Sound Simulation-Based Design Optimization of Brass Wind

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 11

Sound simulation-based design optimization of brass wind instruments

Robin Tournemenne, Jean-François Petiot, Bastien Talgorn, Joël Gilbert, and Michael Kokkolaras

Citation: The Journal of the Acoustical Society of America 145, 3795 (2019); doi: 10.1121/1.5111346
View online: https://doi.org/10.1121/1.5111346
View Table of Contents: https://asa.scitation.org/toc/jas/145/6
Published by the Acoustical Society of America

ARTICLES YOU MAY BE INTERESTED IN

The effect of the cutoff frequency on the sound production of a clarinet-like instrument
The Journal of the Acoustical Society of America 145, 3784 (2019); https://doi.org/10.1121/1.5111855

Analysis on scattering characteristics of Love-type wave due to surface irregularity in a piezoelectric structure
The Journal of the Acoustical Society of America 145, 3756 (2019); https://doi.org/10.1121/1.5102165

An experimental evaluation of the generalized sinusoidal frequency modulated waveform for active sonar
systems
The Journal of the Acoustical Society of America 145, 3741 (2019); https://doi.org/10.1121/1.5113581

Infrasonic source altitude localization based on an infrasound ray tracing propagation model
The Journal of the Acoustical Society of America 145, 3805 (2019); https://doi.org/10.1121/1.5110712

Discriminative feature analysis based on the crossing level for leakage classification in water pipelines
The Journal of the Acoustical Society of America 145, EL611 (2019); https://doi.org/10.1121/1.5113809

Sound radiation and transonic boundaries of a plate with an acoustic black hole
The Journal of the Acoustical Society of America 145, 164 (2019); https://doi.org/10.1121/1.5081680
Sound simulation-based design optimization of brass wind
instruments
Robin Tournemenne,1,a) Jean-François Petiot,2 Bastien Talgorn,3 Joe
€l Gilbert,4 and
3
Michael Kokkolaras
1
Magique 3D Team, Inria Bordeaux Sud Ouest, 200 avenue de la vieille tour, 33405 Talence Cedex, France
2
Ecole Centrale de Nantes, LS2N, UMR CNRS 6004, 1 rue de la No€e, 44321 Nantes Cedex 3, France
3
Department of Mechanical Engineering, McGill University, Montr
eal, QC H3A 0C3, Canada
4
Laboratoire d’Acoustique de l’Universit
e du Mans, UMR CNRS 6613, avenue Olivier Messiaen,
72085 Le Mans Cedex 09, France

(Received 8 January 2019; revised 21 May 2019; accepted 23 May 2019; published online 28 June
2019)
A method for optimizing the inner shape of brass instruments using sound simulations is presented.
This study considers different objective functions and constraints (representative of both the intona-
tion and the spectrum of the instrument) for a relatively large number of design variables. A complete
physics-based model, taking into account the instrument and the musician’s embouchure, is used to
simulate steady regimes of sounds by means of the harmonic balance technique, the instrument being
represented by its input impedance. The design optimization variables are related to the geometrical
dimensions of the resonator. The embouchure’s parameters are varied during the optimization proce-
dure to obtain an average behavior of the instrument. The objective and constraint functions of the
optimization problem are evaluated using the physics-based simulation model, which is computation-
ally expensive. Moreover, the gradients of the objective and constraint functions can be discontinu-
ous, unavailable, or hard to approximate reliably. Therefore, a surrogate-assisted derivative-free
optimization strategy using the mesh adaptive direct search algorithm was employed. One example
of a B[ trumpet’s bore is used to demonstrate the effectiveness of the design optimization approach:
the obtained results improve previously reported objective function values significantly.
C 2019 Acoustical Society of America. https://doi.org/10.1121/1.5111346
V

[JW] Pages: 3795–3804

I. INTRODUCTION determines an instrument’s timbre and playability, notably


due to the height and bandwidth of peaks.
The development of innovative and higher-quality
With this in view, several researchers have used input
designs is crucial to the viability of musical instrument manu-
impedance to design an instrument’s inner shape using an
facturers. Many prototypes are required to ensure that quality
optimization approach. Following Kausel’s successful recon-
attributes such as intonation, ease of playing, timbre, projec-
struction of a trumpet bore using the Rosenbrock algorithm
tion, etc., are adequate at each stage of the development pro-
(Kausel, 2001), Noreland et al. (2010) optimized the instru-
cess. Numerical acoustic models may be helpful in shortening
ment’s intonation with a hybrid scheme for the input imped-
development cycles by minimizing the manufacturing of
ance model and shape constraints. Braden et al. (2009) also
costly prototypes. For example, in the case of brass instru-
optimized the intonation and the input impedance peak
ments, several studies propose the use of numerical modeling
heights of a trombone using a multi-modal input impedance
to predict quality and guide the design process (Campbell,
model, while Macaluso and Dalmont (2011) optimized and
2004; Macaluso and Dalmont, 2011).
built a near-perfect harmonic trumpet. Some studies investi-
The dominant physical quantity impacting the sound qual-
gated the relationship between input impedance features and
ity of a brass instrument is its input impedance (cf. Fig. 1).
psycho-acoustic criteria: Poirson et al. (2007) optimized the
Input impedance is the frequency-dependent quotient of pres-
trumpet using objective functions based on the input imped-
sure and volume flow at the instrument entry plane, and, at first
ance and targets defined by trumpet players preferences.
approximation, is the result of the instrument’s interior shape
(the bore). Many works have focused on modeling input Guilloteau (2015) looked for empirical relations between
impedance, e.g., Causse et al. (1984). playing frequencies and resonance frequencies to optimize
The impact of input impedance on sound quality is clarinets.
acknowledged in Campbell (2004), where, in a first approxi- However valuable, these works focused exclusively on
mation attempt, playing frequencies are governed mainly by instrument performance, neglecting a crucial element in
corresponding impedance peaks (Eveno et al., 2014). Beyond sound production: the musician embouchure. In particular,
influencing the instrument intonation, impedance also the studies of Eveno et al. (2014) showed that the relation-
ship between the resonance frequencies of the impedance
and the actual frequencies of the sounds played by musicians
can vary significantly. Although the impedance of an instru-
a)
Electronic mail: robin.tournemenne@inria.fr ment provides interesting information about sound quality,

