You are on page 1of 242

UNIVERSITE

PIERRE & MARIE


LA SCIENCE A P ARIS
CURIE

THÈSE de DOCTORAT de l’UNIVERSITÉ PARIS 6

Spécialité : Acoustique, Traitement du Signal et Informatique


Appliqués à la Musique

The Physics of Double-reed Wind


Instruments and its Application to Sound
Synthesis

Présentée par : André Almeida


pour obtenir le titre de Docteur de l’Université Paris 6

A soutenir le 26 Juin 2006 devant le jury composé de :

M. Xavier RODET Directeur de thèse (Université Paris 6, Paris)


M. René CAUSSE Directeur de thèse (IRCAM, Paris)
M. Christophe VERGEZ Encadrant (LMA, Marseille)
M. Jean-Pierre DALMONT Rapporteur (LAUM, Le Mans)
M. Murray CAMPBELL Rapporteur (Université d’Edimbourg)
M. Benoı̂t FABRE Examinateur (LAM, Paris)
M. Gary SCAVONE Examinateur (Université McGill, Montréal)
ii
To my parents,

To my sister, Ana,
iv
Acknowledgments

A thesis is not only the product of the PhD student but of many others that followed the
thesis and contributed in a greater or lesser extent to the completion of this project. To
all of them I wish to express my profound gratitude.
In particular, I want to thank Xavier Rodet, who accepted to supervise this thesis and
always trusted my scientific skills throughout this work.To René Caussé I wish to thank
the co-supervision of this thesis, to have hosted me in the Musical Acoustics team at
IRCAM, and for his continuous implication in this thesis especially in the experimental
work.
To Christophe Vergez, I cannot thank enough, for continuously supporting me through-
out this work, for the countless and fruitful scientific discussions both in the more
theoretical parts as in the experimental part, they were always the source of interesting
ideas that contributed to the accomplishment of my work. Thank you also for the sup-
port during the most difficult times of this thesis, and for not having given up during
these moments. To the wonderful colleague, the vigorous scientist and the joyful friend,
I hope we can share many more years of team work!
Still at IRCAM, I would like to leave my word of gratitude to all the colleagues and
friends that accompanied me during this journey. Firstly, a huge and warm thank you
to Claudia Fritz, with whom I shared the experience of the PhD during a great part of this
work, thank you for sharing your ideas, for the great ambiance at ‘‘labo 7’’ and for the
friendship that lasted beyond the scientific work. Many thanks also to Aude Liz ée, for
the joy she brought to our team during her ‘‘DEA’’ internship as well as for her scientific
contribution. To the interns that contributed directly to parts of this work, I thank their
commitment to the subject, Clémént Vern, who contributed to the measurements of the
reed dynamics, to Matthias Coulon, who improved the image analysis technique, worked
in the measurements of the elastic properties of the reed and gave a substantial help
in the anemometry measurements, and finally to Sylvain Hourcade, who contributed in
widening the variety of reeds for which the elastic properties were measured. Big thanks
to Alain Terrier, always ready to improve the experimental setups with new ideas and
for the brilliant execution of many parts needed for the experiments, to Gérard Bertrand
for his help with electronics devices and execution of required apparatuses, ideas for
experiments and hours of friendly discussions. Thanks to all the colleagues at IRCAM
for discussions or help in several specific problems, in particular Thomas H élie, Joël
Bensoam, Axel Roebel, and all those at the IRCAM research team whom I can’t thank
in particular for lack of space. . .
Many other scientists, musicians and instrument makers contributed to this work
through discussions in conferences, visits at IRCAM or collaborations in other labora-
tories. In particular I wish to thank Benoit Fabre, whose collaboration was essential for
vi

the experiments related to the flow, and who helped in many scientific issues in this
thesis. The list continues with Jean-Pierre Dalmont and Joël Gilbert for the interest
they showed for this subject and my work since the beginning of this thesis. Thanks
to Avraham Hirschberg, who contributed with many fruitful discussions and ideas for
the models, to Jean Kergomard, Kees Nederveen and Murray Campbell. I am also very
grateful for the interest shown by the musicians, in particular, David Rachor, Jim Kopp
and Franck Leblois, whose advises and ideas were very important to interpret and relate
many scientific results to the actual use of the double-reed and the instrument.
The support for such a long-term project does not only come from the scientific or
technical domains. Although most of the people referred to above also contributed with
personal support, many other people helped in encouraging me throughout this work.
I can only mention here some persons who provided me with constant human support,
they are my ‘‘parisian family’’, Bruno, Paula and Beatriz, André, Vale and Caterina, Rui
and Cris (even in Lisbon, a true parisian support!), Zé, Catarina and Luca. Thank you
to Kristina for the sincere professional concern and advises.
Thank you so much to my parents who never let me give up and who always have a
kind and effective word to encourage me.
Finally, a special thought to Leonor, whose life is probably, after mine, the most
influenced by this project. To you, thank you for all the words of encouragement and
your belief in my work!
Contents

I Introduction 1
1 The double-reed 3
1.1 Generic description of the double reed . . . . . . . . . . . . . . . . . . . 4
1.2 General working principles . . . . . . . . . . . . . . . . . . . . . . . . . . 6
1.3 Fabrication of the double-reed . . . . . . . . . . . . . . . . . . . . . . . . 6
1.4 Utilisation of the double-reed . . . . . . . . . . . . . . . . . . . . . . . . 9
1.4.1 Soaking . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9
1.4.2 Embouchure . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10
1.4.3 Articulation and dynamics . . . . . . . . . . . . . . . . . . . . . . 10
2 Reed models 13
2.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13
2.1.1 Overview of the chapter . . . . . . . . . . . . . . . . . . . . . . . 13
2.2 Self-sustained instruments . . . . . . . . . . . . . . . . . . . . . . . . . 13
2.2.1 Passive and active parts of the instrument . . . . . . . . . . . . . 14
2.2.2 Wind instruments . . . . . . . . . . . . . . . . . . . . . . . . . . 14
2.2.3 Reed instruments . . . . . . . . . . . . . . . . . . . . . . . . . . . 14
2.3 The resonator . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15
2.3.1 Wave equation and wave propagation . . . . . . . . . . . . . . . . 15
2.3.2 Propagation inside a cylindrical waveguide . . . . . . . . . . . . . 16
2.3.3 Diffusion and dispersion of the waves . . . . . . . . . . . . . . . . 17
2.3.4 Resonator termination and radiation . . . . . . . . . . . . . . . . 18
2.3.5 A complete cylindrical resonator . . . . . . . . . . . . . . . . . . . 19
2.3.6 Conical resonator models . . . . . . . . . . . . . . . . . . . . . . 20
2.3.7 Complex geometries . . . . . . . . . . . . . . . . . . . . . . . . . 22
2.4 The exciter . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23
2.4.1 Double reeds as opposed to single reeds . . . . . . . . . . . . . . 23
2.4.2 Description of the double reed . . . . . . . . . . . . . . . . . . . . 23
2.4.3 Dynamic effects of the reed . . . . . . . . . . . . . . . . . . . . . . 25
2.4.4 Considerations about the elementary model . . . . . . . . . . . . 26
2.5 State-of-the-art in double-reed physical modelling . . . . . . . . . . . . . 26

II Characterisation of the double-reed 29


3 Geometrical and Mechanical aspects of the reed 31
viii Contents

3.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 31
3.1.1 Overview of the chapter . . . . . . . . . . . . . . . . . . . . . . . 31
3.2 Cross-section profile of the double-reed . . . . . . . . . . . . . . . . . . . 32
3.3 Geometry of the reed opening . . . . . . . . . . . . . . . . . . . . . . . . 32
3.3.1 Experimental approaches . . . . . . . . . . . . . . . . . . . . . . 34
3.3.2 Observations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 36
3.3.3 Quantitative analysis . . . . . . . . . . . . . . . . . . . . . . . . . 38
3.3.4 Symmetry . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 40
3.3.5 Conclusions on the reed opening shape . . . . . . . . . . . . . . . 40
3.4 Reed equivalent stiffness . . . . . . . . . . . . . . . . . . . . . . . . . . . 41
3.5 Viscoelastic effects . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 43
3.5.1 Synthetic reed . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 44
3.5.2 Humidified natural cane reed . . . . . . . . . . . . . . . . . . . . 45
3.6 Partial conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 46
4 The nonlinear characteristics of the reed 47
4.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 47
4.1.1 Context . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 47
4.2 Principles of measurement and practical issues . . . . . . . . . . . . . . 48
4.2.1 Volume flow measurements . . . . . . . . . . . . . . . . . . . . . 48
4.2.2 Practical issues and solutions . . . . . . . . . . . . . . . . . . . . 49
4.2.3 Experimental set-up and calibrations . . . . . . . . . . . . . . . . 52
4.3 Results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 56
4.3.1 Double-reeds used in this study and operating conditions . . . . 56
4.3.2 The pressure vs flow characteristics . . . . . . . . . . . . . . . . . 57
4.3.3 Repeatability . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 58
4.3.4 Effect of the added mass . . . . . . . . . . . . . . . . . . . . . . . 60
4.3.5 Effect of humidity . . . . . . . . . . . . . . . . . . . . . . . . . . . 61
4.3.6 Variations between reeds . . . . . . . . . . . . . . . . . . . . . . . 62
4.3.7 Differences between bassoon and oboe reeds . . . . . . . . . . . . 63
4.3.8 Comparison with single-reeds . . . . . . . . . . . . . . . . . . . . 64
4.3.9 Comparison with synthetic reeds . . . . . . . . . . . . . . . . . . 65
4.4 Partial conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 66
4.4.1 Differences relative to the elementary model . . . . . . . . . . . . 66
4.4.2 Other remarks . . . . . . . . . . . . . . . . . . . . . . . . . . . . 67
5 Details of the flow inside the reed 69
5.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 69
5.2 Overview . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 69
5.3 Difficulties in measuring a complete velocity field . . . . . . . . . . . . . 70
5.4 Visualisation of the flow at the reed output . . . . . . . . . . . . . . . . . 71
5.5 Hot-wire measurements . . . . . . . . . . . . . . . . . . . . . . . . . . . 73
5.5.1 Experimental device . . . . . . . . . . . . . . . . . . . . . . . . . 73
5.5.2 Calibration of the hot wire . . . . . . . . . . . . . . . . . . . . . . 75
5.5.3 Double-reeds used in these measurements . . . . . . . . . . . . . 75
Contents ix

5.6 Velocity profiles . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 75


5.6.1 Variation of diametrical profiles with the mouth pressure . . . . . 75
5.6.2 Variation of diametrical profiles with the reed opening . . . . . . . 77
5.6.3 Comparison of two perpendicular diameters . . . . . . . . . . . . 77
5.7 General discussion on the flow profiles . . . . . . . . . . . . . . . . . . . 79
5.8 Volume flow calculations . . . . . . . . . . . . . . . . . . . . . . . . . . . 81
5.8.1 Volume flow deduced from the maximum of the velocity profile . . 81
5.8.2 Variation of flow with pressure . . . . . . . . . . . . . . . . . . . . 83
5.9 Average flows deduced from characteristics measurements . . . . . . . . 87
5.9.1 Small openings . . . . . . . . . . . . . . . . . . . . . . . . . . . . 88
5.9.2 Mean flow velocity estimations . . . . . . . . . . . . . . . . . . . . 89
5.10 Static flow model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 90
5.10.1 Vena contracta . . . . . . . . . . . . . . . . . . . . . . . . . . . . 90
5.10.2 Singular losses . . . . . . . . . . . . . . . . . . . . . . . . . . . . 91
5.10.3 Conical diffuser . . . . . . . . . . . . . . . . . . . . . . . . . . . . 92
5.10.4 Complete model . . . . . . . . . . . . . . . . . . . . . . . . . . . . 96
5.11 Conclusion on the static flow model . . . . . . . . . . . . . . . . . . . . . 100
6 Auto-oscillations in the double-reed 105
6.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 105
6.1.1 Overview of the chapter . . . . . . . . . . . . . . . . . . . . . . . 105
6.2 Discussion on dynamic flow effects . . . . . . . . . . . . . . . . . . . . . 106
6.2.1 Effects of inertia in the air flow . . . . . . . . . . . . . . . . . . . 106
6.2.2 Compressibility . . . . . . . . . . . . . . . . . . . . . . . . . . . . 106
6.2.3 Fluid-structure coupling . . . . . . . . . . . . . . . . . . . . . . . 107
6.3 Periodic motion of the reed . . . . . . . . . . . . . . . . . . . . . . . . . . 108
6.3.1 Two-state movement . . . . . . . . . . . . . . . . . . . . . . . . . 108
6.3.2 Time relations between open and closed states . . . . . . . . . . . 109
6.3.3 Closed to open transition . . . . . . . . . . . . . . . . . . . . . . . 109
6.3.4 Open to closed transition . . . . . . . . . . . . . . . . . . . . . . . 110
6.3.5 Oscillations during the open state . . . . . . . . . . . . . . . . . . 110
6.3.6 Effect of biting . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 110
6.4 Dynamic flow characterization . . . . . . . . . . . . . . . . . . . . . . . . 111
6.4.1 Procedure . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 112
6.4.2 Experimental conditions . . . . . . . . . . . . . . . . . . . . . . . 113
6.4.3 Obtained data . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 113
6.4.4 Sketch of the flow evolution . . . . . . . . . . . . . . . . . . . . . 119
6.4.5 Conjectures on the flow in a coupled reed . . . . . . . . . . . . . 121
7 Retrospective 123
7.1 Summary of part II . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 123
7.1.1 Geometrical aspects of the reed opening . . . . . . . . . . . . . . 123
7.1.2 Elastic properties . . . . . . . . . . . . . . . . . . . . . . . . . . . 124
7.1.3 Vibrational properties . . . . . . . . . . . . . . . . . . . . . . . . 124
7.1.4 Non-linear characteristics . . . . . . . . . . . . . . . . . . . . . . 125
x Contents

