensionless oscillation frequency ff1 (with f1 =
ω1
able v(t̃)/Uj . Symbols × indicate the stable 2π the resonance frequency of the transfer func-
parts of the branches. tion Y (ω)). Symbols × indicate the stable parts
of the branches.
struments playing under normal conditions, θ is
expected to be larger than 4 [25]. Figure 5 shows resonance frequency f1 as θ increases fur-
the same trend for the standard regime. On the ther.
contrary, we observe that all aeolian regimes are • For aeolian regimes, one observes an impor-
found for θ < 4. This feature is in agreement tant frequency deviation just after the os-
with the fact that they occur at low jet veloci- cillation threshold, followed by a second fre-
ties. Indeed, for the same oscillation frequency, θ quency deviation after a plateau around the
is lower for an aeolian regime than for a standard resonance frequency (inflection point at f1 ).
regime (as highlighted in section 3, τ f is larger
than unity for aeolian regimes).
Moreover, the bifurcation diagram displayed 3.3 Experimental illustration
in figure 5 shows a particular amplitude evolu-
Because of the important simplification of the
tion for aeolian regimes. Indeed, one notices,
model, a quantitative comparison between ex-
for such regimes, a bell-shaped curve, whereas
perimental and numerical results is not possible.
for the "standard" regime one can observe a sat-
However, a qualitative comparison is interesting,
uration followed by a slow decrease of the am-
and proves that the dynamical system studied
plitude. Thus, while the standard regime exists
in this paper reproduces some key features of
when τ tends to zero (and thus when the jet ve-
the behaviour of flute-like instruments.
locity Uj tends to infinity - see equation (3)), ae-
olian regimes conversely exist only for restricted
Experimentally, aeolian sounds occur for
ranges of these two parameters.
very low blowing pressures (which correspond
to high values of the delay τ ), that is to say
3.2.2 Frequency evolutions when the musician blows very gently in the
Figure 6 presents the dimensionless oscillation instrument. It is why experimental observation
frequency f /f1 (with f1 = 2π ω1
the resonance fre- of such sounds is particularly complicated. The
quency of the transfer function Y (ω)) as a func- use of a pressure-controlled artificial mouth
tion of θ, along the different periodic solution (which permits to play the instrument without
branches shown in figure 4. Similarly to the am- musician, and to control very precisely the
plitude, it reveals different evolution patterns for mouth pressure) provides new information
aeolian and standard regimes: about the dynamics of the instrument [26].
The simultaneous measurement of the mouth
• For the standard regime, one observes an pressure Palim (related to the jet velocity Uj
important frequency deviation just after the by the Bernoulli equation for stationary flows)
oscillation threshold (at θ = 8.3). Then the and the acoustic pressure under the labium Pin
frequency asymptotically tends toward the (see figure 1) allows to study the influence of

6
80 4

alim
A B C increasing pressure ramp − A1

1
mouth pres− dimensionless frequency f/f

dimensionless internal pressure P /P


5 increasing pressure ramp − A2
60 3.5

in
oscillating amplitude of the
increasing pressure ramp − B
4 3 decreasing pressure ramp − B
(f = 615 Hz)
40
decreasing pressure ramp − C1
3 A A C C
(a) 1 2
1 2 20 2.5 decreasing pressure ramp − C2

2
1

2 0

1 1.5
−20
1
0 −40
sure (Pa)

500 0.5
(b)
0
0 2 3 4 5 6 7
0 20 40 60 80 100 120 θ = Uj / (W.f)
time (s)

Figure 7: Time-frequency representation (a) of Figure 8: Same experimental data as in figure


the sound of a Zen-On Bressan plastic alto 7: representation of the dimensionless acoustic
recorder played by an artificial mouth making pressure amplitude measured in the resonator,
an increasing and decreasing pressure ramp (b). with respect to the dimensionless jet velocity θ.
The dark dot-dashed lines indicate the two first Letters in legend refer to letters in figure 7-(a).
resonance frequencies (f1 ≈ 615 Hz and f2 ≈ 842
Hz) of the resonator. G sharp fingering (4th oc-
the oscillating amplitude of the principal regime
tave).
shows a different evolution, comparable to the
one observed in figure 5 for the standard regime.
some parameters on the characteristics of the As far as the frequency is concerned, large varia-
oscillation regimes. tions can be observed in figure 7-(a), particularly
close to the threshold of the standard regime B.
Time-frequency analysis (figure 7-(a)) of the
sound produced by a Zen-On Bressan plastic alto The experimental observation of aeolian
recorder (previously studied by Lyons [27]) dur- sounds on a recorder is especially interesting.
ing both an increasing and a decreasing ramp Indeed, if these particular sounds are well-known
of the mouth pressure (figure 7-(b)) highlights for flue organ pipes (see for example [4, 2]),
several oscillation regimes (zones A and C) be- it is generally assumed that recorders can not
low the threshold of the principal regime corre- operate on the 2nd (or higher) hydrodynamic
sponding to the expected note (zone B). Figure 8 mode of the jet, due to the small value of their
represents, for the same data, the dimensionless ratio W/h (with W the mouth heigh and h
acoustic pressure amplitude Pin /Palim as a func- the channel exit heigh - see figure 1) [2, 12].
tion of θ. As for aeolian regimes on the numerical Indeed, this ratio is close to 4 in recorder-like
bifurcation diagram (figure 5), we observe that instruments whereas it can reach 12 in flue
one of these regimes is located in a zone defined organ pipes [2].
by θ < 4, whereas the principal regime (zone B)
appears for θ > 4. If the second one (regime Thereby, the very simple model studied here
A2-C1 in figures 7-(a) and 8) appears for higher produces aeolian regimes, which present partic-
values of θ (about 5.25 < θ < 6.25), it never- ular features which recall some experimental ob-
theless can be considered as an aeolian regime. servations on a recorder played by an artificial
Indeed, as can be seen in figure 7-(a), it corre- mouth. Moreover, these observations on the am-
sponds to an oscillation on the first resonance plitude and frequency evolution patterns for aeo-
mode of the pipe, whereas regimes A1-C2 and lian regimes can recall the results of Meissner in
B correspond to oscillations on the second res- the case of a cavity resonator excited by an air
onance mode. Consequently, the oscillation fre- jet [24].
quency of regime A2-C1 being lower than that of However, the confrontation between figures 5
regimes A1-C2 and B, the value of θ is increased. and 8 highlights that the system under study
Moreover, for the two sounds which appear produces aeolian regimes with a larger amplitude
for low values of the mouth pressure, the bell- than those produced by a recorder. Such impor-
shaped evolution of the oscillating amplitude is tant differences between numerical and experi-
clearly visible, recalling the characteristics of mental results (in term of amplitudes) for low
aeolian regimes of the model. On the contrary, values of the dimensionless jet velocity θ have