J. Acoust. Soc. Am. 145 (6), June 2019 0001-4966/2019/145(6)/3795/10/$30.00 C 2019 Acoustical Society of America
V 3795
FIG. 1. Simulated input impedance Z
of a B[ trumpet (magnitude), highlight-
ing resonances 2, 3, 4, and 5 of the
instrument.

prediction of “playability” and sound qualities of brasses 1995), it is able to mimic a large range of playing phenom-
based solely on impedance remains difficult. ena [see, for example, the pioneering work of Elliot and
A second approach is based on a holistic model of the Bowsher (1982) or the more recent works of Petiot and
physical phenomenon, coupling the instrument and the musi- Gilbert (2013) and Velut et al. (2017)]. Similar to Chen and
cian embouchure, to produce sound simulations representa- Weinreich (1996), we argue that while the lips may not be
tive of the instrument quality. Using this approach, the entirely modeled by a 1-DOF model, most characteristic
authors integrate sound simulations in the optimization pro- behaviors of brasses can be reproduced by an outward strik-
cess. These simulations are obtained from a physics-based ing reed.
model to account for the interaction of the instrument with a Our physics-based model of the trumpet is based on
virtual musician embouchure (Tournemenne et al., 2017). In Eqs. (1), (2), and (3), which all depend on three periodic var-
a previous study, the instrument’s intonation was optimized iables: the opening height h(t) of the two lips, the volume
based on simulated playing frequencies (Tournemenne et al., flow u(t) of the air jet through the lip channel, and the pres-
2017). Two examples were considered, optimizing 2 or 5 of sure p(t) in the mouthpiece (cf. Fig. 2).
the bore’s geometrical parameters; results were quite
encouraging. p^ðjxÞ ¼ ZðjxÞ^
u ðjxÞ; (1)
The main objective of the present paper is to extend this d2 hðtÞ 2pf‘ dhðtÞ Pm  pðtÞ
new optimization paradigm in order to assess both its poten- þ þ ð2pf‘ Þ2 ½hðtÞ  h0  ¼ ;
dt2 Q‘ dt l‘
tial and limitation. The two main novelties of this paper are
the optimization of criteria based on the instrument sounds (2)
spectra and the inclusion of constraints in the problem for- sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
2jPm  pðtÞj
mulation. Three other contributions are noticeable: (i) an uðtÞ ¼ bhþ ðtÞ sign½Pm  pðtÞ : (3)
improved version of the optimization method has been q
implemented, (ii) a new solver is introduced, and (iii) the
performance limits of the optimization method are tested by These three equations express the acoustic impedance of
considering ten geometrical variables of the bore. A trumpet the resonator, a simple harmonic oscillator for the lips
is used as a representative brass instrument to demonstrate model, and the coupling between the lips and the trumpet,
the proposed design optimization method. respectively. Equation (3) shows two non-linearities: the
The paper is organized as follows. We first present square root originating from the Bernoulli equation, and the
extensive details on the physics-based model and the simula- positive part of the lip aperture hþ ¼ max(h, 0) modeling the
tion technique. We then formulate the optimization problems closed lips. Several parameters are included in this model:
and describe the principles of the mesh adaptive direct air density q, input impedance Z of the trumpet, the parame-
search (MADS) algorithm and the framework for surrogate- ters concerning musician embouchure which are Pm (the
assisted optimization. Finally, we conduct a case study con-
cerning the shape optimization of a trumpet with ten design
variables and draw conclusions.

II. TRUMPET MODELING


In this study, we utilize an elementary model of a brass
instrument under playing conditions: the vibrating lips are
modeled as a one-degree-of-freedom (1-DOF) outward-
striking valve, non-linearly coupled to the air column of the
FIG. 2. (Color online) Representation of the outward striking model of the
brass instrument. This elementary model is a good compro-
lips, with the definition of the variables of the physical model: Pm (pressure
mise between simplicity and efficiency. While the 1-DOF in the mouth), h (lip aperture), u (volume flow), and p (pressure in the
model cannot model real musician lips exactly (Yoshikawa, mouthpiece).