7.1.5 Flow inside the reed . . . . . . . . . . . . . . . . . . . . . . . . . 125


7.1.6 Dynamic regimes . . . . . . . . . . . . . . . . . . . . . . . . . . . 126
7.1.7 Dynamic flow . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 126
7.2 Retained facts for a double-reed model . . . . . . . . . . . . . . . . . . . 127

III Towards a numerical model of a double reed instrument 129


8 Synthesis 131
8.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 131
8.1.1 Historical background . . . . . . . . . . . . . . . . . . . . . . . . 131
8.1.2 A digital model of a reed instrument . . . . . . . . . . . . . . . . . 132
8.1.3 Overview of the chapter . . . . . . . . . . . . . . . . . . . . . . . 132
8.2 Digital model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 132
8.2.1 Implementation of cylindrical resonators . . . . . . . . . . . . . . 133
8.2.2 Implementation of conical resonators . . . . . . . . . . . . . . . . 134
8.2.3 Other elements of the bore . . . . . . . . . . . . . . . . . . . . . . 134
8.2.4 Exciter model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 141
8.2.5 Coupling the exciter to the resonator . . . . . . . . . . . . . . . . 143
8.3 Application of experimental data . . . . . . . . . . . . . . . . . . . . . . . 144
8.4 Control of the model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 144
8.5 Results of the simulation . . . . . . . . . . . . . . . . . . . . . . . . . . . 145
8.5.1 General remarks on the time-evolution curves . . . . . . . . . . . 147
9 Conclusion 149
9.1 Quasi-static model of the reed . . . . . . . . . . . . . . . . . . . . . . . . 149
9.1.1 The double reed as a nonlinear exciter . . . . . . . . . . . . . . . 149
9.1.2 Reed opening geometry . . . . . . . . . . . . . . . . . . . . . . . . 150
9.1.3 Elastic model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 150
9.1.4 Sound synthesis . . . . . . . . . . . . . . . . . . . . . . . . . . . 151
9.2 Open questions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 151
9.3 Perspectives . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 152
A Organology 157
A.1 Classification of double reed instruments . . . . . . . . . . . . . . . . . . 157
A.2 Variability of reed shapes . . . . . . . . . . . . . . . . . . . . . . . . . . . 159
B Preliminary modal analysis 163
B.1 Mechanical response of the double-reed . . . . . . . . . . . . . . . . . . 163
B.2 Modes of vibration of the reed . . . . . . . . . . . . . . . . . . . . . . . . 163
B.3 An oscillator driven by an external force . . . . . . . . . . . . . . . . . . 164
B.4 Frequency analysis of reed vibration . . . . . . . . . . . . . . . . . . . . 164
B.4.1 Experimental approaches . . . . . . . . . . . . . . . . . . . . . . 164
B.4.2 Data analysis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 165
B.4.3 Comparison of excitation methods . . . . . . . . . . . . . . . . . . 165
B.5 Admittance spectra of bassoon reeds . . . . . . . . . . . . . . . . . . . . 166
Contents xi

B.5.1 Comparison of two different cane reeds . . . . . . . . . . . . . . . 166


B.5.2 Effect of humidity . . . . . . . . . . . . . . . . . . . . . . . . . . . 167
B.5.3 Comparison between plastic and cane reeds . . . . . . . . . . . . 168
B.5.4 Oboe reeds . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 168
B.6 Discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 169
B.6.1 Importance of the dynamic aspects of the reed . . . . . . . . . . . 170
B.6.2 Importance of higher-order modes of vibration of the reed . . . . . 170
C Momentum 173
C.1 Differential form . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 173
C.2 Integral form . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 173
C.3 Vena Contracta . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 173
C.4 Singular losses . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 175
C.5 Potential theory for the Vena contracta . . . . . . . . . . . . . . . . . . . 175
C.5.1 Ecoulements par des orifices . . . . . . . . . . . . . . . . . . . . . 176
C.5.2 Borda tube ( = −π) . . . . . . . . . . . . . . . . . . . . . . . . . 180
D Viscous effects 183
D.1 How important are viscous losses? . . . . . . . . . . . . . . . . . . . . . 183
D.2 Turbulent flows . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 184
D.3 Discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 185
E Effects of gravity 187
F Elasticity data 189
G Artificial mouth 191
G.1 Artificial mouth . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 191
G.2 Stroboscope . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 193
G.2.1 Aim of the device . . . . . . . . . . . . . . . . . . . . . . . . . . . 193
H Image Analysis 195
H.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 195
H.1.1 Image analysis in musical instrument acoustics . . . . . . . . . . 195
H.1.2 Challanges of automatic image analysis . . . . . . . . . . . . . . . 196
H.2 Lighting balance . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 196
H.3 Binarization . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 197
H.4 Identification of the reed opening . . . . . . . . . . . . . . . . . . . . . . 197
H.4.1 Removing thin connections or cracks . . . . . . . . . . . . . . . . 198
H.4.2 Reducing the number of regions . . . . . . . . . . . . . . . . . . . 198
H.4.3 Segmentation and region identification . . . . . . . . . . . . . . . 198
H.5 Geometrical measurements . . . . . . . . . . . . . . . . . . . . . . . . . 200
H.6 Discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 200
H.6.1 Difficulties encountered during image processing . . . . . . . . . 200
H.7 Notions on mathematical morphology . . . . . . . . . . . . . . . . . . . 203
I Signal processing 205
I.1 Signal synchronisation . . . . . . . . . . . . . . . . . . . . . . . . . . . . 205
xii Contents

I.1.1 Splitting the periods of a signal . . . . . . . . . . . . . . . . . . . 205


I.1.2 Period statistics . . . . . . . . . . . . . . . . . . . . . . . . . . . . 206
I.1.3 Image analysis and reed opening area . . . . . . . . . . . . . . . . 206
I.1.4 Flow reversal . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 206
J Topography methods 211
J.1 Hand measurements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 211
J.2 Artisanal scanner . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 211
List of Figures

1.1 Modern double-reed instruments . . . . . . . . . . . . . . . . . . . . . 3


1.2 Double-reeds . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4
1.3 Terms used to describe the regions of double-reeds . . . . . . . . . . . 5
1.4 Reed making . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7
1.5 Thickness regions in oboe reed . . . . . . . . . . . . . . . . . . . . . . . 8

2.1 Cylindrical resonator and associated coordinates and variables . . . . . 17


2.2 Reflection coefficient . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19
2.3 Impedance of a cylindrical resonator . . . . . . . . . . . . . . . . . . . 19
2.4 Coordinates and variables used on conical resonator description . . . . 20
2.5 Impedance of a cylindrical resonator . . . . . . . . . . . . . . . . . . . 22
2.6 Non-linear characteristics of a reed . . . . . . . . . . . . . . . . . . . . 25

3.1 oboe reed internal profiles . . . . . . . . . . . . . . . . . . . . . . . . . 33


3.2 Reed opening designs . . . . . . . . . . . . . . . . . . . . . . . . . . . . 34
3.3 Pressure distribution in the double-reed . . . . . . . . . . . . . . . . . 35
3.4 Front view of an oboe reed during closure . . . . . . . . . . . . . . . . . 37
3.5 Area vs inter-blade distance . . . . . . . . . . . . . . . . . . . . . . . . 39
3.6 Time evolution of the area swept by each blade . . . . . . . . . . . . . . 41
3.7 Stiffness measurements (S vs ∆p) . . . . . . . . . . . . . . . . . . . . . 42
3.8 S(t ) and ∆p(t ) during stiffness measurements . . . . . . . . . . . . . . 43
3.9 Relaxation of the reed . . . . . . . . . . . . . . . . . . . . . . . . . . . . 44
3.10 Relaxation in a characteristics measurement . . . . . . . . . . . . . . . 46

4.1 Diaphragm mounted on the reed . . . . . . . . . . . . . . . . . . . . . . 49


4.2 Reed and reed plus diaphragm characteristics compared . . . . . . . . 50
4.3 Addition of masses to the oboe reed . . . . . . . . . . . . . . . . . . . . 52
4.4 Device used for characteristics measurements. . . . . . . . . . . . . . . 52
4.5 Diaphragm geometry . . . . . . . . . . . . . . . . . . . . . . . . . . . . 53
4.6 Diaphragm calibrations . . . . . . . . . . . . . . . . . . . . . . . . . . . 54
xiv List of Figures

4.7 Pressure acquisitions plotted against the time . . . . . . . . . . . . . . 55


4.8 Reed opening and pressure acquisitions plotted against time . . . . . . 56
4.9 A typical flow vs pressure characteristics . . . . . . . . . . . . . . . . . 58
4.10 Repeatability of the characteristics measurements . . . . . . . . . . . . 59
4.11 Non dimensional comparison of characteristics for a same reed . . . . . 59
4.12 Effect of the added mass on the reed characteristics . . . . . . . . . . . 60
4.13 Effect of humidity on a reed characteristics . . . . . . . . . . . . . . . . 62
4.14 The characteristics of two different reeds compared . . . . . . . . . . . 63
4.15 Comparison of the characteristics for different instrument reeds . . . . 64
4.16 Normalised characteristics of different reeds . . . . . . . . . . . . . . . 65
4.17 The characteristics of natural and synthetic reeds compared . . . . . . 66

5.1 Flow at the reed output (Schlieren) . . . . . . . . . . . . . . . . . . . . 72


5.2 Experimental device . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 73
5.3 Reed positioning device . . . . . . . . . . . . . . . . . . . . . . . . . . . 74
5.4 Definition of the measuring region . . . . . . . . . . . . . . . . . . . . . 74
5.5 Velocity profiles vs pressure . . . . . . . . . . . . . . . . . . . . . . . . 76
5.6 normalised velocity profiles vs pressure . . . . . . . . . . . . . . . . . . 76
5.7 Velocity profiles vs reed opening . . . . . . . . . . . . . . . . . . . . . . 78
5.8 Perpendicular diametral velocity profiles . . . . . . . . . . . . . . . . . 78
5.9 Differences between the measured profile and the real profile . . . . . . 80
5.10 Velocity profile in a cylindrical pipe . . . . . . . . . . . . . . . . . . . . 81
5.11 Volume flow q vs umax . . . . . . . . . . . . . . . . . . . . . . . . . . . 83
5.12 Volume flow q vs pressure pm . . . . . . . . . . . . . . . . . . . . . . . 84
5.13 Input flow velocity vs pressure . . . . . . . . . . . . . . . . . . . . . . . 84
5.14 average velocity u vs pressure pm . . . . . . . . . . . . . . . . . . . . . 85
5.15 Volume flows (hot-wire compared to diaphragm) . . . . . . . . . . . . . 86
5.16 Corrected volume flows from hot-wire measurements . . . . . . . . . . 87
5.17 Extrapolation of the residual opening area . . . . . . . . . . . . . . . . 88
5.18 Flow velocity at the reed entrance . . . . . . . . . . . . . . . . . . . . . 89
5.19 Flow regions in the reed . . . . . . . . . . . . . . . . . . . . . . . . . . 90
5.20 Pressure recovery vs contraction coefficient CD . . . . . . . . . . . . . . 92
5.21 Reed pressure vs mouth pressure . . . . . . . . . . . . . . . . . . . . . 94
5.22 Comparison of flow models . . . . . . . . . . . . . . . . . . . . . . . . . 95
5.23 Recovery coefficient in the staple . . . . . . . . . . . . . . . . . . . . . . 95
5.24 Reed tip non-linear characteristics . . . . . . . . . . . . . . . . . . . . 96
List of Figures xv

5.25 Non-linear characteristics with and without Vena Contracta . . . . . . 97


5.26 Non-linear characteristics with conical diffuser . . . . . . . . . . . . . . 98
5.27 Pressure recovery deduced from experimental characteristics . . . . . . 98
5.28 Reed tip non-linear characteristics determined from experimental data 99
5.29 Stiffness compared in flow and flow-less systems . . . . . . . . . . . . . 100
5.30 Mach number of the flow inside the reed . . . . . . . . . . . . . . . . . 102