7
also been observed by Fletcher [4] in the case of • in the third zone, where τ̃ < 0.1, the first
a flue organ pipe. It can probably be related register is unstable, and the second register
to the very simple modeling of the complex be- is the only stable solution of the system.
haviour of the air jet in flute-like instruments (see
for example [13, 16, 28, 20, 29]), which is simply The existence of a range of the dimensionless
represented, in the system under study, by an delay τ̃ (between 0.1 and 0.7) where two periodic
amplification factor and a delay. solutions are simultaneously stable implies the
existence of an hysteresis phenomenon. Indeed,
starting from a value of τ̃ where the first register
4 Jumps between periodic solu- is the only stable solution, and decreasing the
delay τ̃ (i.e. "blowing harder"), we observe that
tion branches the system follows the branch corresponding to
the first register until it becomes unstable (for
In this section, we study system (6) as in section
τ̃ = 0.1). At this point, the system synchronizes
3, with the addition of a second acoustic mode
with the second register.
(i.e. two terms in equation (4), m = 2). As
Starting from this new point, and increasing
for an open cylindrical resonator, the resonance
the delay τ̃ (i.e. "blowing softer"), the system
frequency of the second mode is chosen close to
follows the branch corresponding to the second
twice the first resonance frequency [12].
register. It is only when this second branch be-
comes unstable (for τ̃ = 0.7) that the system
4.1 Register change and hysteresis comes back to the first register.
phenomenon A time-domain simulation, using a Bogacki-
Shampine method based on a third-order Runge-
The use of a transfer function including two Kutta scheme [30] allows to confirm and high-
modes implies the existence of two standard light this hysteresis phenomenon. In order to
regimes. Each of these regimes results from the compare the different results, we superimpose, in
coupling between one of the two acoustic modes figure 10 (which is a zoom of figure 9-(a)), the bi-
and the first hydronamic mode of the jet (n = 0). furcation diagram and the oscillating amplitude
For a second resonance frequency slightly lower computed with this time-domain solver, for both
than twice the first one, numerical continuation an increasing and a decreasing ramp of the delay
of the corresponding periodic solution branches τ̃ .
leads to the bifurcation diagram shown in fig-
ure 9 - (a). As previously, it represents the os- 4.2 Register change: experimental il-
cillating amplitude of the dimensionless variable
lustration
v(t̃)/Uj as a function of the dimensionless delay
τ̃ . Here again, the stability is addressed by ex- In flute-like instruments, it is well-known by mu-
amining the Floquet multipliers. However, there sicians that blowing harder in the instrument
is a major difference with the case where only eventually leads to a "jump" to the note an oc-
one mode of the resonator is considered: indeed, tave above. Figure 11 represents the frequency
both stable and unstable portions appear (sym- evolution of the sound produced by a Zen-On
bol × indicates the stable parts of the branches). Bressan plastic alto recorder played by the arti-
More precisely, we can distinguish three differ- ficial mouth, during an increasing and a decreas-
ent zones in figure 9 - (a): ing ramp of the alimentation pressure.
These experimental results illustrate the
• the first one is defined for τ̃ > 0.7; in this register change phenomenon, corresponding to
case the first standard regime (called "first the "jump" from a periodic solution branch
register" in the context of musical acous- of frequency f1 to another of frequency f2 ≈ 2·f1 .
tics) is stable, whereas the second standard
regime ("second register") is unstable. An hysteresis is also observed. Recalling that
in flute-like instruments, the blowing pressure is
• for 0.1 < τ̃ < 0.7, the two standard regimes related to the convection delay τ of the hydro-
(i.e. the two registers) are simultaneously dynamic perturbations along the jet, these ex-
stable. perimental results can be compared to numerical

8
10

dimensionless frequency f/f (f =345 Hz)


static solutions Increasing pressure ramp
(a) standard regime (1st register, n = 0) f /f
2 1 Decreasing pressure ramp

dimensionless variable v/Uj


8 standard regime (2nd register, n = 0)
stable solutions 2
oscillation amplitude of