3796 J. Acoust. Soc. Am. 145 (6), June 2019 Tournemenne et al.
static overpressure in the mouth), f‘ (the resonance frequency (defined by its control parameters). The harmonic balance
of the lips), l‘ (the area density of the lips), b (the width of technique considers the sound’s permanent regime (steady
the lips), h0 (the rest value of the opening height of the lips), state); since the latter is periodic, the truncated pressure is
and Q‘ (the quality factor of the resonance of the lips). Input given by
impedance Z is computed using the transfer matrix method
considering plane wave propagation, visco-thermal losses X
N
pðtÞ ¼ A0 þ An ei2pnFt þ An ei2pnFt : (4)
(Causse et al., 1984), and a radiation function under an infi-
n¼1
nite plane baffle hypothesis.
It is important to assess the validity of this 1-DOF model The unknowns, i.e., the amplitudes and phases of the har-
by indicating which behaviors of brasses can be reproduced monics An and the playing frequency F, are determined using
adequately with simulations and which ones cannot. Newton’s method (Gilbert et al., 1989).
Previous results using this elementary model (Petiot and To perform a sound simulation, it is necessary to define
Gilbert, 2013) showed that the trumpet sounds simulated are relevant values (i.e., values that lead to a convergence
dissimilar to the real trumpet sounds played by a musician. toward a steady-state sound for a given note) for the parame-
In particular, sound spectra in permanent regimes are very ters of the musician embouchure. For a given note, experi-
different. This can be explained by inherent limitations of ence shows that countless embouchures may lead to a
the model. A first limitation concerns the linear approxima- steady-state note. The choice of the range of the parameters
tion of sound propagation in the brass resonator defined by is based both on numerical tests of the simulations and on
its acoustic impedance [Eq. (1)]. When the instrument is measurements of real trumpet players. The three variables
played loudly with brassy timbre, this approximation is no
Pm, l‘, and f‘ are considered control parameters of the simu-
longer valid, and the nonlinear propagation needs to be taken
lations, and constitute the virtual embouchure. The pressure
into account (Myers et al., 2012).
Pm in the mouth influences mainly the dynamics of a simu-
A second limitation concerns the lip model, which simpli-
lated sound and ranges from 1 to 12 kPa (Fletcher and
fies the complicated real-lips motion (Bromage et al., 2010;
Tarnopolsky, 1999). In the following numerical experiments,
Martin, 1942; Yoshikawa, 1995) with a 1-DOF [Eq. (2)]. Two-
we partitioned this range into three parts running from 1 to
degrees-of-freedom (2-DOF) models have been considered for
5 kPa for what we call piano (p) dynamics, 5 to 9 kPa for
time domain simulation (Adachi and Sato, 1996; Boutin et al.,
mezzoforte (mf) dynamics, and 9 to 12 kPa for fortissimo (ff)
2015). Furthermore, measured mechanical responses of artifi-
dynamics. The frequency of the lips f‘ enables the selection
cial (Cullen et al., 2000) and real brass-players lips (Newton
of the played regime (note): the higher the value of f‘, the
et al., 2008) revealed that a pair of mechanical resonances
higher the simulated regime. Exploration tests led to a range
requires a 2-DOF model in order to be consistent with near
for f‘ that spans from 130 to 480 Hz to simulate the second,
threshold oscillations (Cullen et al., 2000). However, although
additional terms can theoretically be added to the 1-DOF third, fourth, and fifth regime of the B[ trumpet with no
model, the difficulty lies in selecting realistic values for the valve pressed, the regimes considered in this study. These
additional parameters (Velut et al., 2017). regimes correspond to the musical notes B[3, F4, B[4,
A third limitation relates to the assumption that the volume D5–concert-pitch. Finally, in order to produce many differ-
flow u [Eq. (3)], which controls the valve effect, is proportional ent sounds for every regime, we add variability to the
to the opening height h between the lips. Experimental data embouchure making l‘ a control parameter of the simula-
reported in Bromage et al. (2010) for a large set of playing fre- tions, ranging from 1 to 6 kg/m2 (Cullen et al., 2000). In our
quencies and sound levels showed that the relationship may be study, the values of b, Q‘, and h0 are the same for every sim-
exponential instead of linear, and dependent on the pitch and ulation (Cullen et al., 2000).
dynamic level of the note played. The values of the control parameters considered in this
Nevertheless, even if nonrealistic for the spectrum of study are summarized in Table I.
trumpet sounds, previous studies confirm that this elementary Given that above 3000 Hz the impedance magnitude is
model behaves in agreement with the main physical princi- flat (see Fig. 1), it is not relevant to consider many harmon-
ples that govern the playing of brasses (Petiot and Gilbert, ics for the sound simulation. The highest studied note being
2013; Poirson et al., 2005). In particular, results presented in D5 (587 Hz), we chose to simulate our permanent regime
Petiot and Gilbert (2013) showed that the elementary model is with only N ¼ 6 harmonics. In conclusion, for a given
able to produce differences between instruments according to
playing frequency, spectral centroid (abbreviated hereafter SC), TABLE I. Values of the control parameters for the simulations considered
in the study (virtual musician embouchure).
and evolution of the SC with the playing dynamics that are, on
average, in agreement with the differences noticed when a real Parameter Symbol (units) Value
trumpet is played. These results justify the use of this elementary
Resonance frequency of the lips f‘ (Hz) 130 to 480
model in an optimization process for objective functions based
Mass per area of the lips l‘ (kg/m2) 1 to 6
on intonation, SC, or the evolution of the SC.
Pressure in the mouth Pm (kPa) 1 to 12
Numerical solutions of this system of equations are Width of the lips b (mm) 10
obtained using the harmonic balance technique to simulate Rest value of the opening height h0 (mm) 0.1
the sound created by a given trumpet (defined by its input Quality factor of the resonance Q‘ 3
impedance Z) for a given “virtual musician embouchure”