6.1 Measured opening area S(t ) . . . . . . . . . . . . . . . . . . . . . . . . 108


6.2 S(t ) for two different bitings . . . . . . . . . . . . . . . . . . . . . . . . 111
6.3 Experimental device for dynamic anemometry . . . . . . . . . . . . . . 112
6.4 Impedance curves calculated for the reed duct (reed and staple) . . . . 114
6.5 Time evolution of a diametrical velocity profile . . . . . . . . . . . . . . 115
6.6 Time evolution of an axial velocity profile . . . . . . . . . . . . . . . . . 117
6.7 Snapshots of diametrical profiles for a reed with lips . . . . . . . . . . . 118
6.8 Influence of lips on the flow evolution . . . . . . . . . . . . . . . . . . . 119
6.9 Influence of lips on the opening area evolution . . . . . . . . . . . . . . 120
6.10 Sketch of the flow evolution . . . . . . . . . . . . . . . . . . . . . . . . 120

8.1 Scattering junction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 135


8.2 Scattering junction used for a tone hole . . . . . . . . . . . . . . . . . . 140
8.3 Simulated pressure and reed position at note attack . . . . . . . . . . . 145
8.4 Simulated pressure and reed position . . . . . . . . . . . . . . . . . . . 146
8.5 Simulated pressure and reed position . . . . . . . . . . . . . . . . . . . 146
8.6 Simulation for variable mouth pressures . . . . . . . . . . . . . . . . . 148
8.7 Simulation for variable dampings . . . . . . . . . . . . . . . . . . . . . 148

A.1 Reed dimensions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 161

B.1 Response of a bassoon reed . . . . . . . . . . . . . . . . . . . . . . . . 166


B.2 Comparison of two different cane reeds. . . . . . . . . . . . . . . . . . . 167
B.3 Comparison between different levels of humidification. . . . . . . . . . 168
B.4 Comparison between plastic and cane reeds. . . . . . . . . . . . . . . . 168
B.5 Mechanical response of an oboe reed. . . . . . . . . . . . . . . . . . . . 169

C.1 Flow into a Borda tube . . . . . . . . . . . . . . . . . . . . . . . . . . . 174


C.2 Singular losses . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 175
C.3 Flow in the complex spatial plane (z) . . . . . . . . . . . . . . . . . . . 176
C.4 Hodograph — Set of velocities (u) found in the flow . . . . . . . . . . . 177
xvi List of Figures

C.5 Transformed hodograph (ζ ) . . . . . . . . . . . . . . . . . . . . . . . . . 178


C.6 Flow potential set (w . . . . . . . . . . . . . . . . . . . . . . . . . . . . 178
C.7 Auxiliary variable v . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 179
C.8 Jet profile . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 181
C.9 Flow shape for the Borda tube . . . . . . . . . . . . . . . . . . . . . . . 181

F.1 Fibercane reed stiffness . . . . . . . . . . . . . . . . . . . . . . . . . . . 189


F.2 Dry cane reed stiffness . . . . . . . . . . . . . . . . . . . . . . . . . . . 190
F.3 Fibercane reed stiffness . . . . . . . . . . . . . . . . . . . . . . . . . . . 190

G.1 Artificial mouth . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 192


G.2 Reed plcaed inside the artificial mouth . . . . . . . . . . . . . . . . . . 192

H.1 Image analysis procedure . . . . . . . . . . . . . . . . . . . . . . . . . . 196


H.2 Original image and binarisataion . . . . . . . . . . . . . . . . . . . . . 197
H.3 Morphological operations . . . . . . . . . . . . . . . . . . . . . . . . . . 198
H.4 Inversion and flood-filling . . . . . . . . . . . . . . . . . . . . . . . . . . 199
H.5 Problems arising for small openings . . . . . . . . . . . . . . . . . . . . 199
H.6 Complete image analysis procedure . . . . . . . . . . . . . . . . . . . . 201
H.7 Fitting an ellipse to the reed opening . . . . . . . . . . . . . . . . . . . 202

I.1 Flow velocity cartography . . . . . . . . . . . . . . . . . . . . . . . . . . 208


I.2 Unfolding the velocity signal . . . . . . . . . . . . . . . . . . . . . . . . 209

J.1 Scanner setup . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 212


J.2 Analysis of a scanner image . . . . . . . . . . . . . . . . . . . . . . . . 212
J.3 Three-dimensional reconstitution of a reed cast . . . . . . . . . . . . . 213
List of Symbols

Observations,
Symbol Description Dimensions
Typical values
Sj
α Coefficient of Vena Contracta – α= S

f Frequency s−1

h Reed opening height m 0 < hoboe . 10−3


k Reed stiffness N m−1 = kg s−2 ?
νair = 1.44 × 105 ,
ν Kinematic viscosity m2 s−1
νCO2 = 7.6 × 104

m Reed mass kg ?
µair = 1.8 × 105 ,
µ Dynamic viscosity kg m−1 s−1
µCO2 = 1.48 × 105

ω Angular frequency rad/s ω = 2πf

patm Atmospheric pressure Pa (kg m−1 s−2 ) patm ' 105 Pa

pm Mouth pressure 1
Pa (kg m−1 s−2 ) 0 . p m . PM
pr Pressure at the reed output 1
Pa (kg m−1 s−2 ) −pm . pr (t ) . pm
PM ' 25kPa
PM Reed beating pressure Pa (kg m−1 s−2 )
(in the oboe)
PT maximum-flow2 pressure Pa (kg m−1 s−2 ) -

∆p Pressure drop in the reed Pa ∆p , pM − pR


Pressure drop in the di-
(∆p)d Pa (∆p)d , pR − patm
aphragm (chap. 4)
  1/ 2
q Volume flow m3 s−1 0 < q . 332/2 2ρ k 1/2 S03/2

r Reed damping kg s−1 ?


ud
Re Reynolds Number – Re = ν

S Reed opening area m2 0 < S . S0

1
relative to the atmospheric pressure
2
in a non-linear characteristics curve
xviii List of Figures

S0 Equilibrium reed opening area m2 0 < S0 ' 2 × 10−6


Effective area of pressure ap-
Sa m2 Sa ' 50 × 10−6
plication
Sj Jet cross-section m2

Sdiaph Diaphragm cross-section m2


ρair = 1.2,
ρ Volumic mass kg m−3
ρCO2 = 1.82
10 – 200 m s−1
u Flow velocity m s−1
(in the oboe reed)
Zin Resonator input impedance kg m−4 s−1
Preface

This work started as an attempt to propose a synthesis algorithm for double-reed in-
struments. In general, reed instruments, especially the clarinet have been the subject
of numerous studies, starting off with [Backus, 1961]. A model of the clarinet based on
basic physical assumptions about the reed and the flow, and the acoustic propagation
in the bore was implemented as early as the 1980 by [Schumacher, 1981]. Since then,
a lot of works have studied both the behavior of simple models, and the refinements to
bring to these models to improve their sound quality and realism.
Unfortunately, double-reed instruments such as the oboe or the bassoon were not
the object of such a great number of studies as the clarinet. One of the reasons for this,
is that these instruments are not as popular as single-reed instruments (starting with
the saxophone, an to a lesser extent the clarinet), so that the industrial and scientific
investment in this kind of instruments is much more scarce than the former.
Oboes and bassoons (the most common double-reed instruments in music perfor-
mances nowadays) use conical bores as their resonators. Since the acoustical behavior
of conical resonators is more difficult to interpret than straight ones, it is natural that
scientists base their work on cylindrical resonator instruments while researching for the
details in the exciters, and studying the instrument behavior. Nevertheless, the interest
in conical resonators is considerable, mainly because of the popular saxophone, so that
there exist resonator models that can either replace the conical resonator with others
of similar response , implement conical resonators or even more realistic models that
take into account the deviations from standard shapes.
Today, the most important difficulty that one faces when trying to implement a model
of an oboe or bassoon, is the lack of experimental and theoretical data about the double-
reed, or even arguments in favor or against the utilisation of a generic reed model for
the case of double-reeds. Other than the lesser interest on double-reeds, there is a
difficulty that explains this lack: the small dimensions of the reed. These make it very
difficult to study both mechanical and aerodynamic properties of these exciters.
It is thus natural that the work of this thesis followed a more experimental path, in
order to fill in some of the gaps existing in scientific literature.
It was soon realised that the interest in this kind of experimental data went beyond the
application to sound synthesis. In fact, the particular relation of double-reed musicians
with the reeds makes the art of reed making a passionating as much as puzzling subject.
For musicians, reed-making is a matter of empirical learning, much as the instrument
performance itself. Not only it is difficult to translate their experience on the techniques
and their effects into scientific terms, it is also a fact that the technique is widely
variable from one musician to another, and in certain cases, one particular gesture can
be described as having opposite effects, when different musicians are questioned.
xx List of Figures

Due to the time constraints of a PhD work, this work had to concentrate more on
general aspects of the double-reeds than on the details of different scrapping or gouging
techniques, so that most of this work can seem insufficient to oboists and bassoonists.
The author and the team that collaborate in this work are aware of this weakness, and
of the need of bringing this work to a level of detail that can be useful to the musicians
in their practice.

Overview
This work is divided into three main parts, an introduction (part I) describing the double-
reed and generic models of double-reed instruments, an experimental characterisation
of the double-reed (part II) and finally an attempt to put the experimental data together
with the classical model of reed instruments in order to propose a synthesis algorithm
for double-reed instruments (part III).
In chapter 1, we describe the double-reed, its fabrication and its use in musical
contexts. Chapter 2 provides the generic model of reed instruments and a review of the
studies found in scientific literature about the double-reed.
Chapter 3 introduces some basic aspects about the double-reed, such as a character-
isation of the reed opening, and its variations in different opening states, measurements
of basic characteristics like effective stiffness and viscoelasticity effects, and finally an
overview of the modes of vibration of the reed independently from the flow aspects.
In chapter 4 the static non-linear characteristics of the reed is measured and com-
pared for various kinds of reeds, and different environment conditions. Chapter 5 tries
to investigate deeper the aspects of the static flow inside the reed and proposes a flow
model to explain the measured non-linear characteristics.
Chapter 6 briefly studies the periodic motion of the reed, and investigates the details
of the flow in dynamic regimes.
The main experimental results are finally collected in chapter 7, which also introduces
the main features applied to the synthesis model.
Finally, chapter 8 collects all the experimental data from the previous chapters, dis-
cusses the modifications to be brought to the generic model of the reed in order to build
a model of the double-reed and finally presents the implementation of a double-reed
instrument
Part I.

Introduction
1. The double-reed
There are several instruments in the double reed family. Undoubtedly, the more well-
known ones are the modern oboe and bassoon, which take part in the modern orchestra
(fig. 1.1). The resonators in both these instruments are mainly conical throughout most
of their lengths, and this fact contributes to the specific timbre which we associate to
double reed instruments. For instance, while in the clarinet most of the sound energy
is concentrated in odd harmonics, giving it a round sound, notes played by oboes or
bassoons present both even and odd harmonics in similar proportions.

Figure 1.1.: The oboe and the bassoon, the two most common double-reed instruments
used in the modern orchestra

The first question that should be asked concerns the particularity of double-reed
instruments. They are particular from the morphologic point of view. But is it possible
to associate a particular sound to these instruments? The fact is that their sound can
be very variable. From the nasal sound of the oboe, to a more strident sound associated
to instruments played without lip support (like the bombarde), or the surprisingly flute-
like sound of the guanzi ( ), their timbre is all but homogeneous. This however does
not imply that their physics is inherently different, because the way they are blow is
also highly variable.
However the conical resonator may not be the only explanation for the specific color
of their sound. As an example, the saxophone is a case that mixes single-reeds to
conical resonators, and still its sound is very easily distinguishable from instruments
associating double-reeds to conical resonators.
There are also examples of musical instruments using double reeds attached to
straight bores, such as some chanters used in bagpipes, or ancient instruments such
as the crumhorn (see appendix A.1 for a listing of double reed instruments using double
reeds). Their sounds are usually very different from those of a clarinet, but they have
some significant differences from orchestral double-reeds: in general the lips do not di-
4 Chap. 1: The double-reed

rectly touch the reed. Such instruments tend to have a brighter and more harsh sound
than orchestral instruments, probably because of the damping and increased inertia
introduced by the lips.
It is also worth mentioning the development of single-reed mouthpieces for oboes and
bassoons. These were commercialised until a few years ago [Bate, 1975], but to our
knowledge are not fabricated any longer. However, one recent attempt by Nederveen 1
shows that they can be used to produce a sound almost indistinguishable from a real
double-reed. It is interesting to notice however that although these mouthpieces are
based on clarinet and saxophone ones, their dimensions approach those of oboe or
bassoon double-reeds, and this may give some insight into for further considerations
that will be made to the double-reed (sect. 2.4.1).
The facts listed above may suggest that only the resonator determines the timbre
of a reed instrument and not the kind of reed exciters used to play them. However
differences may exist in the coupling between the exciter and the resonator, which is
often more continuous in double-reed instruments.

1.1. Generic description of the double reed


Despite the variety of instruments using double-reeds as their exciters, their basic
configuration remains the same for all of them.

Figure 1.2.: Two double-reeds used in oboes (1) and bassoon (2).