1
6

1
1,8

4
1,6

2
1,4

0
0 0.5 1 1.5 2 2.5 3 3.5 4 1,2
dimensionless delay τ ⋅ ω1 (ω1 = 2764)

10 f /f
static solutions 1 1
(b) standard regime (1st register, n = 0)
standard regime (2nd register, n = 0) 1 1.2 1.4 1.6 1.8 2
8
stable solutions dimensionless delay τ ⋅ ω (ω =2168)
dimensionless variable v/Uj

1 1
oscillation amplitude of

4 Figure 11: Dimensionless playing frequency f /f1


2 of a Zen-On Bressan plastic alto recorder blown
0
0 0.5 1 1.5 2 2.5 3 3.5 4
by an artificial mouth, during an increasing (solid
dimensionless delay τ ⋅ ω1 (ω1 = 2764)

10
static solutions
line) and decreasing (dashed line) blowing pres-
(c) standard regime (1st register, n = 0)
standard regime (2nd register, n = 0) sure ramp, showing jumps between the two stan-
j

8
dimensionless variable v/U

stable solutions
oscillation amplitude of

6
dard regimes. Two first resonance frequencies of
the resonator: f1 = 345 Hz, f2 = 733 Hz. F
4
fingering (third octave).
2

0
0 0.5 1 1.5 2 2.5 3 3.5 4
dimensionless delay τ ⋅ ω (ω = 2764)
1 1
results presented in figure 10. It shows again
Figure 9: Bifurcation diagrams of the system (6), that the system under study reproduces classical
obtained with numerical continuation, for differ- behaviour of recorder-like instruments.
ent values of inharmonicity of the two resonance
modes. Parameter values: m = 2 ; α = 340 ; 5 Influence of the ratio of the
a1 = 10 ; a2 = 30 ; Q1 = 100 ; Q2 = 100 ;
ω1 = 2764. (a) : ω2 = 5510, ωω12 = 1.99. (b) : two resonance frequencies
ω2 = 5528, ωω12 = 2. (c) : ω2 = 5654, ωω21 = 2.05.
Numerical continuation allows a systematic in-
Abscissa: dimensionless delay τ̃ = ω1 τ . Ordi-
vestigation of the parameters influence on the os-
nate: oscillation amplitude of the dimensionless
cillation regimes. In this section, we focus on the
variable v(t̃)/Uj . Symbols × indicate the stable
influence of the ratio of the two resonance fre-
parts of the branches.
quencies. Indeed, this parameter is well-known
15
to instrument makers and players for its influence
DDE Biftool − static solutions
DDE Biftool − standard regime (1st register, n = 0) on the sound characteristics [31]. Moreover, its
DDE Biftool − standard regime (2nd register, n=0)
j
dimensionless variable v/U

time−domain simulation: decreasing delay ramp influence has been proved for other musical in-
oscillation amplitude of

10 time−domain simulation: increasing delay ramp


stable solutions struments (for example, in [32]).

5
5.1 Stability of periodic solution
branches
0
0 0.2 0.4 0.6 0.8 1 1.2 1.4
dimensionless delay τ ⋅ ω (ω = 2764)
1 1
1.6 1.8 2
Figures 9 present bifurcation diagrams obtained
for different values of this ratio (respectively
ω2 ω2 ω2
Figure 10: Comparison between the bifurcation ω1 = 1.99 ; ω1 = 2 and ω1 = 2.05). Only the
diagram of the system (6), obtained with nu- second resonance frequency ω2 varies from one
merical continuation, and results provided by a case to the other; all the others parameters are
time-domain solver. Same parameter values as kept constant.
in figure 9-(a). Abscissa: dimensionless delay For all these bifurcation diagrams, we found
τ̃ = ω1 τ . Ordinate: oscillating amplitude of the an aeolian regime (see section 3.2) in a zone de-
dimensionless variable v(t̃)/Uj . Symbols × indi- fined by about 0.8 < τ̃ < 5.5 . For sake of read-
cate the stable parts of the branches. ability, the corresponding curves are not repre-
sented here. The dark gray curve corresponds to
the first register, the light gray dashed one cor-
responds to the second register, and symbols ×

9
indicate stable parts of the branches. stable, to a final value for which the first register
Although the resonance frequency of the sec- is stable. We note three different regimes: zones
ond mode is only slightly modified from one case A (τ̃ < 0.71) and C (τ̃ > 0.9) correspond
to the other, one observes significant changes respectively to the second and first registers.
in the stability properties of the two registers. In zone B (0.71 < τ̃ < 0.9), the presence of
In the case of two perfectly harmonic acoustic multiple frequencies reveals the existence of a
modes (ω2 = 2ω1 ), represented in figure 9 - (b), quasiperiodic regime, which agrees with the
the first register is initially stable, and becomes results of Floquet analysis (presented in figures
unstable when the delay τ̃ decreases (i.e. when 9 - (b) and 12).
the blowing pressure increases). On the contrary,
the second register is first unstable and becomes
stable when the delay τ̃ decreases. For the range 15
A B C

50
of τ̃ located between the loss of stability of the

1
dimensionless frequency f/f
first register and the stabilization of the second 10 0

register (i.e. between the two vertical dot-dashed

(f = 440 Hz)
−50
lines in figure 9 - (b)), there is no stable solution,
5

1
neither static, nor periodic. −100

In order to determine the nature of the differ-


0 −150
ent resulting bifurcations, we propose to study 0.6 0.65 0.7 0.75 0.8 0.85
dimensionless delay τ.ω (ω = 2764)
1 1
0.9 0.95