J. Acoust. Soc. Am. 145 (6), June 2019 Tournemenne et al. 3797
trumpet (characterized by its input impedance Z) and for a
virtual musician embouchure (characterized by the parame-
ters Pm, l‘, f‘, b, h0, and Q‘), the simulation may generate
one note (if the system converges), corresponding to one of
the regimes 2, 3, 4, and 5 of the trumpet. Each note is charac-
terized by its playing frequency F and the complex ampli-
tudes of its six first harmonics.
It is important to mention that the computed sound p(t)
corresponds to the sound in the mouthpiece. According to FIG. 3. Flowchart of the optimization process.
Benade (1966), a relevant spectrum transformation function
could be defined to compute the sound outside the instru-
future line of research. Consequently, for the P notes (P ¼ 4
ment. The difficulty lies in the definition of the radiated pres-
in this study), the harmonic balance technique simulates
sure, relevant from a perceptual point of view. For the
many sounds based on the calculated input impedance Z(x)
optimization considered in this work, the well-defined pres-
and a set of virtual embouchures ui for each note i. The set
sure in the mouthpiece is deemed sufficient. It is also impor-
tant to mention that the convergence of the simulation [
P
toward auto-oscillations is not ensured for a given shape of u¼ ui ; (6)
the resonator and embouchure. The search of convenient i¼1
embouchures is a complex task, described in Sec. III.
represents the entire set of embouchures used to evaluate
III. OPTIMIZATION PROBLEM FORMULATION one instrument (see Sec. III C for details). An objective func-
tion J is then computed using the playing frequencies F or
The design optimization problem of an instrument can the harmonics amplitudes An produced by the card(u) simu-
be formulated as the search for the optimal geometry mini- lations [card() denotes cardinality of a set, i.e., the number
mizing an objective function:
of elements in a set]. The value of the objective function
min JðxÞ; (5) (and the values of the constraint functions introduced later)
x2X are provided to the optimization algorithm [MADS, imple-
mented in the NOMAD software package (Le Digabel,
where J : Rn ! R is the objective function, and the vector 2011)], which will propose a new design vector x (see Sec.
x 2 Rn includes the design optimization variables. The IV for details). The new design candidate x is evaluated in
design space X is a subset of Rn delimited by box (bound) the same manner, and the process is iterated until the optimi-
constraints. The design optimization variables are the geomet-
zation algorithm termination criterion is satisfied.
ric parameters that define the inner shape of the bore. To
facilitate the input impedance calculations, the bore is approx- 1. Objective function based on intonation
imated by a series of conical and cylindrical waveguide seg-
ments. Consequently, x is a vector of geometric quantities For each note i, the average (across different embou-
such as the lengths and radii of cylinders or cones. The design chures) playing frequency Fi ðx; ui Þ at mezzoforte dynamics
space X may be modified to obtain viable trumpet shapes. is computed. The intonation of the note is assessed by the
Two classes of objective functions are available consid- deviation of the actual playing frequency, as simulated using
ering our physics-based model and the harmonic balance the physics-based model, from the expected playing fre-
technique: descriptors based on playing frequencies and quency. This objective function relies on a reference note
descriptors based on sounds spectra. Many quantities based from which the ideal and actual musical distance is com-
on frequency and sound spectrum can be found in the litera- puted. We chose for reference the fourth regime of the trum-
ture; however, there is no consensus in the community pet with no valve pressed (B[4, concert pitch), given that it
regarding their influence on the instrument’s musical quality. is the usual tuning note of the instrument. The equal-
These disagreements notwithstanding, this paper considers tempered scale is used to define the ideal distance between
three different objective functions based on intonation and the studied note i and the reference note (Tournemenne
SC given their recognized impact on the instrument quality et al., 2017).
(Deutsch, 2013). For every note i, we compute the equal-tempered devia-
tion (ETD) between the average frequency of the ith note
A. Objective functions Fðx; ui Þ and the reference frequency Fðx; ur Þ as
Figure 3 describes the flowchart of the process for opti- !
F ðx; ui Þ
mizing the shape of a trumpet bore using physics-based ETDi ðx; uÞ ¼ ar!i  1200 log2 ; (7)
sound simulations. Input impedance is computed for a design F ðx; ur Þ
vector x representing the resonator’s geometry. This study
focuses on the average behavior of the instrument across a where ar!i is the ideal difference between the reference note
panel of embouchures for each note (B[3, F4, B[4, D5). r and the targeted note i given by the equal-tempered scale
Another approach based on ideal embouchures for each note (500 cents, for example, between B[4 and F4). The objec-
could have been adopted, which would be an interesting tive function J1 ðx; uÞ for the whole instrument is the average

3798 J. Acoust. Soc. Am. 145 (6), June 2019 Tournemenne et al.
of the absolute deviation across the (P  1) notes (notice that B. Optimization problems
the deviation between the reference note and the fourth note
In summary, we consider three different optimization
is always equal to zero, given that it is the tuning note)
problems labeled Int, SC Int, and SC Dyn. They represent
1 X instrument design problems that are realistic from a musi-
J1 ðx; uÞ ¼ jETDi ðx; uÞj: (8) cians point of view:
P  1 i2Notes
• Int: intonation improvement:

min J1 ðxÞ; (13)


2. Objective function based on the SC x2X

The average SC, SCðx; uÞ, at mezzoforte dynamics for • SC Int: SC improvement under intonation constraint:
every note is computed according to the six harmonic ampli-
tudes An of each simulated sound max J2 ðxÞ subject to J1 ðxÞ  J1max ; (14)
x2X

1 X
SC ðx; uÞ ¼ SC½x; Ev ðPm ; ll ; fl Þ; (9) • SC Dyn: SC dynamics improvement:
cardðuÞ Ev 2u

max J3 ðxÞ: (15)


where Ev is an embouchure of the virtual musician with x2X

X
6
The constraint function of the SC Int problem is the
jn An ðx; Ev Þj objective function of the Int problem. It represents the sce-
n¼1
SCðx; Ev Þ ¼ : (10) nario where musicians are willing to trade some of their
X6
jAn ðx; Ev Þj instrument’s intonation quality for a brighter timbre.
n¼1
C. Finding playable embouchures
Consequently, SC(x, Ev) spans from 1 to 6, representing the
The main challenge lies in the simulation of many dif-
normalized SC. In this work, we decided to look for the
ferent sounds (represented by their permanent regimes) dur-
instruments having the highest SC. These instruments would
ing the numerical evaluation of the objective function.
generally be considered as bright by musicians (Poirson
Practically, for each geometry x, we need to find suitable vir-
et al., 2005). This is a somewhat arbitrary choice and other
tual embouchures leading to convergence toward a perma-
relevant descriptors/targets may be found, although such
nent regime. Furthermore, just as a beginner would blow
consideration is out of the scope of the paper. Consequently,
even the most well-designed trumpet with a terrible sound,
the objective function J2(x, u) is
the virtual embouchure must be carefully selected in order to
J2 ðx; uÞ ¼ SCðx; uÞ: (11) produce realistic sounds. No analytical approach exists to
deal with this challenge and simple solutions always fall
short. For example, it is not possible to define, a priori, a
fixed list of virtual embouchures that will be used for every
3. Objective function based on the SC dynamics geometry x, because experience shows that the intersection
This descriptor represents the ability of the instrument of the sets of virtual embouchures leading to convergence
to maximize the SC difference between a piano (p) and a for- toward a permanent regime for each geometry may be
tissimo (ff) dynamics. The idea is to find the instrument pro- empty. It is far too expensive to process a complete fine grid
ducing bright notes for high dynamics (high SC) while of the three virtual embouchure parameters for every geome-
keeping a dark sound (low SC) for low dynamics. The aver- try x. Consequently, a rigorous preprocessing of the simula-
age SC for each note and each piano [SCp ðx; ui Þ] and fortis- tions is undertaken to help the simulations obtain a set of
simo [SCf f ðx; ui Þ] dynamics are computed. The objective appropriate embouchures that converge toward adequate
function is then formulated as sounds for every geometry x.
This preprocessing is based on an exploration of the area
X SCf f ðx; ui Þ  SCp ðx; ui Þ of the design space X augmented by the three embouchure
J3 ðx; uÞ ¼
P parameters leading to convergence of the sound simulation. If
i2Notes
x is in R2 , the space to explore has five dimensions: two geo-
X DSC i
¼ : (12) metric variables and three embouchure variables (Pm, l‘, f‘).
i2Notes
P To explore this space, a five-dimensional Latin hypercube is
built and the harmonic balance technique tries to simulate
For this descriptor, the note D5 concert-pitch has been every sample.
discarded because of the difficulty in simulating it for low In order to discard the simulations of unrealistic sounds
dynamics. A more application-oriented study considering mentioned above, we use a criterion representing the ampli-
the entire instrument’s tessitura should account for these tude of the simulated sound relative to the pressure in the
kinds of difficulties. mouth. If the amplitudes of the harmonics are large enough