The term ‘‘double-reed’’ usually refers to the device used to modulate the air-flow
entering a wind instrument such as the oboe or the bassoon. This device can be
assimilated to a valve, which by the action of pressure controls the flow.
The double-reed is traditionally constituted by two blades cut off natural cane (usually
of the species Arundo Donax). The two blades are bound together by their base forming
1
Although no publication was made on this device, several musical acoustics colleagues had the chance
to listen to and try it out at SMAC, 2003. Also, the commercialisation (by Runyon) of mouthpieces for
bassoons using clarinet single-reeds has been reported.
1.1 Generic description of the double reed 5

Staple
Cork

Butt
Oboe
Ligature
Blades
Rails
Tip

Wires
Bassoon

Figure 1.3.: Terms used to describe the regions of double-reeds


6 Chap. 1: The double-reed

a duct into which the air is driven by the difference of pressure between the mouth and
the duct. The duct starts as an oblong cross-section in the tip (in the upstream part of
the reed), progressively evolving into a quasi-circular cross-section towards the butt in
the downstream region. This end is introduced in the main bore of the instrument (see
fig. 1.3). The tip of a blase is thinner than the remaining part, and this makes it soft
and flexible.
In orchestral double-reeds, the musician gently presses its lips against the reed.
When compared to some ancient double-reed instruments, such as the shawm, the
crumhorn or popular ones such as some bagpipes, the latter have very bright, sometimes
strident sounds, whereas oboes and bassoons have fewer high-frequency components,
sounding mellower.

1.2. General working principles


The softer part of the reed (at the tips) can be forced to bend by the action of an external
pressure, imposed by the lips, or by a difference between the external and internal air
pressure. At a given value of the applied pressure the reed ends up closing, so that
ideally no flow should cross the reed.
In a regular musical use, there are periodical variations of the pressure inside the reed,
imposed by the resonator’s modes of vibration. These pressure oscillations drive the
position of each blade and the flow crossing it. In turn, these flow variations regenerate
the acoustic oscillations inside the instrument.

1.3. Fabrication of the double-reed


The fabrication of the double reed is a complex issue, and the details involved in the
preparation and tuning of the reed are the subject of a large number of works (see
for example [Kopp, 2003] [Smith, 1992]). Such details are out of the scope of this
document, and could be the subject of a complete new work involving the knowledge
of instrumentists and reed makers, their experience on the effects of different reed
scrapings and tunings of the reed, side to side with vibrational analysis of these scrapes,
and its effects on the intonation of the instrument, as an example.
To the subtleties of reed makers’ skills is added the fact that oboe and bassoon reeds
are played between the instrumentist’s lips. This changes the vibrational properties of
the reed, in a way that also depends on the instrumentist. An ideal scrapping to an
oboist may thus not be convenient to another oboist.
However, a quick glance at the procedure of fabrication can provide some general
ideas on the physics behind the functionning of the double reed. Several stages can be
identified in this procedure:
The double reed is usually fabricated from a single sample of cane, which is a cylin-
drical segment split into equal strips using a splitting arrow (fig. 1.4, 1). For an oboe
reed a dry tube of cane of 9.5 to 11 mm in diameter is split into 3 equal strips. For a
bassoon the initial diameter is 24-27 mm and it is split into 4 strips [Ponthot, 1988].
1.3 Fabrication of the double-reed 7

1
Pith

Bark
2

Figure 1.4.: Stages of reed making: 1 Raw tube of cane, with a strip cut off; 2 Folded
cane;3 Shaped cane placed on the staple; 4 Cane bound to the staple
8 Chap. 1: The double-reed

The cane is then shaved (gouged) from the inside (the pith), in order to remove part of
the soft matter, until the base thickness of the reed is achieved. Because the diameter of
the gouge can be different from the cane diameter, the basic cross-section of the gouged
strip consists of a crescent shape, uniform over the strip length [Yefchak, 2005]. The
thickness of the gouge dictates the reed hardness, while the diameter influences the
final opening of the reed [Goossens and Roxburgh, 1980].
After letting the reed soak for a few hours, an incision is made at half of the reed’s
length in the exterior part of the reed, breaking the hard fibers of the outer cane shell
(the bark).
The cane is then folded at the middle (2), and placed around a shaper, a hard steel
plate whose shape is roughly the final shape of each blade as seen from above. The
shaper is used to guide the trimming of the blades’ sides. The two blades are now
visible and present a tapering from the tip (in the middle of the strip) to the extremities.
They are still connected at the tip by the softer matter in the reed (3).
Depending on the instrument, the two blades are attached to each other by the butt
(in the bassoon) or tied to the staple (in the oboe). At this extremity they are forced to
increase their curvature, whereas at the tip, the junction between the two blades still
forces them to be straight one over the other.
The tip is then scraped to remove the bark, and cut at about 5 mm from the bending
(dashed line in 3). The two blades separate from each other at the tip. Depending on
the force attaching the two reeds together by the butt, the two blades are usually in
contact with each other along the sides (rails). This provokes an initial pre-constraint
on the reeds, and should assure air-tightness of the reed.
The final stage is to scrape the blades’ tips in order to give them a decreasing thickness
towards the tip of the reed. Usually the scraped part is limited to 7 to 9 mm near the
tip in the case of an oboe reed [Goossens and Roxburgh, 1980].

Lay

Heart Back

Figure 1.5.: Regions of thickness in an oboe blade (after


[Goossens and Roxburgh, 1980])

This is a very rough description of the scraping phase. Usually there is a thicker part
of the reed near the tip at the center called the ‘‘heart’’, and the corners of the reed (‘‘lay’’)
are thinner than the remaining tip. Some musicians prefer to leave a thicker ‘‘spine’’ in
1.4 Utilisation of the double-reed 9

the center of the reed, and they refer to this as a ‘‘W’’ scrape, and others prefer to scrape
more evenly the back of the reed (‘‘U’’ and ‘‘V’’ scrapes) [Goossens and Roxburgh, 1980].
However, the scraping technique and the final thickness profile of the reed depend
much on the musician’s practice. There are extensive works dedicated to the art of
making good reeds [Ledet, 1999] [Shalita, 2004], as an example. However, the perfect
technique does not exist, it depends on the physiognomy of the musician, on his playing
technique and on the learning path he followed.
As a final remark, [Goossens and Roxburgh, 1980] stresses the fact that symmetry
should remain a constant throughout the fabrication of the reed, so that for instance
the reed opening should present a symmetrical shape.

1.4. Utilisation of the double-reed

Due to the fragile nature of the double-reed, and the fact that it is used in straight
interaction with the musician’s mouth (in most modern instruments), the technique
used for playing a double-reed instrument plays a crucial role in the behavior of the
instrument, and the timbre of the notes produced.
The performance technique of both oboes and bassoons has been the object of a large
number of works dedicated to beginner musicians. These are usually intended as a
complement for the apprentice, because a close supervision of an experienced musician
is required to transmit an empirical knowledge that can be very difficult to translate
into words.
The following sections will give a quick glance on the general performance technique,
mainly as a review of what can be found in the musical literature, and intended to
provide some understanding on how the action of the musician can affect the physics
of the reed.

1.4.1. Soaking

Before playing the instrument the reed is usually soaked for a few (about 5 [Rader, 2004])
minutes in water so that the reed softens and its shape assumes a correct opening shape
(more closed than when dry for [Rader, 2004], however [Sprenkle, 1961] and others say
that an over-soaked reed has too large a tip opening).
Soaking also helps to seal the rails of the reed, so that the flow does not escape
from the sides while playing the instrument [Sprenkle, 1961]. Eventual leaks can be
corrected by wrapping the base of the reed in gold-beaters skin2 .
The accuracy of advised soaking timings (never beyond 5 minutes) suggests that in
fact the reed should not be saturated in water content in order to produce a good sound.
In fact, over-soaking is considered to open the tip too wide and making the reed
less responsive [Sprenkle, 1961] [Roscoe, ]. This has a double-sided effect of softening
2
the outer membrane of ox intestine. The advantages of this material are its strength and fairly uniform
thickness.
10 Chap. 1: The double-reed

the reed too much and forcing the musician to bite the reed too hard, vitiating his
embouchure.
After this initial soaking the water content is maintained by the natural humidity
provided by the musician’s breath. This is why an oboist or a bassoonist keeps the reed
in its mouth even during pauses.

1.4.2. Embouchure
Embouchure is understood to be the basic position of the mouth (lips, teeth and tongue)
that should be assumed to start playing the instrument. It should optimise the air
supply to the reed, and a correct position should allow the musician to control such
reed properties as the opening, its damping or inertia, as well as the amount of reed
that is left free to vibrate inside the mouth.
The contact between the lips and the reed should be air-tight (once again we see the
absence of leakages in and around the reed as a crucial factor).
The lips should roll over the teeth until the transition between red and white flesh
is in contact when the two jaws are brought together [Sprenkle, 1961]. This provides
a soft and damping material whose tension and hardness can be controlled via the lip
muscles (changing the biting should be avoided and jaws left at a constant position
[Goossens and Roxburgh, 1980]), which is referred to as ‘‘lip-pressure’’. Lip-pressure
can vary according to pitch and dynamics. [Campbell et al., 2004] states that some
players can hold the reed directly on their teeth in order to reach the highest notes,
which suggests that these notes require reduced damping while maintaining a strict
control over the reed opening.
The contact between the lips and the reed should be symmetrical in both blades
[Goossens and Roxburgh, 1980], so that the angle at which the instrument is held is
important.
A tip of 1 to 2 mm should be left inside the mouth out of lip contact [Goossens and Roxburgh, 1980],
[Campbell et al., 2004], so that at least a short tip of the reed is able to vibrate freely.
As a remark, we can suppose that the part of the reed which is in contact with the lips
can also vibrate, although its motion is expected to be reduced by the damping and
inertia of the lips. In fact, since the lips are not attached to the blades and they have no
counterpart inside the reed, they should be able to separate from the lips during part
of the motion induced by the reed tip.
Given that only the lower jaw can move with respect to the remaining skull, it is this
part of the face that can exert the greater control over the reed [Goossens and Roxburgh, 1980].
A special attention is thus paid to the lower jaw and lip. We can suppose that this re-
mark is mostly intended for the control of the lip-pressure, because the symmetry of
the reed implies an equally important contribution of both lips.

1.4.3. Articulation and dynamics


In this section we shall discuss briefly the recommendations on the gestures used during
intonation, both concerning transitions between notes, attacks of new notes from silence
1.4 Utilisation of the double-reed 11

and musical dynamics.


In general, it is advised to maintain a good and constant breath support (by keeping
the respiratory muscles under tension) even if the air pressure is to change inside the
respiratory tract [Goossens and Roxburgh, 1980]. This also ensures a good steadiness
of the force applied to the air even if the pressure is required to change in the mouth.
Pressure changes occur very frequently while playing the instrument, because the av-
erage flow varies depending on the regime, for example, flow can be important before
the oscillation starts but it decreases abruptly once the reeds starts to oscillate.
While attacking a note, the position of the jaws and the lip-force should not change.
The tongue is placed against the corner of the reed, preventing reed oscillations. Breath
is asserted before starting the note by removing the tongue from the reed [Sprenkle, 1961].
The tongue should block the air flow before the note is attacked [Goossens and Roxburgh, 1980].
We conclude that before the tongue is removed the air pressure in the mouth is thus
higher than once the note is about to start. Pressure should increase naturally as
the oscillation is established, if the breath support is maintained constant. This was
verified in analysis of the time-variations of the pressure during basic musical tasks
[Fuks and Sundberg, 1996].
During articulation (transition between to notes), the tongue touches the reed gently
but quickly, in order to stop the oscillation for a short interval. Breath support must
remain constant throughout this procedure.
Double-reeds are known to have a smaller dynamic range than other instruments
[Goossens and Roxburgh, 1980]. Similarly as what was said for articulation and at-
tacks, the breath support should be constant for different dynamics. The amplitude
of the reed oscillations (and thus the sound produced by the instrument) can be con-
trolled by the lip-pressure [Sprenkle, 1961]. A slight movement of the reed away from
the mouth is allowed [Goossens and Roxburgh, 1980] in order to reduce the vibrating
length of the reed during diminuendi.
12 Chap. 1: The double-reed
2. Reed models
This chapter is dedicated to describing the general principles of functioning of reed
instruments, and unveiling some clues about the particularities of double-reed instruments.
In particular, we review some propositions and studies existing in the scientific literature.

2.1. Introduction
Double-reed instruments share many features with other reed instruments, for in-
stance, the resonance of an air column, the pressure-controlled valve... It is thus
natural to start our study with a description of what is already known for reed instru-
ments, providing a basic mathematical model that can describe the generic behavior of
the instrument.
In the way, we indicate some key-points where this model is more susceptible to fail
for double-reeds. We will finish with an overview of some of the propositions already
found in the literature.

2.1.1. Overview of the chapter


The chapter starts with a general presentation on the physical approach to musical
instrument modelling, gradually focusing, first on the large group of auto-oscillating
instruments, then wind instruments and the more particular case of reed instruments.
Section 2.3 introduces the physical model of resonators, both cylindrical and conical.
Section 2.4 focuses on the part of the instrument that will be the central subject of
this thesis: the exciter.
Finally, section 2.5 reviews the present-state of knowledge on double-reeds.