Floquet multipliers in the vicinity of stability


changes. As shown in figure 12, stabilization of Figure 13: Time-frequency representation ob-
the second register occurs, at τ̃ = 0.68, by the tained by a time-domain simulation of the system
crossing of the unit circle by two conjugate com- (6), using a third-order Runge-Kutta scheme. A
plex Floquet multipliers. slowly increasing ramp of the dimensionless delay
τ̃ = ω1 τ is achieved. Zone A corresponds to the
0.06
Unit circle
Floquet multipliers
second register, zone C corresponds to the first
0.04 register, and zone B to a quasiperiodic regime.
0.02
Parameter values are the same as in figure 9 -
0
(b). The dark dot-dashed lines indicate the two
ℑ(µ)

−0.02

−0.04
first resonance frequencies of the transfer func-
−0.06 tion (f1 = 440 Hz and f2 = 880 Hz).
−0.08

−1.015 −1.01 −1.005 −1


ℜ(µ)
−0.995 −0.99 −0.985
As shown in figure 9 - (a), a small decrease
of the second mode resonance frequency (of
Figure 12: Representation in the complex plane about 0.6 %) compared to the harmonic case
of the Floquet multipliers for a point located just ω2 = 2ω1 does not alter the general form of the
before the stabilization of the second periodic so- bifurcation diagram: the first register is stable
lution branch (related to the second register), at for high values of τ̃ , whereas the second register
τ̃ = 0.68. Parameter values are the same as in is first unstable and becomes stable when the
figure 9 - (b). delay τ̃ decreases. However, the stability ranges
are significantly modified. Compared to the
In the case of a direct bifurcation, this Hopf previous case, the first register is stable for
bifurcation of the limit cycle leads to a quasiperi- smaller values of the delay τ̃ , which leads to the
odic regime [23]. The numerical tools used here existence of a range of τ̃ where the two registers
do not allow numerical continuation of this kind are simultaneously stable. In the later case, the
of solutions. On the other hand, time-domain hysteresis phenomenon emphasized in section 4
integration methods allow to study such regimes. becomes possible.
Figure 13 shows the time-frequency analysis of
the signal obtained with the same time-domain These stability ranges are affected again by an
solver as previously, by increasing the dimen- increase of the second mode resonance frequency
sionless delay τ̃ in a quasi-static way from an (leading to a difference of about 2% between
initial value for which the second register is ω2 and 2ω1 ). Figure 9 - (c) shows the resulting

10
bifurcation diagram. One can notice that the For a larger negative inharmonicity of the
first register is now stable for all values of τ̃ resonance frequencies ( ωω21 ≈ 1.92), the same
between 2 and 0. Hence, once the system is experiment leads to a very different behaviour.
synchronized on this branch of solutions, no As shown in figure 15, we observe a transition
register change is possible, and the only way to from the first register to a quasiperiodic regime
make the system oscillate on the second register (in zone D) without hysteresis effect.
is to change initial conditions. Time-domain
simulations results (not presented here) show 80

1
D

mouth pres− dimensionless frequency f/f


good agreement with these characteristics. 6
70

60

(f = 458.7 Hz)
50
(a) 4
40

1
2 30
5.2 Experimental illustration 20

0 10

As emphasized in section 1, the simplicity of the

sure (Pa)
(b) 500

studied system prevents a quantitative compari- 0


0 20 40 60 80 100 120
son between numerical results and experimental time (s)

datas. However, we can qualitatively compare


some experimental phenomena with numerical Figure 15: Time-frequency representation (a) of
results presented in the previous subsection. the sound of a Yamaha (YRA-28BIII) plastic alto
recorder played by an artificial mouth during an
The use of an artificial mouth with a real in- increasing and decreasing blowing pressure ramp
strument highlights the influence of the inhar- (b). The white dot-dashed lines indicate the two
monicity on the sound characteristics. Indeed, first resonance frequencies (f1 ≈ 458.7 Hz and
the measured inharmonicity depends on the fin- f2 ≈ 883.3 Hz) of the resonator. B-flat fingering
gering. (3th octave), with a significant negative inhar-
Thus, for a small positive inharmonicity ( ωω21 ≈ monicity ( ωω21 ≈ 1.92).
2.04), an increasing mouth pressure ramp (corre-
sponding to a decrease of the convection delay τ ) Comparison of these results with those of fig-
causes a regime change from the first register to ures 9 - (a), 9 - (b) and 9 - (c) proves that the
the second register, including an hysteresis effect behaviour of both the studied system and the
(as shown in figure 14). real instrument strongly depends on the inhar-
monicity of the two first resonance frequencies.
80
A small change of this parameter value is
1
dimensionless frequency f/f

4 60 enough to alter the oscillation thresholds of the


(f = 387.3 Hz)

3
40
different regimes, their stability properties, or
(a)
2 even the nature of the oscillation regimes. De-
1

1
20
pending on the case, the system can "jump" be-
0
t t
0 tween branches of periodic solutions (with hys-
mouth pres−

1 2

teresis effect), or jump from a periodic solution


sure (Pa)