J. Acoust. Soc. Am. 145 (6), June 2019 Tournemenne et al. 3799
relative to the mouth pressure Pm produced by the virtual IV. SURROGATE-ASSISTED DERIVATIVE-FREE
musician, the sound is considered appropriate. This is OPTIMIZATION
expressed by the following inequation [Eq. (16)]:
We use the rigorous derivative-free MADS optimization
vffiffiffiffiffiffiffiffiffiffiffiffiffi algorithm, which has convergence properties (Audet and
u 6
uX 2 Dennis, 2006) and has been implemented in the NOMAD
t An
n¼1 software package (Le Digabel, 2011). Every iteration of the
 a: (16) MADS algorithm consists of two steps: the optional search
Pm
and the mandatory poll. The search step can implement any
Given the exploration and the criterion, an empirical user-defined strategy to obtain promising candidates. The
method finds adequate virtual embouchures for any geometry poll step determines candidates around the incumbent solu-
x of the design space. There are two key differences between tion; it ensures the convergence of the algorithm toward a
the preprocessing presented here and that reported in local optimum. Our strategy in the search step is to formulate
Tournemenne et al. (2017): (i) the definition of the criterion and solve a surrogate problem to obtain a promising candi-
threshold a based on live recordings and (ii) a more robust date, i.e., we use surrogate models of the computationally-
technique used to define the set of virtual embouchures u for intense simulation procedure to evaluate the objective and
any bore, both summed up in the following paragraph. constraint function values. We then evaluate the real poten-
The live recordings of three volunteer trumpeters play- tial of this promising candidate using the physics-based sim-
ing several times the four notes allowed us to estimate the ulations. In addition, when the MADS algorithm needs to
standard deviation of playing frequency for each note. We proceed to the poll step, we use the surrogate models to
then defined a in order to obtain simulations having approxi- rank-order the poll-generated candidates and then evaluate
mately the same standard deviation of playing frequency. In them opportunistically using the physics-based simulations.
practice we defined one a per note and dynamic (p, mf, ff); In this manner, we generate a large amount of information
its value ranges from 0.85 (D5) to 1.17 ðB[ 3Þ. During the using computationally inexpensive surrogate models but
optimization, the procedure defining u for every bore relies make algorithmic decisions using the high-fidelity simula-
on a maximal distance from the corresponding cloud of suc- tions. More details are provided in Secs. IV A and IV B.
cessful embouchures found during the preprocessing, above
which a virtual embouchure is discarded. In the interest of
keeping the length of this paper reasonable, we point the A. MADS
interested reader to the manual accompanying the source
At each iteration k of the MADS algorithm, the trial
code repository for a detailed description of the procedure
points must lie on a mesh Mk. The mesh size Dm k depends on
(Tournemenne, 2019a).
the iteration number k and gets smaller as the optimization
It is important to mention that for numerical reasons, the
converges.
quantities Fðx; ui Þ and SCðx; uÞ are average values across a
During each search step, a surrogate model J^ is built
finite set of embouchures u, randomly chosen and selected
using previous evaluations of the objective function J. Then,
by the preprocessing. The consequence is that the objective
a second instance of MADS is used to obtain the design that
function J(x) is non-deterministic, i.e., different objective ^ This candidate design is then projected on the
minimizes J.
function values may be obtained for the same x. In practice,
mesh Mk and J is evaluated. If this candidate leads to an
a set of 1000 embouchures are simulated per note per
dynamic range in order to keep the standard deviation on the improvement of the solution, the surrogate model J^ is
numerical estimation of J(x) as low as possible according to updated and the search is repeated. Otherwise, the algorithm
the law of large numbers. This choice of 1000 embouchures continues with the poll step. Two possible surrogate model-
is validated a posteriori given the small error bars in Fig. 5. ing techniques are described in Sec. IV B.
Additional “blackbox” properties of the objective functions During each poll step, a set of candidates Pk is generated
under consideration include: on the mesh Mk. The distance between the incumbent solution
and the candidates Pk is controlled by the poll parameter Dpk
• The evaluation of J(x) may fail due to difficulties to simu- which, as Dm k , gets smaller as the optimization converges. The
late notes (find virtual embouchures). interested reader can refer to Audet and Dennis (2006) for
• It is not possible to reliably compute the gradient of J(x) details. As mentioned earlier, we first rank-order the points of
because of the random selection process in the selection of the set Pk using the surrogate model J. ^ The physics-based
the virtual embouchure. sound simulation model J is then used to evaluate the points
• The evaluation of the objective function can be computa- of the set Pk using an opportunistic strategy: If a point is feasi-
tionally expensive (between 3 and 20 min depending on ble and leads to an improvement of the objective function, the
the processor). evaluation process is aborted and the algorithm iterates. Note
• We cannot assume smoothness of the objective (or con- that if a more feasible point is found during the poll step, the
straint) functions.
mesh and poll parameters are increased so that the algorithm
To address these issues, we resort to the use of can explore other areas of the design space. Otherwise, these
derivative-free optimization algorithms and a surrogate- parameters are reduced, which means that the iteration will
assisted modeling strategy. operate in a closer neighborhood of the design space.