2.2. Self-sustained instruments


In a general way, sound production in musical instruments can be seen as a way of
taking advantage of the natural oscillations of an object, or several objects in contact.
Most objects can naturally vibrate according to different modes, depending on the
method that is used to excite these vibrations.
Reed instruments belong to a large group of instruments where oscillations are self-
sustained. This means that the natural vibrations of the resonator (the air column
present inside the bore of the instrument) are maintained throughout the playing of one
note by the internal mechanisms of the instrument and a continuous energy supply.
The sound produced by self-sustained instruments is characterized by a stable output
of sound whose intensity can be maintained as long as the musician wishes.
14 Chap. 2: Reed models

2.2.1. Passive and active parts of the instrument


In a classical approach, it is usual to distinguish a passive part of the instrument, called
the resonator, which has its modes of vibration, and receives the energy supply from the
exciter to set these modes into vibration and eventually maintain them through time (in
the case of self-sustained instruments).
This formal separation proves to be useful because the resonator can be described
separately from the exciter and the two systems linked through a set of common vari-
ables.
Usually, the two systems depend on each other, so that the time evolution of the
state of the instrument can only be found by solving the two systems together. Only
in some special cases, which do not concern our study, can the state of the exciter be
found independently, providing an input signal which is transformed (filtered) by the
resonator.

2.2.2. Wind instruments


In instruments that concern this thesis’ study, the resonator is assimilated to the air
column contained inside the instrument. It is limited by the walls of the bore, roughly
a cylindrical or conical pipe. The length of the air column is usually much larger
than the transverse dimensions (the diameter), a fact that simplifies the mathematical
description of the resonator as will be seen below (sect. 2.3)
When the instrument is sounded, some of the properties of the air-column oscillate
around their values at rest. Pressure (p) and flow velocity (u) are the usual quantities
that are used to describe the vibration of the air, but other properties periodically os-
cillate and transmit these oscillations to the neighboring regions of air. These are, for
example the density (ρ) and temperature (T ) of the air. In general, however, the oscilla-
tions of these variables inside the air column are much stronger than those present in
the propagation of sound in free-air.
The modes of oscillation of these properties in the air-column determine roughly the
frequency at which the instrument will sound, although the mechanical properties of
the reed can also influence it to a lesser extent [Brown, 1990].
Several methods exist to create and maintain the oscillations of the air-column. Ex-
citation can be described as a set of conditions imposed on the oscillating properties of
the air-column in a determined region.

2.2.3. Reed instruments


In reed and lip-driven instruments the excitation mechanism is a solid component that
periodically blocks or limits the airflow into the instrument (a valve). The main difference
between these two kinds of valves is that lips are blown-open (or outward-striking: an
increase in the mouth pressure forces the valve to open) whereas reeds are blown-closed
(or inward-striking).
The next sections describe the general principles of functioning of a reed instrument,
and how these can be transported into a physical model of the instrument. The usual
2.3 The resonator 15

conceptual separation between exciter and resonator is used.


In practice, the resonator is the region where the air oscillations are sufficiently small
to be described by the equations of acoustic propagation (a linearised version of the
general equations of fluid motion for small variations of pressure, flow velocity, density
and temperature around average values of these variables). Eventually, there may be
some cases (for loud playing levels) where non-linearities may be important, because
the variations of the referred variables may become too strong to be described by a
linear model. Such non-linearities were shown to be important in brass instruments
[Vergez, 2000], but recent studies [Gilbert et al., 2005] prove that in the case of wood-
winds these non-linearities have no perceptible effects.
The exciter is the part of the instrument where the approximations used in acoustics
are not valid, whether it be due to the variations in the boundary conditions (movement
of the reed) or to massive variations of the flows. This includes the vibrating reed and
the acceleration of the flow into the instrument.
It must be kept in mind that this separation between exciter and resonator is a con-
ceptual one, which simplifies the mathematical description of the instrument. However,
in the case of wind instruments, the air-column is also the continuation of the air flow
involved in the excitation. This can make the separation between both systems delicate
and somewhat artificial, requiring special care in the interpretation of the results of
experiments and models when the exciter or the resonator are studied independently.

2.3. The resonator


As stated above, the study of the resonator is based on the linear propagation of acous-
tic waves. The deduction of the acoustic wave equations from the equations of fluid
motion can be found in any introductory acoustics book [Fletcher and Rossing, 1991],
[Morse and Ingard, 1968] and will not be reproduced here. Our study of the resonator
will start with simple solutions of this wave equation for the propagation of planar waves
in a given direction.

2.3.1. Wave equation and wave propagation

The solutions to the linear acoustic pressure wave equation (for plane waves) are a linear
combination of the following:

p+ (z, t ) = A(ω)e (ωt −kz ) (2.1)


− (ωt +kz )
p (z, t ) = B(ω)e (2.2)

In this equation, z is the direction of propagation of the wave (perpendicular to the


wavefronts), ω represents the angular frequency measured in rad/s (omega = 2πf ,
where f is the frequency of the wave). k is the wave number which in ideal propagation
is related to the angular frequency by the relation c = ωk , c representing the velocity of
sound. It is found from the resolution of the linearized Navier-Stokes equation and the
16 Chap. 2: Reed models

adiabatic compression relation:


r
γpatm
c= (2.3)
ρ

where γ = Cp /Cv = 1.4 is the ratio of constant pressure and constant volume specific
heats, patm is the average atmospheric pressure and ρ is the density of the air.
In equation (2.1), p+ represents a wave travelling from left to right (that is, in the
positive direction of z)1 .
The propagation of a pressure wave (p ± (t )) implies a propagating flow velocity wave
±
(u (t )), which follows the pressure wave with a phase offset.

2.3.2. Propagation inside a cylindrical waveguide


Inside a duct, the solid walls impose boundary conditions on the flow of the acoustic
wave. In particular for an ideal fluid, the velocity component perpendicular to the wall
has to be zero. In the case of a cylindrical duct, and in fact of any duct with straight,
parallel walls (constant cross-section), one possibility for fulfilling these conditions is
that the wave propagates parallel to the walls. Wave fronts are thus perpendicular to
the walls. Since the acoustic wave is a longitudinal wave, the flow will thus always be
parallel to the walls.
Of course this is not the only possibility. Even if the propagation is perpendicular to
the walls, standing waves can exist that have q nodes (no oscillation of the flow velocity)
at the reed walls.
For the study of wind instruments we shall restrict ourselves to waves that propagate
longitudinally (along the axis of the bore). Due to the dimensions of the bore, transverse
standing waves tend to correspond to frequencies much higher than usual playing
frequencies of the instruments.
In a waveguide, the acoustic flow is defined as q(z, t ) = S(z )u (z, t ), where u (z, t ) is
the acoustic velocity, supposed to be homogeneous over a cross-section (S(z )) of the
waveguide at position z (usually defined to start at one of the resonator’s ends, as long
as only the axial modes are considered. Acoustic flow can then be related to the pressure
waves according to the formula:

1
q± (z, t ) = ± p± (2.4)
Zc (z )

with
ρc
Zc = (2.5)
S
which is the so-called characteristic impedance of the acoustic wave, constant through-
out the resonator length because its cross section S is constant.
A wave of frequency ω propagates along the resonator with a velocity of c = ωk , but
as we’ll see below (section 2.3.3) the wavenumber k may need to be considered as
1
This can be verified by following a constant phase (φ = ωt − kz): constant phase implies that dφ = 0, so
that ωdt − kdz = 0 and ∂z ∂t
= + ωk .
2.3 The resonator 17

x
r (z )
z

Figure 2.1.: Cylindrical resonator and associated coordinates and variables

depending on the frequency. The propagation can thus be described by the following
formula:
P ± (ω, x ) = e ∓ık (x −x0 ) P ± (ω, x0 ) (2.6)

2.3.3. Diffusion and dispersion of the waves

In the previous section we described mathematically the propagation of an ideal wave.


Equations (2.1) state that the amplitude of the wave is maintained through time and
space, and the same is true for the acoustic flow.
In practice, in a free field, some of the energy of the wave is lost to a temperature
increase through viscosity. When confined to a cylindrical waveguide, to this energy
dissipation must be added the dissipation through viscous effects due to the no-slip
condition on the boundaries (friction against the resonator walls) and thermal conduc-
tion through the resonator walls which implies that compressions and expansions are
not really adiabatic. Contributions from losses through the walls are much more im-
portant than through viscosity and conduction through the air, so that the latter will
be neglected.
For large resonator diameters (when compared to the boundary layer width), these
effects can be introduced as perturbation (Kirchhoff’s theory [Rayleigh, 1945]), so that
the propagating waves can still be described by equations (2.1) and (2.4) with a variable
wave number:
ω i 3/2
k (ω) = − ηcω1/2 (2.7)
c 2
with
√ √
   
2 cp
η= lv + −1 lt (2.8)
Rc 3/2 cv

and R is the bore radius, lt and lv the thermal and viscous characteristic lengths, and
cp /cv the ratio of the constant volume to constant pressure specific heats.
18 Chap. 2: Reed models

2.3.4. Resonator termination and radiation

When a travelling wave arrives at one end of the resonator, it does not follow unchanged
in the same direction. A part of the wave is reflected back to the resonator (in the
opposite direction) and the rest is transmitted to the free acoustic field outside of the
resonator. The relationship between the reflected and the incident wave can be written
in frequential form using the reflection coefficient (Rrad ):

P + (ω) = Rrad (ω)P − (ω) (2.9)

A particularly simple case of this situation exists when a boundary condition is im-
posed at the end which states that the pressure field outside of the resonator is constant
(corresponding to the atmospheric pressure). In first approximation this is true, consid-
ering that the pressure fields inside the resonator are much higher than outside. In this
case, the total acoustic pressure at the end of the resonator is zero, so that 0 = P + + P − ,
which means that the outgoing wave (p − ) is reflected back symmetrically (P + = −P − ).
The key to describing the part of the wave which is radiated is to know the charac-
teristic impedance of the radiated wave exactly at the resonator ending (transition from
resonator to the free field). This radiated wave is assimilated to a wave produced by
a plane circular piston with the same cross section as the resonator. In fact, there is
no analytical solution to this problem with the particular boundary conditions of the
cylindrical resonator (which are boundaries for the free field as well as for the internal
field), but two particular solutions exist for the flanged and unflanged planar piston.
For the flanged planar piston, the load impedance seen by the planar wave arriving
at the termination is written [Levine and Schwinger, 1948]:

ρc
Zrad = (Rr (2ka ) + ıXr (2ka )) (2.10)
S
with

2J1 (x )
Rr (x ) = 1− (2.11)
x
2H1 (x )
Xr (x ) = (2.12)
x

and J1 (x ) the first order Bessel function of the first kind, and H1 (x ) the first order
Struve function. S denotes the cross-section area of the tube at the exit, and k = ω/c
the wavenumber of the sinusoidal wave.
These models provide a radiation impedance Zrad (ω) which can be used to calcu-
late the reflection and transmission coefficients seen by the outgoing wave (p − ) at the
extremity of the bore:
Zrad /Zc − 1
Rrad = (2.13)
Zrad /Zc + 1
where Zc is the characteristic impedance of the planar wave inside the bore.
2.3 The resonator 19

Reflection Coeff.
0.8

0.6

0.4

0.2

0
Frequency HHzL
100 1000 10000

Figure 2.2.: Reflection coefficient corresponding to a flanged pipe termination (absolute


value)

2.3.5. A complete cylindrical resonator

These three key effects (acoustic propagation, visco-thermal losses and radiation) are
essential to describe a complete resonator model. We are now able to calculate the
input impedance of an ideal cylindrical resonator (without visco-thermal losses and
ideal end reflection) to a more realistic resonator, including losses and radiation of a
flanged piston.

160
150
Impedance HdBL

140
130
120
110
100
90

Frequency HHzL
0 1000 2000 3000 4000 5000

Figure 2.3.: Cylindrical resonator impedances. In dashed red, the simplest model con-
sisting of a perfect reflection at the bore’s end, and no propagation losses.
In solid blue, a flanged pipe termination is supposed, and visco-thermal
losses through the walls are taken into account
20 Chap. 2: Reed models

2.3.6. Conical resonator models


Most modern double reed instruments such as oboes or bassoons use resonators whose
shapes are conical throughout most of their length. Although these are slightly more
complicated than the resonators described in section 8.2.1, an exact description using
spherical waves is possible (neglecting visco-thermal losses, non-linearities and non-
axial modes).

x
θ r (z )
z

Figure 2.4.: Coordinates and variables used on conical resonator description

If we consider conical walls as boundary conditions of the wave equation, it can be


written simply in spherical coordinates. If we only consider variations in the radial direc-
tion (along r, which is a similar restriction as the restriction to longitudinal propagation
seen in 2.3.2), it can be written as:

1 ∂2 p
 
1 ∂ ∂p
r2 = (2.14)
r 2 ∂r ∂r c 2 ∂t 2
whose solutions are:
rp(r, t ) = P (r, t ) = P + (r, t ) + P − (r, t ) (2.15)

and

A(ω) (ωt −kr )


p+ (r, t ) = e (2.16)
r
B(ω) (ωt +kr )
p− (r, t ) = e (2.17)
r

The acoustic flow waves q ± (r, t ) are related to the pressure waves p ± (r, t ) by the
following formulas:

1
q+ (z, t ) = p+ (z, t ) (2.18)
Zc (ω, z )
1
q− (z, t ) = − p− (z, t ) (2.19)
Z ∗ (ω, z )
c
2.3 The resonator 21

ρc 1
Zc (ω, z ) = 1 (2.20)
S 1 + ıkz

These are the spherical travelling waves which will be used to describe the conical
bore.