500
(b)
250
0
0 20 40 60 80 100 120
branch to a quasiperiodic regime.
time (s)
Such quasiperiodic sounds were previously ob-
served experimentally on flute-like instruments
Figure 14: Time-frequency representation (a) of and obtained through numerical simulations, for
the sound of a Yamaha (YRA-28BIII) plastic example by Fletcher [4] and Coltman [3]. Basing
alto recorder played by an artificial mouth during on the comparison between passive resonance
an increasing and decreasing blowing pressure frequencies of the pipe and sounding frequencies
ramp (b). The white dot-dashed lines indicate of the instrument, Coltman [3] explained this
the two first resonance frequencies (f1 ≈ 387.3 phenomenon by a beat between two frequen-
Hz, f2 ≈ 793.3 Hz) of the resonator. The dark cies: a first one corresponding to an harmonic
dashed lines indicate register changes. G finger- (exact multiple of the fundamental frequency),
ing (3th octave), with a slightly positive inhar- and a second corresponding to an oscillation
monicity ( ωω12 ≈ 2.04). on an higher resonance mode of the pipe.

11
Our numerical and experimental observations, However, it would be hazardous to use the
highlighting that a slight shift of the second studied dynamical system as a predictive tool,
resonance frequency is enough to make the for example for the design of musical instru-
system (respectively the instrument) produce ments. Indeed, some parameters of the sys-
quasiperiodic regimes, show good agreement tem can not be related to mesurable physical
with such observations. quantities. Moreover, as we emphasized in sec-
tion 1, some important physical phenomena have
Moreover, these results seem to be consistent not been taken into account. These two ele-
with the experience of a recorder maker [31]: ac- ments make vain any attempt of quantitative
cording to him, the first register is stable for a comparison between numerical and experimen-
wider range of alimentation pressure in the case tal results. Particularly, we did not discussed
of strong inharmonicity and the case of a perfect about the sound timbre, which corresponds to
harmonicity should be avoided to prevent insta- the spectrum of the sound produced by the in-
bilities. It is interesting to note that this feature strument. This characteristics is essential in a
is an important difference between recorders and musical context since it allows us to distinguish,
transverse flutes. Indeed, unlike recorder mak- in the case of a steady-state regime, the sound of
ers, flute makers seek a good harmonicity to al- a flute from that, for example, of a trumpet. The
low flutists to play in tune on different octaves system studied here can not be used to predict
using a same fingering. the sound timbre, since some elements known for
their significant influence on the spectrum (but
also on the amplitude of the different regimes)
6 Conclusion have been neglected:
We have studied a nonlinear delay dynamical • the offset between the position of the labium
system with small number of degrees of freedom. and the channel exit. Due to symmetry
We focused on periodic solutions. Indeed, this properties of the nonlinear function, this pa-
system is inspired by flute-like instruments rameter controls the ratio between even and
and in most cases, notes produced by these odd harmonics [33, 3, 2].
instruments are periodic oscillations.
• nonlinear losses due to flow separation and
Because of the drastic simplifications, the vortex shedding at the labium, which have
studied system can not be considered a priori an important influence on the sound level
as a model for these instruments. However, [34, 35, 20, 2, 6].
as shown in this paper, the use of numerical
continuation tools, providing a more global • the dipolar character of the pressure source,
vision of the dynamics of the system, highlights created by the oscillation of the jet around
that the system not only presents a consid- the labium [15, 13, 36].
erable variety of periodic regimes, but also • the jet velocity profile [13, 37], and an accu-
has a second interest. Indeed, it qualitatively rate modeling of the behaviour of the unsta-
reproduces phenomena observed experimentally ble jet ([17, 16] for example), particularly for
on flute-like instruments such as amplitude and low values of the dimensionless jet velocity
frequency evolutions for both standard and θ [35, 13].
aeolian regimes, regime changes with hysteresis
effect, and quasiperiodic oscillations. Taking into account these two first elements
do not compromise the use of the approach pre-
We can furthermore investigate the influence sented in this paper. Only the number of param-
of some parameters related to instrument eters and degree of freedom (and thus the com-
maker’s issues. We focused here on the role putation time) increases. However, taking into
of the inharmonicity of the two first resonance account the third element involves the presence
frequencies. Analysis of bifurcation diagrams of a delayed derivative term in the right-hand
leads to results that are consistent with the side of the second equation of system (1). This
empirical knowledge of an instrument maker. change transforms the delay system into a neu-
tral delay dynamical system, and thus introduces

12
additional difficulties in numerical continuation A DDE Biftool: Reformulation
and system analysis [38]. of the studied system
To be implemented in the software DDE Biftool,
7 Acknowledgements
system (1) needs to be reformulated as follows:
The authors would like to thank Philippe Bolton
for interesting and helpful discussions during the ẋ(t) = f (x(t), x(t − τ ), γ) (9)
completion of this work. where τ is a delay, x the vector of state variables,
and γ the set of parameters.

We detail here such a reformulation in the case


of a single-mode transfer function (equation (4)):
jω · a1
Y (ω) = ω1 (10)
ω12 − ω 2 + jω Q 1

where ω is the pulsation, and a1 , ω1 and Q1 are


respectively the modal amplitude, the resonance
pulsation and the quality factor of the resonance
mode.