3800 J. Acoust. Soc. Am. 145 (6), June 2019 Tournemenne et al.
B. Surrogate modeling strategy optimization jobs (3 problems  2 surrogate approaches 
10 starting points).
We consider two surrogate model approaches in this
In order to undertake this considerable computational
study:
endeavor, we relied on a high performance computing clus-
• An ensemble of surrogate models approach where several ter, parallelizing the sound simulations on two Haswell
R R
surrogates of different types and with different modeling IntelV XeonV E5-2680 v3 2.5 GHz processors providing 24
parameters are built and updated while selecting the model cores for each of the 60 jobs. This equipment is part of the
that fits the data best at each iteration. We use the same PlaFRIM experimental testbed (PlaFRIM, 2019). One evalu-
ensemble of surrogates as in Audet et al. (2018) and ation of the objective function took between 2 min 40 s and
Tournemenne et al. (2017): six polynomial regression 3 min 50 s for Int and SC Int (4000 sounds simulated),
models, five kernel smoothing models, and six radial basis depending on the difficulty to simulate sounds for the con-
function models. The best model is chosen according to sidered geometry x. We provide sounds of the studied notes
the order-error with cross-validation metric presented in of the initial trumpet as supplementary material.1 These
Audet et al. (2018). sounds are made of the first six harmonics of the permanent
• A LOcally WEighted Scatterplot Smoothing (LOWESS) regime (no transients).
(Talgorn et al., 2018) surrogate model. This model consists
of building a local polynomial regression around the point x A. Design optimization problem with ten variables
(10-d problem)
where we wish to predict the objective function. In the con-
struction of this local regression, data points close to x are The ten design variables represent the geometry of the
given more importance than those further away from x. leadpipe which is an important part of the bore that connects
the mouthpiece to the tuning slide. The leadpipe, roughly
conical, has a significant influence on the intonation and tim-
V. APPLICATION bre of the instrument (Petiot and Gilbert, 2013). Eleven parts
of equal length (l ¼ 20 mm) are considered. The design vari-
We consider three problem with ten design optimization
ables are the inner radii of the leadpipe at the connection
variables (10-d problem): Int, SC Int, and SC Dyn. We
between two parts (10 variables out of 12 control points
have also considered these problems with two design varia-
because the initial and last control points are fixed at 4.64
bles similar to Tournemenne et al. (2017), but we omit them
and 5.83 mm, respectively). These 10 inner radii values span
to keep the paper length reasonable. Since it can be useful to
from 4.5 to 6 mm. The rest of the instrument corresponds
optimization novices, we have made it available at approximately to a standard trumpet.
Tournemenne (2019b). In this realistic design space, the high dimensionality
The initial bore x0 is close to an existing trumpet bore requires efficient optimization: a discretization of the space
(whose internal diameter has been measured with different with a granularity of 0.5 mm (20% of the range of each
balls of decreasing diameter inserted in the trumpet, and a dimension) would necessitate 410 function evaluations (more
gauge to measure their position inside the instrument). than 1  106, which is not tractable in a reasonable computa-
The design problems are solved with a budget of 200 tion time).
function (or blackbox) evaluations. This maximum number Figure 4 presents the performance of the optimization
of function evaluations is defined empirically by observing approaches for the three 10-d problem formulations. The ini-
the evolution of the objective functions in order to minimize tial value for J1 is 8.7 cents. The initial values for J2 and J3
computational cost (i.e., avoid unnecessary evaluations that are 2.23 and 0.04 (value of SC), respectively, as can be seen
do not improve the function value significantly, see Fig. 4). in Fig. 5.
Moreover, to ensure a reliable quantification of the efficiency On average, the LOWESS surrogate approach yields
of the optimization method, each problem is solved 10 times slightly better designs for J1(x, u) and better results for J3(x,
(with different starting points) for each of the 2 surrogate u). For Int (J1) the best objective function value obtained is
modeling approaches (using either an ensemble of surrogates 0.1 cents, and the best intonation improvement is 8.6 cents.
according to Audet et al. (2018) or locally weighted regres- This represents a 99% improvement. For SC - Int (J2) the
sion models according to Talgorn et al. (2018). Given that 3 best objective function value obtained is 2.27 (value of SC),
design problems are studied, there is a total of 60 and the average SC improvement is 0.04. This represents a

FIG. 4. (Color online) Evolution of the objective function values for the three 10-d problem formulations; from left to right: Int, SC Int, and SC Dyn.

J. Acoust. Soc. Am. 145 (6), June 2019 Tournemenne et al. 3801
7.5 (ETD2) and 13.6 cents (ETD3), respectively, which are
superior to the classical just-noticeable difference (JND) of 5
cents. B[ 4 being the relative reference, it is always consid-
ered perfectly in tune. Since the initial value of D5 is already
low (4.8 cents) the improvement of 4.7 cents is slightly
lower than the JND. The sound simulation results concerning
intonation are in agreement with the impedance peak fre-
quency values except for D5 (ETD5), for which the optimal
trumpet seems less in tune when considering the impedance
peaks frequency values. For SC - Int, the highest improve-
ment is 0.08 (value of SC) for B[ 4 (SC 4) which is at the
same level than the JND of 0.1 reported in Jeong and Fricke
(1998). Consequently, for B[ 3, F4, and D5, the improve-
ments seem negligible, rising to 0.04, 0.01 (error bars level),
and 0.5, respectively. For SC Dyn, the improvements for
B[ 3 and F4 are above the JND [0.13 and 0.15 (DSC 2 and
DSC 3)], contrary to B[ 4 (0.01 DSC 4).
Figure 6 shows the leadpipes yielding the best value for
each design problem.
As in Tournemenne et al. (2017), the optima are counter
intuitive since the leadpipe does not have a positive slope
along the whole trumpet axis. This kind of shape for a lead-
pipe is not common among trumpets because it is very diffi-
cult to manufacture. The optimization algorithm was able to
explore the design space in order to find unusual designs.
Finally, the geometrical differences between the three
optima and the initial leadpipe are on the order of a millime-
ter for several control radii, which will lead to noticeably dif-
ferent instruments when manufactured.