Diffusion and dispersion

The inclusion of viscosity and thermal diffusion terms in the spherical wave equation
greatly complicates its resolution. While for the cylinder the analytical solutions can
be written (section 2.3.3), for the cone no exact solution has been found yet, to our
knowledge.
However, if we consider that the tapering angle of the resonator is small, the plane-
wave Kirchhoff theory should be a good approximation to the total losses by using an
value of R which stands between the input and the output radii in equations (2.7) and
(2.8). The best approximation for this value is the radius at one third of the length from
the smaller radius.

Resonator termination and radiation

For conical resonators, and in a general way, for a resonator which terminates with a
bell, the plane piston theory of section 2.3.4 is no longer exact, because the wavefronts
on the edge of the bell are curved. In a conical waveguide, the flanged piston case turns
into a pulsating dome (circular portion of the sphere) oscillating in the remaining rigid
sphere.
Thomas Helie’s theory [Helie and Rodet, 2003] provides radiation impedances for spher-
ical wave terminations compatible with the one-dimensional descriptions of wind instru-
ment resonators. These impedances (Zrad ) can be used to calculate the reflected and
radiated waves in a similar way as given by equation (2.13), but since we are now work-
ing with spherical waves, the appropriate characteristic impedances have to be used,
giving:

Zrad /Zc − 1
Rrad = (2.21)
Zrad /Zc∗ + 1

Impedance of a truncated cone

In practice, a conical resonator needs to be truncated at its apex, because a non-zero


input cross-section is required for the coupling to the exciter, or to a consecutive bore
segment. Figure 2.5 shows the impedance of a truncated cone.
In a truncated cone, impedance maxima are not harmonically related. For this reason,
an instrument composed of a conical resonator associated to a reed is unstable, and
usually some alterations are made to its conical shape to bring the impedance peaks
back into a harmonic relation.
22 Chap. 2: Reed models

180
170

Impedance HdBL
160
150
140
130
120

Frequency HHzL
0 1000 2000 3000 4000 5000

Figure 2.5.: Impedances of a truncated conical resonator. In dashed red, the simplest
model consisting of a perfect reflection at the bore’s end, and no propagation
losses. In solid blue, a flanged pipe termination is supposed, and visco-
thermal losses through the walls are taken into account

In real instruments, in addition to the geometrical corrections to the bore there are
further aspects that balance this inharmonicity. In particular for reed instruments, the
cavity inside the mouthpiece, the reed damping and the flow induced by the reed motion
(‘‘pumped flow’’) are found to provide the necessary correction for the overall instrument
impedance so that resonance frequencies are harmonically related [Nederveen, 1998].
In a mathematical model this correction can be introduced by adding a small cylin-
drical section between the exciter and the conical resonator [Scavone, 2002].

2.3.7. Complex geometries


In general, the resonator of a wind instrument has a more complex cross-section profile
than that presented in the previous sections. This profile can be described by a function
S(x ), the cross-section area as a function of the position along the resonator (x).
One way to implement an arbitrary cross-section is to divide the whole resonator into
segments where it can be approximated by a conical or cylindrical profile. These are
then coupled together by assimilating the acoustic variables (p and q) in the extremities
in contact between two segments. This is the approach used in our synthesis models
(see sect. 8.2.3 for details).
A different approach consists in considering that the wavefront is approximately
spherical throughout the resonator, and writing the wave equation as a function of the
resonator profile S(x ). This kind of wave propagation description is called the Webster
equation.
Recently, it was shown [Helie, 2003] that the spherical wavefront approximation can
be dropped, and the wavefront shape can be determined simultaneously to the wave
equation, providing better results for the calculation of input impedances or reflection
functions, for instance.
2.4 The exciter 23

2.4. The exciter


The main object of this thesis remains the double-reed used as an exciter for double-reed
instruments (such as oboes and bassoons).
There is no obvious way to distinguish the exciter from the rest of the instrument.
This should be regarded as a conceptual distinction, so that it is not always possible
to associate a physical part of the instrument to the exciter or to the resonator alone,
but it may be possible to distinguish some processes that are clearly related to the
creation and maintenance of the oscillation (needed for all self-sustained instruments,
as described in section 2.2).

2.4.1. Double reeds as opposed to single reeds

The name ‘‘double-reed’’ suggests that the number of oscillating reeds has an important
influence in the behavior of the exciter.
However, the strong coupling between both blades in double-reeds, imply that their
behavior is in fact closer to that of a single vibrating solid than to two independent
vibrating structures. In fact, this will be the subject of some observations shown in
section 3.3.4 for example, where we give experimental evidence of the joint behavior of
the two blades that compose the double reed.
A distinction has to be made between the object traditionally associated with the
double-reed (because it is detachable from the remaining instrument as a single piece)
and the exciter as functionally independent from the resonator. In fact, blowing into an
oboe double-reed for instance reveals the acoustical influence of the duct downstream
of the vibrating reeds, allowing us to classify this part of the duct as a resonator. For
example, shortening the staple has obvious consequences on the frequency of vibration
of the reed, and the sound produced by it, very much like the consequences of changing
the length of the main bore of an oboe, for example.

2.4.2. Description of the double reed

At the end of the fabrication of the reed, the tip is open showing an oblong eye shape.
This opening is due to the natural curvature of the cane from which the reed was made
and to the pre-constraint induced by tying the reeds together at the butt extremity, but
the radius of curvature of each blade is usually less pronounced than that of the original
cane, because of the force of contact between the sides of the blade.
The elasticity of the cane allows the reed to close when the blades are forced against
each other. This can be done either by pressing the fingers against the reed, or by
applying a pressure difference between the outside (pm ) and the inside of the reed (pr ).
This behavior is similar to that of a spring, whose displacement from the position at
rest (x0 ) increases with the force applied to it. The simplest way of describing the rela-
tion between the force and the displacement is to suppose a linear and instantaneous
relation between both quantities. This model, which will be tested in section 3.4, can
be written in the form of a simple equation:
24 Chap. 2: Reed models

pm − pr = kS (x − xO ) (2.22)

where the stiffness ks is given per unit surface. Given its surface-averaged value, ks can
be defined from the traditional stiffness k used in the study of a spring by dividing it by
the effective surface Seff over which the pressure is exerted:

k
ks = (2.23)
Seff

and
F
Seff = (2.24)
pm − p r
where F is the total force applied on the reed.
Equation (2.24) provides a way of determining Seff , if necessary. In practice we will
see that the whole equivalent oscillator can be described in terms of surface averaged
parameters, such as ks .
A difference in pressure between the outside and the inside causes the reed to change
its opening. If this difference is sufficiently high, the two blades will bend until they are
in contact with each other throughout the whole length of the reed opening.
The velocity of the flow entering the reed also depends on the pressure difference
between the inside and the outside of the reed. Using the simplest possible model, one
can say that while entering the reed, the total energy of the air is conserved, leading to
the Bernoulli condition:

1 2 1
pm + ρum = pr + ρur2 (2.25)
2 2
The volume flow entering the reed can be found by supposing that the velocity is
homogeneous over a cross-section of the flow. In the reed we would have qr = Sr ur and
in the mouth qm = Sm um . If the flow is uncompressible (which is evaluated in section
6.2.2) both volume flows are identical, because of the condition of mass conservation.
Since the flow is distributed over a much bigger surface inside the mouth than at the
reed entrance, the flow velocity is usually neglected (um ' 0).
The flow law (after replacing the velocity ur by the corresponding flow expression Sqr )
from equation (2.25) can be combined with the elasticity law from equation (2.22) to
obtain a curve relating the flow to the pressure difference applied to the reed (fig. 2.6),
and described by the equation:
s
pM − (∆p)r 2(∆p)r
q= (2.26)
ks ρ

Rearranging the parameters in equation (2.26) it is possible to derive a simpler non-


dimensional formulation for the same model:

q̃ = (1 − p̃)p̃1/2 (2.27)
2.4 The exciter 25

PT PM

0.25

(l/s) 0.2

0.15
Flow

0.1

0.05

0
0 5 10 15 20
Pressure (kPa)

Figure 2.6.: A typical non-linear characteristics curve for a reed of dimensions similar
to an oboe reed

with
r
ks ρ
q̃ = 3/2
q (2.28)
pM 2
p̃ = (∆p)r /pM (2.29)

This formula shows that the shape of the non-linear characteristic curve of the elemen-
tary model is independent of the reed and blowing parameters, although the curve is
scaled along the pressure p and volume flow q axis by the stiffness ks and the beating
pressure pM = kS0

2.4.3. Dynamic effects of the reed


In a normal utilisation of the double-reed – as an exciter of a wind instrument – the reed
system is far from a static case.
In section 2.4 we have stated the simplest model possible for the reed elasticity.
Beyond the fact that the relation between pressure and reed opening might not be
linear, there are several other factors that are neglected in this expression in what
concerns the vibration of the reed without the coupling to the flow.
Due to the inertia and damping, the reed can be seen as a harmonic oscillator driven
by the external forces, which are imposed by the difference between the nearly static
mouth pressure applied by the musician pm 2 and the time-varying pressure inside the
reed pr (t ).
If we consider the external forces to be distributed along an effective surface S eff ,
the harmonic oscillator can be described in terms of an effective mass per unit surface
2
In fact time variations of the mouth pressure can be considered and can explain variations in timbre and
attack times which musicians can control by varying the shape of their vocal tract [Fritz, 2004]
26 Chap. 2: Reed models

m r
ms = Seff
and damping rs = Seff
:

∂x ∂2 x
pr (t ) − pm = kS (x (t ) − xO ) + rs + ms (2.30)
∂t ∂t 2
This equation can be somewhat simplified if we consider only the variations of x (t )
relative to its equilibrium position x0 , written as ∆x (t ) = x (t ) − x0 , and the pressure
difference between the inside and the outside of the reed ∆p(t ) = pr (t ) − pm . Moreover,
the whole equation can be divided by the surface mass ms :

∆p(t ) ∂(∆x ) ∂2 (∆x )


= ωr2 ∆x (t ) + 2α + (2.31)
ms ∂t ∂t 2
Equation (2.31) should replace equation (2.22) as a simplified mechanical description
of the reed.
Eventually, the reed will have to be studied and modelled as a vibrating solid body
with its own modes and frequencies of vibration. A preliminary study is presented in
appendix B, but higher modes of vibration will not be considered in the present work.

2.4.4. Considerations about the elementary model


For the sake of simplicity, several factors have been disregarded in the former analysis,
and we will try to discuss some of them in detail, giving an idea on whether or not
they may be important in the description of the instrument, and why. As much as
possible, experimental data accompaigns the analysis, trying to justify the importance
or unimportance. In other cases, when experiments were not possible with the avail-
able equipment, possible hypothesis are advanced, based on theoretical or dimensional
analysis.

2.5. State-of-the-art in double-reed physical modelling


Reed instruments have been the subject of numerous studies in the past, and they still
arise a continuing interest, mainly because of the relative simplicity of their modeling.
In fact, for example the form of the non-linearity present in reed exciters is much
simpler than that of bowed strings, a family included in the wide group of self-sustained
instruments.
Present studies vary from analysis of the dynamic system constituted by the reed and
the resonator [Kergomard, 1995], period-doublings and route to chaos [Maganza et al., 1986]
[Lizée, 2004], arbitrary resonator geometries for the resonator [Ducasse, 2001] [Helie, 2003],
the effect of conical geometries [Scavone, 1997b] or the inclusion of non-linear effects
in the propagation inside the resonator [Gilbert et al., 2005]. The particular aspect of
the propagation in the vocal tract was also included in classical models [Fritz, 2004] in
order to test its importance.
In the particular domain of single-reed instruments, the exciter model has been re-
fined for example with respect to the mechanical properties of the reed such as the visco-
elastic properties of the material [Casadonte, 1995] [Obataya and Norimoto, 1999], the
2.5 State-of-the-art in double-reed physical modelling 27