Let us inject equation (10) in the third equation


of system (1):

 
ω1
V (ω) · ω12 2
− ω + jω = jω · a1 · P (ω). (11)
Q1
An inverse Fourier transform leads to:
ω1
v̈(t) + ω12 · v(t) + · v̇(t) = a1 · ṗ(t). (12)
Q1
The right-hand side of this expression can be de-
veloped using the second equation of system (1):

p(t) = α · tanh [z(t)] . (13)

Differentiating with respect to time, and using


the first equation of system (1), we obtain:

ṗ(t) = α · ż(t) · 1 − tanh2 [z(t)]



(14)
= α · v̇(t − τ ) · 1 − tanh2 [v(t − τ )]


Injecting this expression in equation (12) leads


to:
ω1
v̈(t) + v̇(t) + ω12 v(t) = a1 α · v̇(t − τ )
Q1 (15)
1 − tanh2 [v(t − τ )]


To improve numerical conditioning of the prob-


lem, it is convenient to make the temporal vari-
ables dimensionless. Let us introduce the new
variables: 
t̃ = ω1 · t
τ̃ = ω1 · τ

13
Equation (15) becomes: B Linear analysis around the
equilibrium solution
d2 v(t̃) ω1 d(v(t̃))
+ · + ω12 · v(t̃) B.1 Eigenvalues of the Jacobian
d( ωt̃1 )2 Q1 d( t̃ )
ω1
(16) Linearisation of system (6) around equilibrium
d[v(t̃ − τ̃ )] 
· 1 − tanh2 [v(t̃ − τ̃ )] .

= a1 α solution (7) allows to compute the eigenvalues

d( ω1 ) λ of the Jacobian matrix of the studied system
along the branch of equilibrium solutions. Their
It leads to:
real parts Re(λ) determine the stability proper-
ties of the static solution: it is stable if and only
1 a1 α if all the eigenvalues have negative real parts [23].
v̈(t̃) + · v̇(t̃) + v(t̃) = · v̇(t̃ − τ̃ )
Q1 ω1 (17) Considering, for sake of clarity, a single acous-
1 − tanh2 v(t̃ − τ̃ ) , tic mode of the resonator (i.e. m = 1 in equa-
  
tion (4)), we represent, in figure 16, Re(λ) with
where v̇ now define the derivative of v with re- respect to the continuation parameter τ̃ , along
spect to dimensionless time t̃. the branch of equilibrium solutions. This rep-
We define a new variable: resentation highlights that this static solution is
stable for 1.8 < τ̃ < 4.1 and for 9.1 < τ̃ < 9.5.
y(t̃) = v̇(t̃), (18) Analysis of both real and imaginary parts of the
eigenvalues shows that the equilibrium solution
which leads to the correct form of the system
is a focus point when it is stable, and a saddle
(given by equation (9)):
point when it is unstable [23]. Moreover, inspec-

 v̇(t) =y(t) tion of the intersection points with the x-axis (de-

 a1 α fined by Re(λ) = 0) determines Hopf bifurcations
· y(t̃ − τ̃ ) · 1 − tanh2 v(t̃ − τ̃ )
   
ẏ(t̃) =

ω1 [23], which correspond to the birth of the differ-

 1 ent periodic solution branches (here highlighted
− v(t̃) − y(t̃).

by circles at τ̃ = 1.8 ; τ̃ = 4.1 ; τ̃ = 9.1 and

Q1

(19) τ̃ = 9.5).
0.2

0.1
Generalization to a system including m res-
0
onance modes is straightforward. In this case,
Re(λ)

equation (4) contains m terms (with m > 1): −0.1

−0.2
m
X jω · ak
Y = 2 ωk (20) −0.3

ωk − ω 2 + jω Q −0.4
k=1 k
0 2 4 6 8 10 12 14 16
Dimensionless delay τ ⋅ ω (ω = 2260)
1 1
and it leads to a system of 2m equations (i.e. 2
equations by resonance mode), of the following
Figure 16: Real part of the eigenvalues λ of the
form:
Jacobian matrix of system (6) linearised around
static solution (7), with respect to the dimen-
sionless delay τ̃ = τ · ω1 . Hopf bifurcations (in-

 ∀i =[1, 2..., m] :
tersections with the horizontal line Re(λ)= 0) are



 v̇i (t̃) =yi (t̃)
highlighted by circles. Parameters value: m = 1,


 ( "m #)
ai α


α = 10 ; a1 = 70 ; ω1 = 2260 ; Q1 = 50.
X
1 − tanh2

 ẏi (t̃) = vk (t̃ − τ̃ )


ω1
k=1
m  2
ωi

 X
  

 yk (t̃ − τ̃ ) − vi (t̃) B.2 Open-loop gain
ω

 1
k=1


ωi In a feedback loop system such as the one pre-




 − yi (t̃). sented in section 2, another approach to get infor-
ω 1 Qi
(21) mation about the destabilization mechanisms of

14
the equilibrium solution is to study the open-loop 8 G=1 (instability condition)
modulus G (Uj = 6.5 m/s and Uj = 15.7 m/s)

modulus G
6 birth of oscillations (n =0)
gain Gol (ω) of the linearised system. Indeed, the (a)
4
birth of oscillations (n =1)

emergence of auto-oscillations from the equilib- 2


0
rium solution (corresponding to a Hopf bifurca- 200 400 600 800
frequency (Hz)
1000 1200 1400

phase P (radians)
tion) is possible under the two following condi- 0

tions [39]: (b) −10


phase = n ⋅ 2π (instability condition)
phase (Uj = 6.5 m/s)
−20
phase (Uj = 15.7 m/s)

• modulus G of the linearised open-loop gain 200 400 600 800


frequency (Hz)
1000 1200 1400

Gol must be equal to 1.