B. Discussion
Concerning intonation, the objective function value based
FIG. 5. (Color online) On the top part, the columns represent from left to
right the J1 value and the non-zero ETDs. Two bores are evaluated, the ini- on sound simulations is always below the one obtained only
tial bore of all the optimization jobs and the best bore from the most suc- with the input impedance peaks (cf. black dots of Fig. 5). This
cessful job (out of the 20 available). The error bars equal to 2 standard has been verified on more examples (Tournemenne, 2017), and
deviations of the estimated quantity. Equivalent plots for the design prob-
lems SC - Int and SC Dyn are drawn below. SC i is SC restricted to ui. In
would mean that the musician always plays more in tune that
addition, the values of J1 and the corresponding ETDs, obtained with the would suggest the resonance frequency values of the input
input impedance peaks frequency values, are presented in black dots. impedance. This is necessarily an effect of the non-linear part
of the model which requires further study. The history plot of
1.8% overall improvement [(2.27 - 2.23)/2.23)]. For SC Dyn the Int design problem demonstrates a significant intonation
(J3) the best objective function value obtained is 0.14 (value improvement, above JND: the objective function is decreased
of SC), and the SC difference improvement is 0.1. This opti-
mal bore improves 35 times the J3 performance of the initial
bore (0.14/0.04). Yet, measurements on an expert trumpet
player playing B[4 concert pitch show a SC increase around
0.8 between a p and ff note for the six first harmonics, which
is more than 5 times superior to the simulated values of the
optimum. These results are in accordance with Petiot and
Gilbert (2013). Even if the method optimizes the SC varia-
tion, the model does not allow the prediction of realistic SC
values. Non-linear propagation in the bore should be taken
into account.
Figure 5 provides a finer acoustical analysis of these
results showing the contributions of each note to the objec-
tive functions for the initial bore and the optimal bore. The
FIG. 6. (Color online) Representation of the leadpipe inner radius along the
error bars are computed following the guidelines of the Joint
instrument axis; the black dotted line to the initial geometry (measured on
Committee for Guides in Metrology (JCGM, 2008). For Int, our trumpet); each other line corresponds to the best design found by each
the B[ 3 and F4 show significant improvements of objective over 20 jobs.

3802 J. Acoust. Soc. Am. 145 (6), June 2019 Tournemenne et al.
by 99% to 0.1 cents. Such a level of performance has been understanding may be translated to computational principles
achieved for a small price (200 evaluations compared with the and models.
million evaluations needed by brute force). Future studies may
consider even more design variables. Concerning the SC, the ACKNOWLEDGMENTS
results show improvements of the objective function J2 and J3,
The authors are grateful to the NOMAD team (Professor
even if they remain limited and close to the JND. More marked
Charles Audet, Professor Sebastien Le Digabel, and Dr.
improvements would certainly occur on real trumpets when
Christophe Tribes), Professor Sa€ıd Moussaoui for his help
played.
with the optimization problem formulation, Professor Jean-
Regarding the surrogate modeling approach, it is hard to
Pierre Dalmont for his precious help with the input
tell whether one is consistently better than the other as they
impedance models, and Juliette Chabassier for the many
seem to exhibit the same exploration to exploitation ratio of
helpful discussions. Numerical experiments were carried out
the design space. It should be noted that the locally weighted
using the PlaFRIM experimental testbed (PlaFRIM, 2019).
regression models take longer to compute than the ensemble
of surrogates in high-dimensional problems (in our case, the 1
See the supplementary material at https://doi.org/10.1121/1.5111346 for
difference is 1 h vs a few minutes for ten variables). the sounds.

VI. CONCLUSION AND PERSPECTIVE Adachi, S., and Sato, M.-A. (1996). “Trumpet sound simulation using a two-
dimensional lip vibration model,” J. Acoust. Soc. Am. 99(2), 1200–1209.
In this paper, we extended a new paradigm for design Audet, C., and Dennis, J. E., Jr. (2006). “Mesh adaptive direct search algorithms
optimization of brass instruments. We consider the optimiza- for constrained optimization,” SIAM J. Optimization 17(1), 188–217.
tion of objective functions (possibly subject to constraints) Audet, C., Kokkolaras, M., Le Digabel, S., and Talgorn, B. (2018). “Order-
based error for managing ensembles of surrogates in derivative-free opti-
based directly on the sounds spectrum, which, to the best of mization,” J. Global Optimization 70(3), 645–675.
our knowledge, is a novel approach. The main contribution Benade, A. H. (1966). “Relation of air-column resonances to sound spectra
of this paper is to demonstrate how physics-based sound sim- produced by wind instruments,” J. Acoust. Soc. Am. 40(1), 247–249.
ulations can be integrated in an iterative optimization algo- Boutin, H., Fletcher, N., Smith, J., and Wolfe, J. (2015). “Relationships
between pressure, flow, lip motion, and upstream and downstream impe-
rithm, which requires that simulations converge dances for the trombone,” J. Acoust. Soc. Am. 137(3), 1195–1209.
automatically toward auto-oscillations for every considered Braden, A. C. P., Newton, M. J., and Campbell, D. M. (2009). “Trombone
bore of the design space, without any assistance of the user. bore optimization based on input impedance targets,” J. Acoust. Soc. Am.
Applied to the optimization of a trumpet, the results show 125(4), 2404–2412.
Bromage, S., Campbell, M., and Gilbert, J. (2010). “Open areas of vibrating
that the optimization method, based on the MADS algo- lips in trombone playing,” Acta Acust. Acust. 96(4), 603–613.
rithm, is efficient to define optimal solutions in a reasonable Campbell, M. (2004). “Brass instruments as we know them today,” Acta
computation time, with or without constraints, for problems Acust. Acust. 90(4), 600–610.
Causse, R., Kergomard, J., and Lurton, X. (1984). “Input impedance of brass
up to ten design variables.
musical instruments—comparison between experiment and numerical
While the approach shows promising potential, there is models,” J. Acoust. Soc. Am. 75(1), 241–254.
room for improvement. Even if it is efficient to reproduce Chen, F.-C., and Weinreich, G. (1996). “Nature of the lip reed,” J. Acoust.
differences between instruments concerning intonation and Soc. Am. 99(2), 1227–1233.
Cullen, J., Gilbert, J., and Campbell, M. (2000). “Brass instruments: Linear
SC, the elementary model could be improved to generate a
stability analysis and experiments with an artificial mouth,” Acta Acust.
more realistic sound spectrum. The sound optimized in this Acust. 86(4), 704–724.
work is the sound in the mouthpiece. Even if this does not Deutsch, D. (2013). The Psychology of Music, 3rd ed. (Academic Press,
change the principle of the method presented, it could be New York).
Elliott, S., and Bowsher, J. (1982). “Regeneration in brass wind
interesting to define a relevant radiated pressure outside the
instruments,” J. Sound Vib. 83(2), 181–217.
instrument, and to include it in the optimization consider- Eveno, P., Petiot, J.-F., Gilbert, J., Kieffer, B., and Causse, R. (2014). “The
ations. Another valuable contribution to this numerical study relationship between bore resonance frequencies and playing frequencies
concerns the manufacturing of the optimal instruments and in trumpets,” Acta Acust. Acust. 100(2), 362–374.
Fletcher, N. H., and Tarnopolsky, A. (1999). “Blowing pressure, power, and
their objective and subjective study. This would help evalu-
spectrum in trumpet playing,” J. Acoust. Soc. Am. 105(2), 874–881.
ate actual improvement. Regarding implementation, the Gilbert, J., Kergomard, J., and Ngoya, E. (1989). “Calculation of the steady-
influence of the selection process of the virtual embouchure state oscillations of a clarinet using the harmonic balance technique,”
could be investigated further, as it may provide more robust J. Acoust. Soc. Am. 86(1), 35–41.
Guilloteau, A. (2015). “Conception d’une clarinette logique” (“Design of a
descriptors of the ease of playing the considered instruments.
logical clarinet”), Ph.D. thesis, Aix-Marseille.
Regarding the methodology, a study of temporal sound sim- JCGM (2008). “Evaluation of measurement data—guide to the expression
ulations could lead to new classes of objective functions, of uncertainty in measurement,” Technical Report.
such as attack times. More ambitious still, the inclusion of Jeong, D., and Fricke, F. R. (1998). “The dependence of timbre perception
on the acoustics of the listening environment,” in Proceedings of the 16th
non-linear propagation in simulations would produce more
International Congress on Acoustics and the 135th Meeting of the
realistic sounds improving the objective functions consid- Acoustical Society of America, Vol. 3, pp. 2225–2226.
ered in this work. Finally, the main challenge in the design Kausel, W. (2001). “Optimization of brasswind instruments and its applica-
of musical instrument lies in the definition of judicious tion in bore reconstruction,” J. New Music Res. 30(1), 69–82.
Le Digabel, S. (2011). “Algorithm 909: NOMAD: Nonlinear optimization
objective functions providing actual insight of the instrument with the MADS algorithm,” ACM Trans. Math. Softw. 37(4), 44:1–44:15.
intrinsic quality. This task may be accomplished by working Macaluso, C. A., and Dalmont, J.-P. (2011). “Trumpet with near-perfect har-
side-by-side with instrument makers whose empirical monicity: Design and acoustic results,” J. Acoust. Soc. Am. 129(1), 404–414.