vibrational properties of the reed [Pinard et al., 2003] or the effect of the particular
geometry of the table which supports the reed [Gazengel, 1994]. Flow details were
also addressed, investigating the effects of the formation of a jet at the reed input
[Hirschberg et al., 1990].
However, the details of the exciter seem to have arisen a greater interest in single-
reeds than for the particular case of double-reeds. One of the reasons for this condition
is without doubt the relative difficulty of carrying out experiments and measurements
in the relatively tight double-reed.
A certain number of articles can be found about double-reed instruments though, on
some particular aspects.
The details of the bores and the reed used in double-reed instruments such as the
oboe or the bassoon is discussed and compared to other woodwinds in [Nederveen, 1998].
Experimental measurements of the parameters of the double-reed exciter can be found
for example in [Almeida et al., 2004b], with a first attempt to obtain the non-linear
characteristics. Aspects connected to the performance of the oboe were approached by
[Fuks, 1998].
Other aspects include the particular oscillation cycle of several double-reed instru-
ments, which is very asymmetric when compared to that of the clarinet [Gokhshtein, 1981]
[Rocaboy, 1989] [Agulló, 2001]. This is a particularity of conical resonators, more than
of the double-reed itself, and of the fact that in these kind of resonators the acoustical
cycle is determined by a single symmetric reflection at the end of the resonator rather
than a sequence of 3 reflections (at the end, at the top and again at the end) in cylindrical
resonators [Rocaboy, 1989].
These authors agree that the fraction of the period during which the reed remains
open is determined by the acoustical propagation in the resonator (variable according
to the playing frequency) and that the time of closure is independent of the playing fre-
quency. However, the authors diverge as to what determines this time. For some, it is
the frequency of vibration of the volume of air trapped inside the reed [Rocaboy, 1989].
Others consider that it is the truncation length of the cone [Dalmont et al., 2002]. How-
ever, none of these justifications explain the increase in the closure time as the mouth
pressure is increased [Gokhshtein, 1979].
[Brown, 1990] uses Rocaboy’s theory to explain the change in tuning that can be
achieved by changing the embouchure parameters. According to him, these have a
primary effect on the value in the pressure wave at which the transitions between open
and closed regimes is observed.
The existence of two symmetric movements of the reed motion was demonstrated for
the simplified Helmholtz motion in conical resonators [Dalmont et al., 2002], inverting
the durations of open and closed states.
Moreover, the existence of other regimes of vibration (in particular with more than
one closed period per cycle and quasi-periodic regimes) was observed in the bassoon
[Shimizu et al., 1989]. In this article, the authors analyse the parameters that cause
transitions between regimes.
For the Tenora, a complete physical model has been proposed and simulated in the
time-domain [Barjau and Agulló, 1989]. The reed model remains similar to the classical
28 Chap. 2: Reed models

model presented in section 2.4, but the relation between the reed position and the
opening area is changed from linear to quadratic. The simulated pressure waves are
compared to waves measured in a real instrument.
The reed motion, and its relation to the pressure signals inside the reed is analysed
in [Gokhshtein, 1979]. The same author proposes a model for the pressure distribution
inside the reed, and discusses the effects of changing the Vena Contracta at the reed
input by rounding off the reed’s edges [Gokhshtein, 1981].
Lately, the flow model inside the reed was the subject of some studies [Wijnands and Hirschberg, 19
proposing that a constriction downstream of the reed increases the resistance seen
by the flow. Although this model was proposed for some particular kinds of reeds
[Hirschberg, 1995], it has been since then adopted as a model for double-reeds in the
musical acoustics literature [Fletcher and Rossing, 1998]. An investigation on the ef-
fects of the hysteresis inherent to this model on simulated instruments is given in
[Vergez et al., 2003].
Part II.

Characterisation of the double-reed


3. Geometrical and Mechanical aspects of the
reed
In this chapter we introduce some basic experimental data on the mechanical properties
of the double reed, and the shape of the reed opening. Firstly, the geometry of the
reed opening is studied as a function of pressures applied to it. Then, we address the
elastic properties of the reed, including viscoelastic effects in the reed material. Finally,
the vibrating properties of the reed are analysed through a scanning of its mechanical
response.

3.1. Introduction
The double-reed is a complicated system involving two main components that influence
each other: the elastic reed blades, which can be deformed by external forces, and the
air flowing between them [Hirschberg, 1995]. These two components can be modelled
independently, and the interaction between both established through a set of coupling
variables.
The aim of this chapter is to gather some experimental informations about the solid,
elastic reed, as much as possible as an independent system from the air flow that runs
through it in normal utilisation. A first investigation concerns the variations in the
reed opening shape for typical pressure variations expected in a musical use, and the
following experiments are dedicated to the elastic and vibrational properties of the reed.
Most of the experiments described in this chapter try to minimise the effects of the
flow. However, for measurements of the evolution of the reed shape we compared static
to dynamic regimes.

3.1.1. Overview of the chapter


The chapter begins with measurements of the cross-section profile along the reed (sect.
3.2).
In section 3.3 we study the variations in the shape of the reed opening for a range
of pressures found in normal uses of the reed. The main aim is to identify a single
variable that can describe the reed opening state in a one-to-one relation, during normal
operating regimes. In particular, a model for the double-reed requires the knowledge of
the relation between the opening area and the displacement of the reed.
Section 3.4 is dedicated to measuring and discussing the relation between the pres-
sure applied to the reed and the variation in the reed opening area, allowing to propose
a spring model for the reed.
32 Chap. 3: Geometrical and Mechanical aspects of the reed

Section 3.5 discusses the non-instantaneous character of the previous relation, trying
to link it to the viscoelasticity in the reed material.
Finally, section B.1 investigates the vibrational properties of the reed for small driving
forces and negligible induced flow effects.

3.2. Cross-section profile of the double-reed

Along this work, it will be necessary to know the internal geometry of the reed. This
information is important for instance when discussing the evolution of the flow inside
the reed (chapter 5).
The internal geometry of the reed was measured indirectly, by filling it with a casting
material after having applied a de-moulding liquid, so that the cast would not stick
to the reed. For the casts, several materials were tried out and finally, two revealed
successful: one solidified completely into a hard mould requiring the destruction of the
reed, while another more supple material (Rhodorsil) required only that the reed blades
were unmounted.
For the digitalisation of the cast surface, an automatic topography method was tried
out (described in appendix J). The development of this method would require further
investment in time and material, so that finally a semi-manual method was used to
extract some key-points in the cast surface (see appendix J.1).
The profiles extracted for a synthetic oboe double-reed are presented in figure 3.1. For
natural cane-reeds, it is more difficult to assure that the material used for the casting
does not deform the reed blades.
The cross-section area is seen to increase constantly from the reed tip till the reed
output, with a slight discontinuity at the staple input.
In the upstream part of the reed, bounded by the blades, the width of the reed
decreases from the tip till the staple input, whereas the height increases in the same
direction.

3.3. Geometry of the reed opening

In this section, we investigate the shape of the reed entrance for different opening states
of the reed. The reed entrance, often called reed opening in this document, is the space
between the tip of the two blades, through which the air flows in a normal utilisation of
the instrument.
Among other observations, we examine whether there is a characteristic shape of the
reed opening for a particular opening area, and consequently we can describe the reed
opening state through a single variable.
These investigations are firstly driven by the interest of having a simple model of the
reed input, so that ideally a single variable is sufficient to describe the shape of the
reed opening. Its area can therefore be related to this variable, and eventually the flow
entering the reed can also be linked to the same variable.
3.3 Geometry of the reed opening 33

8
height
width
6
dimension (mm)

0
0 10 20 30 40 50 60 70
axial distance (mm)

15

10
area (mm2)

0
0 10 20 30 40 50 60 70

Figure 3.1.: Measurements of the cross-section profile on an oboe double reed


34 Chap. 3: Geometrical and Mechanical aspects of the reed

There is still a more general interest in characterising the reed opening shape. Musi-
cians and reed-makers usually have an idea of the shape the reed should have in order
to guide the preparation of the reed. In addition to this idea of the reed opening shape
at rest, they test the reed opening shape by pressing the reed between the fingers and
observing the variation of the shape.
In literature dedicated to musicians and reed makers, some information can be found
about the shape of the reed. No consensus seems to exist about an ‘‘ideal’’ shape of
the reed opening. This subject, like many other issues on the reed making, seems to
be greatly dependent on the kind of instrument, the formation of the musician, and the
musician’s preferences as well.
As an example, two modes of transition between closed and open states are shown in
figure 3.2. These can be found in some works about reed manufacture [Smith, 1992].

h
h

1 – proportional 2 – diamond-shaped

Figure 3.2.: Transition between open and closed states of the reed opening for two dif-
ferent designs (according to Heinrich [Heinrich, 1979])

Still, some of the indications left by musicians in these works were transported into
scientific works, for example in order to propose physical models of double-reed instru-
ments [Barjau and Agulló, 1989]. In scientific works it is common to read [Nederveen, 1998]
that double-reeds differ intrinsically from single-reeds in the geometrical model of the
flow input. In particular, it is proposed that while closing, both dimensions (width
and height) of the reed opening are reduced, implying a 2nd order power law for the
dependence of the area upon the reed height (S(h ) ∝ h 2 ).

3.3.1. Experimental approaches

Because the reed is a flexible solid, the deformation of the reed blades depends on the
distribution of the force over the reed surfaces.
In a common utilisation of the reed, there is a constant pressure applied by the lips
over a region of the reed. When blowing the reed, the pressure increases in a the part of
the reed that remains inside of the mouth (fig. 3.3, 1). Moreover, when a note is played,
between the two reed blades there are periodical variations of the pressure.
When testing the reed during its preparation, the force is distributed differently over
the blades (fig. 3.3, 2).
Any of these ‘‘usual’’ distributions of pressure can be difficult to reproduce exactly in
experiments, so that more homogeneous distributions were used in practice.
3.3 Geometry of the reed opening 35

pm Lip force

Reed pressure

1 – Musical performance 2 – Preparation

pm
pm

Reed pressure

3 – experimental, non-oscillating 4 – experimental, oscillating

Figure 3.3.: Some common pressure distributions used in the double-reed, on top for
common uses of the reed by musicians and on the bottom for our experi-
ments.

Some measurements and observations were tried out using artificial lips in the arti-
ficial mouth. For most experiments however, the pressure was evenly distributed along
the reed external surface and left free to adapt itself in the internal surface (fig. 3.3, 4),
and the results did not vary much from the results shown in the following sections.
In fact, the force acting on the reed is expected to have a greater influence on the
reed tip than on the remaining reed for two main reasons: the reed blades are thinner
there, and since the reed can be seen as a cantilever beam fixated on the staple edge,
the moment of the force increases towards the reed tip.

Operating Regimes

For the analysis of the reed opening shape, it was judged important to compare some
of the observations for non-oscillating (static) and oscillating (dynamic) regimes, so
that some conclusions can be applied later to the movement of the double reed, and
eventually to a physical model of the double reed.

Static regimes One way to access the whole range of reed openings in a statice regime
is to prevent the reed oscillations by covering the input section of the reed with a plastic
film (fig. 3.3, 3), blocking the airflow. The reed is thus submitted to a pressure difference
which can be adjusted easily by increasing the blowing.
One other way is to use a diaphragm at the reed output, which was done in the
measurements of chapter 4. Similar observations can be made without the application
36 Chap. 3: Geometrical and Mechanical aspects of the reed

of a plastic film, however, flow effects are more important in this case.
In fact, the plastic covering can influence the results in two particular aspects:

• Increasing the actual force on the reed, because the surface upon which the pres-
sure acts is slightly increased. However the difference can be considered small (the
surface covering the reed opening is small when compared to the blades’ area).

• Changing the distribution of forces on the reed, which could influence the way the
reed closes. Nevertheless, observations of the reed geometry while closing with
and without the plastic film covering (in a non-oscillating case) were compared to
check that the geometry is basically the same.

Because a light source placed in front of the reed would produce specular reflections
on the plastic covering, the experiment is lit from behind. The reed opening thus
appears as a bright region on a dark background (left-hand side of fig. 3.4), which has
the additional advantage of producing an image with better contrast, and an easier
identification during the subsequent image analysis (see sect. 3.3.3).

Dynamic regimes For dynamic observations, the artificial mouth is also used to generate
the pressure that drives the reed into self-sustained vibration, but now the input section
of the reed is left uncovered.
The reed is illuminated with a stroboscopic light, synchronised with the period of the
sound wave produced by the instrument (or by the reed alone), as described in section
G.2.

Reeds used in this study

The results presented in this section refer to natural cane oboe double-reeds, used as
bought from Glottin (no further scraping applied).
However, these are basic analysis that can be performed on any reed, so that most
results apply and were observed on other kinds of oboe double-reeds. Substantial
differences observed in certain reeds are mentioned when needed.

3.3.2. Observations
Figure 3.4 shows frontal views of the reed opening using two different methods of ap-
plying the pressure. In sequence 1, a plastic film covers the reed so that no air flow
can enter the reed. The reed does not oscillate during observations. In sequence 2, the
photographs were captured while the reed was oscillating, with the help of a stroboscope
synchronised with the sound produced by the instrument.
These observations show that the contact between the blades is almost simultaneous
along the reed opening length, in both static and dynamic regimes (see figure 3.4). In
fact, as the reed opening closes, its shape changes as if it was scaled along its minor
axis, emulating the behavior shown in the left of figure 3.2.
3.3 Geometry of the reed opening 37

1 – Non-oscillating 2 – Oscillating (strobe)

Figure 3.4.: Observations of the reed opening at different closing states. 1 – Variations
are slow; 2 – different phases of a high frequency oscillation
38 Chap. 3: Geometrical and Mechanical aspects of the reed

Due to the utilisation of back-lighting, the left-hand side of figure 3.4 can highlight
small openings that remain even for high mouth pressures, which are difficult to observe
with front lighting. Moreover, a slight reduction of the reed opening length is observed
in the start of this series, which is probably linked to the fact that the tip of some natural
cane reeds are not in contact in the edges when no pressure is applied to them. In fact,
a regular use of the reeds and the consequent humidification tends to bring the edges
together at rest.
After the first stages of the reed closure, the reed length remains practically constant,
a different behavior from the model proposed in [Barjau and Agulló, 1989].
The reed shape does not vary considerably from the shape of the reed opening at rest,
suggesting that one single coordinate describing the distance between the two blades
is sufficient to parameter the shape of the reed at any time, for example the central
distance between blades h or the reed opening area itself S.
One implication of this observation is that no transverse modes of vibration of the
reed are observed. In fact, a transverse mode would imply that the shape of the reed
opening changes throughout one period, with a higher frequency.