Figure 17: Open-loop gain modulus G (a) and
• phase P of the linearised open-loop gain Gol
phase P (b) defined by eq. (24), for two different
must be a multiple of 2π.
values of the jet velocity Uj . Emergence of self-
sustained oscillations is possible if the modulus is
Writting system (1) in the frequency domain
equal to one (dash line, in (a)), and if the phase
leads to:
crosses a straight line with equation P = −n · 2π

 Z(ω) = V (ω)e−jωτ (dash lines, in (b)).
P (ω) = α · tanh(Z(ω)) (22)
V (ω) = Y (ω) · P (ω)

of n, an instability may emerge from the equilib-
and linearisation of the second equation around rium branch.
the equilibrium solution (7) leads to the open- From the physics point of view, the integer n
loop gain: represents the rank of the hydrodynamic mode
of the jet involved in the emergence of the insta-
bility.
Gol (ω) = αY (ω) · e−jωτ , (23)
The conditions of emergence of auto-oscillations
are then given by:

G(ω) = α · |Y (ω)| = 1
P (ω) = arg(Y (ω)) − ωτ = −n · 2π
(24)
where n is an integer.

To exemplify the conditions given by equation


(24), let us consider a transfer function Y (ω)
representing the first two modes of a cylindri-
cal resonator (m = 2 in equation (4)). Figure
17 shows the variables G (a) and P (b) with
respect to ω. Two different values of the de-
lay τ are considered (corresponding to two dif-
ferent values of the jet velocity Uj = 6.5ms−1
and Uj = 15.7ms−1 ) for phase P (G is indepen-
dent of τ ). The points where equations (24) are
fulfilled are marked with circles (◦) when n = 0
and squares () when n = 1. Plots of P for lower
values of Uj (i.e. for higher values of τ ) would
have revealed other intersections with horizontal
lines P = −n · 2π, with larger values of n.
Provided the amplification α is large enough,
this example highlights the existence of an infin-
ity of solutions of equations (24) for different val-
ues of Uj , each solution being related to a given
value of the integer n. Therefore, for each value

15
References [11] K. Engelborghs, “DDE Biftool: a Matlab
package for bifurcation analysis of delay dif-
[1] M. E. McIntyre, R. T. Schumacher, and ferential equations,” tech. rep., Katholieke
J. Woodhouse, “On the oscillations of musi- Universiteit Leuven, 2000.
cal instruments,” Journal of the Acoustical
Society of America, vol. 74, no. 5, pp. 1325– [12] A. Chaigne and J. Kergomard, Acoustique
1345, 1983. des instruments de musique (Acoustics of
musical instruments), ch. 10, pp. 469–510.
[2] M. P. Verge, B. Fabre, A. Hirschberg, and Belin (Echelles), 2008.
A. P. J. Wijnands, “Sound production in
recorderlike instruments. I. Dimensionless [13] B. Fabre and A. Hirschberg, “Physical mod-
amplitude of the internal acoustic field,” eling of flue instruments: A review of
Journal of the Acoustical Society of Amer- lumped models,” Acta Acustica united with
ica, vol. 101, no. 5, pp. 2914–2924, 1997. Acustica, vol. 86, no. 4, pp. 599–610, 2000.
[3] J. W. Coltman, “Jet offset, harmonic con- [14] H. von Helmholtz, On the sensation of
tent, and warble in the flute,” Journal of tones. Dover New York, 1954.
the Acoustical Society of America, vol. 120,
no. 4, pp. 2312–2319, 2006. [15] J. W. S. Rayleigh, The theory of sound sec-
ond edition. New York, Dover, 1894.
[4] N. H. Fletcher, “Sound production by organ
flue pipes,” Journal of the Acoustical Soci- [16] J. W. Coltman, “Jet behavior in the flute,”
ety of America, vol. 60, no. 4, pp. 926–936, Journal of the Acoustical Society of Amer-
1976. ica, vol. 92, no. 1, pp. 74–83, 1992.
[5] R. T. Schumacher, “Self-sustained oscilla- [17] A. Nolle, “Sinuous instability of a planar jet:
tions of organ flue pipes: an integral equa- propagation parameters and acoustic exci-
tion solution,” Acustica, vol. 39, pp. 225– tation,” Journal of the Acoustical Society
238, 1978. of America, vol. 103, no. 6, pp. 3690–3705,
1998.
[6] R. Auvray, B. Fabre, and P. Y. Lagrée,
“Regime change and oscillation thresholds [18] P. de la Cuadra, C. Vergez, and B. Fabre,
in recorder-like instruments,” Journal of “Visualization and analysis of jet oscilla-
the Acoustical Society of America, vol. 131, tion under transverse acoustic perturba-
no. 4, pp. 1574–1585, 2012. tion,” Journal of Flow Visualization and Im-
[7] J. W. Coltman, “Time-domain simulation of age Processing, vol. 14, no. 4, pp. 355–374,
the flute,” Journal of the Acoustical Society 2007.
of America, vol. 92, no. 1, pp. 69–73, 1992. [19] F. Silva, V. Debut, J. Kergomard,
[8] S. Karkar, C. Vergez, and B. Cochelin, “Os- C. Vergez, A. Deblevid, and P. Guillemain,
cillation threshold of a clarinet model: a nu- “Simulation of single reed instruments
merical continuation approach,” Journal of oscillations based on modal decomposition
the Acoustical Society of America, vol. 131, of bore and reed dynamics,” in Proceedings
no. 1, pp. 698–707, 2012. of the International Congress of Acoustics,
(Madrid, Spain), Sept. 2007.
[9] E. J. Doedel, “AUTO: A program for au-
tomatic bifurcation analysis of autonomous [20] B. Fabre, A. Hirschberg, and A. P. J. Wi-
systems,” Congressus Numerantium, vol. 30, jnands, “Vortex shedding in steady oscilla-
pp. 265–284, 1981. tion of a flue organ pipe,” Acta Acustica
united with Acustica, vol. 82, no. 6, pp. 863–
[10] B. Cochelin and C. Vergez, “A high or- 877, 1996.
der purely frequency-based harmonic bal-
ance formulation for continuation of peri- [21] K. Engelborghs, T. Luzyanina, and
odic solutions,” Journal of Sound and Vi- D. Roose, “Numerical bifurcation analy-
bration, vol. 324, no. 1-2, pp. 243–262, 2009. sis of delay differential equations using