J. Acoust. Soc. Am. 145 (6), June 2019 Tournemenne et al. 3803
Martin, D. W. (1942). “Directivity and the acoustic spectra of brass wind Talgorn, B., Audet, C., Le Digabel, S., and Kokkolaras, M. (2018). “Locally
instruments,” J. Acoust. Soc. Am. 13(3), 309–313. weighted regression models for surrogate-assisted design optimization,”
Myers, A., Pyle, R. W., Jr., Gilbert, J., Campbell, D. M., Chick, J. P., and Optimization Eng. 19(1), 213–238.
Logie, S. (2012). “Effects of nonlinear sound propagation on the character- Tournemenne, R. (2017). “Optimisation d’un instrument de musique de type
istic timbres of brass instruments,” J. Acoust. Soc. Am. 131(1), 678–688. cuivre basee sur des simulations sonores par modèle physique” (“Brass
Newton, M. J., Campbell, M., and Gilbert, J. (2008). “Mechanical response instrument optimization based on physics-based sound simulations”),
measurements of real and artificial brass players lips,” J. Acoust. Soc. Am. 
Ph.D. thesis, Ecole Centrale de Nantes.
123(1), EL14–EL20. Tournemenne, R. (2019a). “Brass instruments optimization based on sound
Noreland, J. O. D., Udawalpola, M. R., and Berggren, O. M. (2010). “A simulations,” https://framagit.org/rtournem/BrassOptimUsingSounds/ (Last
hybrid scheme for bore design optimization of a brass instrument,” viewed June 12, 2019).
J. Acoust. Soc. Am. 128(3), 1391–1400. Tournemenne, R. (2019b). “Description of the 2-d optimization problem of
Petiot, J.-F., and Gilbert, J. (2013). “Comparison of trumpets’ sounds played a trumpet,” https://framagit.org/rtournem/2D/blob/master/2D_study.pdf
by a musician or simulated by physical modelling,” Acta Acust. Acust. 98, (Last viewed June 12, 2019).
629–641. Tournemenne, R., Petiot, J.-F., Talgorn, B., Kokkolaras, M., and Gilbert, J.
PlaFRIM (2019). Plateforme Federative pour la Recherche en Informatique (2017). “Brass instruments design using physics-based sound simulation
et Mathematiques (Federative Platform for Research in Computer Science models and surrogate-assisted derivative-free optimization,” J. Mech.
and Mathematics), https://www.plafrim.fr/ (Last viewed June 12, 2019). Design 139(4), 041401.
Poirson, E., Petiot, J.-F., and Gilbert, J. (2005). “Study of the brightness of Velut, L., Vergez, C., Gilbert, J., and Djahanbani, M. (2017). “How well can
trumpet tones,” J. Acoust. Soc. Am. 118(4), 2656–2666. linear stability analysis predict the behaviour of an outward-striking valve
Poirson, E., Petiot, J.-F., and Gilbert, J. (2007). “Integration of user percep- brass instrument model?,” Acta Acust. Acust. 103(1), 132–148.
tions in the design process: Application to musical instrument opti- Yoshikawa, S. (1995). “Acoustical behavior of brass player’s lips,”
mization,” J. Mech. Design 129(12), 1206–1214. J. Acoust. Soc. Am. 97(3), 1929–1939.

3804 J. Acoust. Soc. Am. 145 (6), June 2019 Tournemenne et al.

You might also like