16
DDE-Biftool,” ACM Transactions on Math- [30] P. Bogacki and L. F. Shampine, “A 3(2) pair
ematical Software, vol. 28, no. 1, pp. 1–21, of Runge-Kutta formulas,” Applied Mathe-
2002. matics Letter, vol. 2, no. 4, pp. 321–325,
1989.
[22] K. Engelborghs, T. Luzyanina, K. J.
In’t Hout, and D. Roose, “Collocation meth- [31] P. Bolton, “recorder maker.” personal com-
ods for the computation of periodic solu- munication.
tions of delay differential equations,” SIAM
Journal on Scientific Computing, vol. 22, [32] J. P. Dalmont, B. Gazengel, J. Gilbert, and
no. 5, pp. 1593–1609, 2000. J. Kergomard, “Some aspects of tuning and
clean intonation in reed instruments,” Ap-
[23] A. H. Nayfeh and B. Balachandran, Applied plied Acoustics, vol. 46, no. 1, pp. 19–60,
Nonlinear Dynamics. Wiley, 1995. 1995.
[24] M. Meissner, “Aerodynamically excited [33] N. H. Fletcher and L. M. Douglas, “Har-
acoustic oscillations in cavity resonator ex- monic generation in pipes, recorders, and
posed to an air jet,” Acta Acustica united flutes,” Journal of the Acoustical Society of
with Acustica, vol. 88, no. 2, pp. 170–180, America, vol. 68, no. 3, pp. 767–771, 1980.
2001.
[34] M. S. Howe, “Contributions to the theory
[25] F. Blanc, Production de son par couplage
of aerodynamic sound, with application to
écoulement/résonateur acoustique : étude
excess jet noise and the theory of the flute,”
des paramètres de facture des flûtes par ex-
Journal of Fluid Mechanics, vol. 71, no. 4,
périmentations et simulations numériques
pp. 625–673, 1975.
d’écoulements (Sound production by the in-
teraction between a jet flow and a resonator: [35] A. Hirschberg, J. Kergomard, and G. Wein-
study of geometrical parameters of flutes reich, Mechanics of musical instruments,
through experimentation and numerical flow ch. 7, pp. 291–361. Springer-Verlag, 1995.
simulations). PhD thesis, Université Pierre
et Marie Curie - Paris 6, 2009. [36] J. W. Coltman, “Jet drive mechanisms in
edge tones and organ pipes,” Journal of the
[26] D. Ferrand, C. Vergez, B. Fabre, and Acoustical Society of America, vol. 60, no. 3,
F. Blanc, “High-precision regulation of a pp. 725–733, 1976.
pressure controlled artificial mouth: the
case of recorder-like instruments,” Acta [37] M. P. Verge, R. Caussé, B. Fabre,
Acustica united with Acustica, vol. 96, no. 4, A. Hirschberg, A. P. J. Wijnands, and
pp. 701–712, 2010. A. van Steenbergen, “Jet oscillations and
jet drive in recorder-like instruments,”
[27] D. H. Lyons, “Resonance frequencies of Acta Acustica united with Acustica, vol. 2,
the recorder (english flute),” Journal of the pp. 403–419, 1994.
Acoustical Society of America, vol. 70, no. 5,
pp. 1239–1247, 1981. [38] D. Barton, B. Krauskopf, and R. E. Wil-
son, “Collocation schemes for periodic solu-
[28] M. P. Verge, B. Fabre, W. E. A. Mahu, tions of neutral delay differential equations,”
A. Hirschberg, R. R. van Hassel, A. P. J. Journal of Difference Equations and Appli-
Wijnands, J. J. de Vries, and C. J. Hogen-
cations, vol. 12, no. 11, pp. 1087–1101, 2006.
doorm, “Jet formation and jet velocity fluc-
tuations in a flue organ pipe,” Journal of [39] N. H. Fletcher, “Autonomous vibration
the Acoustical Society of America, vol. 95, of simple pressure-controlled valves in gas
pp. 1119–1132, 1994. flows,” Journal of the Acoustical Society of
America, vol. 93, no. 4, pp. 2172–2180, 1993.
[29] M. S. Howe, “The role of displacement thick-
ness fluctuations in hydroacoustics, and the
jet drive mechanism of the flue organ pipe,”
Proceedings of the Royal Society of London,
vol. A374, pp. 543–568, 1981.

17

You might also like