3.3.3. Quantitative analysis

Although the observations above already provide interesting information about the reed
opening geometry, methods described in appendix H allow us to determine the function
relating the reed coordinate z (the maximum distance between the two blades at the reed
tip) to the corresponding opening cross-section S(z ), and quantify the non-linearity in
the geometric relationship S(z ). These measurements can usually be done on the same
images as those used in qualitative observations.
Figure 3.5 represents the reed opening area as a function of h, for different openings.
A rectangular area in which only one of the dimensions is altered (like the clarinet)
would produce a linear distribution, with the other dimension as slope. On the other
hand, a variation like the one represented on the right hand of figure 3.2, would produce
an almost quadratic distribution, because both dimensions vary at the same time when
reeds close.
For the double-reed, the obtained distribution (figure 3.5) is mostly linear, which
confirms the direct observations described in the beginning of this section.
The slope of the graph corresponds to the effective length of an equivalent rectangular
reed opening (with the same area), which is smaller than the length measured on a real
reed. The coefficient between effective and real lengths is a parameter to be used in
mathematical models of the double-reed, since it controls the magnitude of the volume
flow that enters the instrument for a given distance between the blades of the reed.
Note that all the reeds (both natural and synthetic) showed a linear relationship be-
tween their opening area and width. However, certain reed makers claim that depending
on the reed, both cases depicted in figure 3.2 are possible. In practice, should the be-
havior shown on the right part of figure 3.2 be observed, an exponent (usually between
1 and 2) would be included as a parameter for the fitting to the data in figure 3.5.
3.3 Geometry of the reed opening 39

−6
x 10
4

3.5

2.5
Area (m )
2

1.5

0.5 Area=5.183492e−03*width+−6.253252e−08

0
0 1 2 3 4 5 6 7
Width (m) −4
x 10
−6
x 10
2.5

1.5
Area (m )
2

0.5 Area=5.354350e−03*width+−1.044430e−07

0
0 1 2 3 4 5
Width (m) −4
x 10

Figure 3.5.: Opening area measurements on a natural cane reed (a, on top – without
lips and b, with artificial lips).
40 Chap. 3: Geometrical and Mechanical aspects of the reed

3.3.4. Symmetry

In oboe-like instruments, the two reed blades control the air flow into the instrument.
One fundamental question is to know if the two blades oscillate synchronously and
symmetrically in order to check whether both reeds can be considered as oscillators
with identical mechanical parameters.
With the help of the stroboscope, periodic reed motion can be inspected by eye. For
relatively low frequencies (compared to the reed natural frequency), the two blades os-
cillate in phase, and their motion looks symmetric at first sight (see figure 3.4). This
conclusion stands for many recorded sequences of images and was also observed di-
rectly at playing frequencies non-multiple of the camera recording rate, for which video
recording is difficult. We propose to draw quantitative information from the image anal-
ysis.
First, the main-axis of the reed is determined as the greatest straight line that can
be traced inside the reed opening region. The sub-area on each side of the reed main
axis are then calculated. The left sub-area and the opposite of the right sub-area are
displayed in figure 3.6 against time (case of a plastic reed). It is obvious that for that
particular reed, both blades oscillate synchronously and symmetrically. This results
stands for other plastic reeds. However, in some natural cane reeds, asymmetry can be
noted: the two blades motion, while synchronous, sometimes have different amplitudes.
This indicates an ill-constructed reed, as mentioned in section 1.3. It is frequent in fact
to observe that the two blades have slightly different curvatures at rest.
Also, for other regimes (quasi-periodic, multiphonics or squeaks) symmetry is not
always observed. However, these regimes are only used in special occasions, or not
used at all in musical interpretation. Note that all observations were made in steady
states, so that transients have been disregarded.

3.3.5. Conclusions on the reed opening shape

The results shown in this section indicate that for a normal musical use it is accurate
enough to describe the double-reed as a system with one degree of freedom. This does
not necessarily mean that both blades are identical, but that the coupling between both
is sufficiently strong to make them work as a single oscillator.
Even in cases where the double-reed showed a clear asymmetry (results not shown in
this section), the movement of both blades was observed to be simultaneous. However,
as mentioned in section 1.3, such asymmetric reed-openings should be considered to
be ill-designed or unfinished.
None of the observed reeds has shown a tendency for the quadratic law relating the
distance between reeds (h) and the reed opening are S(h ), which was proposed by
[Barjau and Agulló, 1989] for the Tenora (however Tenora reeds were not studied by
us).
Such quadratic laws for S(h ) are also implicit in some designs proposed by reed
makers and musicians (fig. 3.2). As described above, these observed closing modes
of the reed can be due to the force distribution applied for testing during the design
3.4 Reed equivalent stiffness 41

−6
x 10
1.5
right
left
1
Opening area (m2)

0.5

−0.5

−1

−1.5
0 2 4 6 8 10 12
Time (s)

Figure 3.6.: Time evolution of the half areas of the slit (on each side of the reed opening
axis)

(fig. 3.3). Otherwise, the fact that such closing modes were not observed by us may
be justified by the relative scarceness of finished reeds in the samples tested by us,
revealing a lack of variety in the analysed reeds.

3.4. Reed equivalent stiffness


When modeling the double reed as a harmonic oscillator, the quasi-static regime cor-
responds to an instantaneous relation between the reed displacement (z) and the force
applied (F ), because time derivatives implied in inertia and damping can be neglected.
In such a model, the force is distributed over the reed surface, so that it is more straight-
forward to measure a surface stiffness k as a ratio between the reed displacement and
the pressure applied k = pz .
For this measurement, a synthetic reed is inserted in the artificial mouth, being
subjected to a controlled pressure from its outside. Stiffness is estimated through the
instantaneous and static relation between the pressure and the displacement, so that
oscillations have to be prevented in order to keep inertia and damping negligible. We
use the static method as described in section 3.3.1 to prevent the oscillations.
Pressure is increased progressively until the reed closes (points A to B in fig. 3.7).
When the reed is almost closed, pressure can be increased with very small consequences
on the reed opening (hook on the upper right end in figure 3.7, corresponding to points
42 Chap. 3: Geometrical and Mechanical aspects of the reed

Reed area (mm −− difference from equilibrium)


3
Measured
Fit
2.5 C
B

1.5

1
2

0.5
D
0
A

−0.5
−5 0 5 10 15 20
Pressure (kPa)

Figure 3.7.: Measurement of the applied pressure as a function of the reed displacement
from its rest position, and fitted linear function to the decreasing pressure
branch (points C to D).

B to C): it is in fact very difficult that the measured area reaches zero because light
tends to diffuse when the reed is almost closed.
After the reed was kept closed for a few seconds, the same measurement is performed
while decreasing the pressure (points C to D in figures 3.7 and 3.8). The reed opening
area decreases at a rate similar to that found while increasing the pressure, but the
curve shows an offset towards slightly smaller reed openings. In fact when the pressure
is released the reed opening has a larger area than in the beginning of the experiment.
A likely interpretation is that the reed blades are slightly deformed by the continued
force applied to the reed, suggesting that a more complicated model should be used to
explain this behavior (see sect. 3.5).
Besides, the relation between the opening area and the pressure applied on the reed
is mostly linear, which is quite unexpected for such a complex system. Given that the
slope of the graphic does not vary much between increasing and decreasing pressures,
stiffness can be calculated as this slope (neglecting the offset). In figure 3.7 it is found
to be approximately 6.15 × 109 Pa · m −2 for decreasing pressures. This value can be com-
pared with that found for the clarinet, about 0.8 × 109 Pa · m −2 , based on measurements
by S. Ollivier and J.-P. Dalmont [Dalmont et al., 2003], which means that the opening
area is more sensible to pressure variations in clarinet reeds than in oboe reeds. This
may account to some extent to the results obtained for the comparison of oboe and
clarinet non-linear characteristics (sect. 4.3.8).
The measurements presented in this section were obtained for a synthetic reed. The
3.5 Viscoelastic effects 43

20 4
pressure
reed displacement
15 C 3
B
Pressure (kPa)

opening (mm )
2
10 2

5 1

0 0
A D

−5 −1
0 10 20 30 40 50 60
Time (s)

Figure 3.8.: An example of time variations of pressure and reed displacement during a
stiffness measurement.

linearity of the elasticity curve observed in figure 3.7 is not perfectly maintained in
natural cane reeds. In fact, the stiffness is seen to increase for small reed openings in
dry reeds. Wetting the reed seems to bring the curve back to a linear elastic regime (see
appendix F)

3.5. Viscoelastic effects


The reed deformation suggested by the observations in the previous section can be taken
into account by means of a model of viscoelasticity, a phenomenon frequently observed
in material science.
A simple model can be used to describe the reed deformation according to the time
evolution of the force (expressed as the convolution with a kernel function, for example a
decreasing exponential) [Christensen, 2003]. Such a model states that when a constant
force is applied to the material, its deformation increases with time, tending to a constant
value that is proportional to force. Similarly, after this force is suddenly released, the
deformation is still observed, but it tends back to zero as time evolves in a way that can
be modeled considering an exponential kernel as:

P
x0 = e −t/τ (3.1)
kve
The parameters of the model are then the relaxation time τ, and the viscous stiffness
kve . P is the pressure step that forced the reed.
44 Chap. 3: Geometrical and Mechanical aspects of the reed

2.7

Reed Opening(mm )
Measured

2
Exponential sum fit
2.6
2.7

2.5 2.6

2.5
2.4
2.4
0 10 20 30
2.3
0 200 400 600 800 1000
Time (s)

Figure 3.9.: Measurement of the viscoelastic relaxation time and fitted sum of two
exponentials (inset shows a zoom for the short-term relaxation).

3.5.1. Synthetic reed

In order to estimate the parameters τ and kve , a constant pressure is applied to the reed
in the artificial mouth (28 kPa), keeping it shut for 30 minutes. The pressure is then
suddenly reduced to 0, so that the reed should re-open to its rest position. The reed
opening is recorded on video for several minutes after relaxing the pressure.
Image analysis is used to measure the reed opening area (vertical axis in figure 3.9)
along the time (horizontal axis). Measurements are affected by a certain amount of
noise, because the opening variations are small when compared to the absolute reed
opening, but the overal trend of the reed opening can be distinctively recognized. The
source of the noise is mainly the grain of the image (fluctuations in the intensity of each
pixel from one image to the other).
When observing the data for the full length of the measurement (about 15 minutes)
the opening area is seen to increase slowly at a decelerating rate. A measurement
realized several hours after the acquisition in figure 3.9 shows that the opening area
stabilized at 2.83 mm 2 . This long-term evolution can be modelled using equation (3.1)
with parameters τ = 850 s and kve = 150 × 109 Pa · m −2 .
However, this model fails to describe the evolution of the reed opening area in the
short-term (figure 3.9, inset), because there is a quicker transition during the first
few seconds after the pressure release. This relaxation period can be described by
adding another function in the form of equation (3.1) with parameters τ = 2.5 s and
kve = 150 × 109 Pa · m −2 . This fact is also common in the study of complex materials,
3.5 Viscoelastic effects 45

where several relaxations are observed with different time constants, corresponding to
different scales in the material structure.
Note that the similarity of the values found for kve in both relaxations is merely a
coincidence and was not imposed in the analysis of the results.
In clarinet reeds, a similar behavior is observed [Dalmont et al., 2003], with a long-
term relaxation close to 900 s, and a short-term relaxation of longer duration than for
double reeds (about 8 s).
Both relaxation times are several orders of magnitude larger than the period of oscil-
lation of the reed (usually smaller than 10−2 seconds), so that viscoelastic effects can be
neglected for a simple model which describes the instrument behavior (for sound syn-
thesis, for instance). In practice, viscoelastic effects tend to slightly close the reed after
some time of playing, because there is an average force over the reed applied during
large intervals of time.
This effect is probably balanced by the musician, who slightly releases the force of the
lips over the reed, but its effects are observed on experiments with the artificial mouth:
when playing during long intervals, changes in frequency and intensity are observed
as if the reed parameters were changing. In fact, if the reed is not stable enough and
visco-elastic effects cause strong variations in the reed opening at rest, the musician
periodically presses the reed by its rails in order to re-increase the opening area.
As another example, during the experiments carried out in section 3.4, the closing
and opening of the reed lasted for one minute in average. The short-term relaxation can
contribute to the difference in opening area values before and after the reed is forced for
some seconds (see figure 3.7).

3.5.2. Humidified natural cane reed

The previous experiment was not repeated for a humidified cane reed, however, the
relaxation of the reed can also be studied after a different time-distribution of the stress.
In figure 3.10 are presented the reed-opening variations after a 3-minute measurement
typical run for measuring the reed characteristics, as done in chapter 4.
There are two main differences from the previous analysis with respect to the time
evolution of the force applied to the reed. Firstly, the stress is not constant before
the release of the pressure, and second, the time during which the stress is applied is
smaller than the typical values of the slowest relaxation times found previously. This
means that some ‘‘memory’’ of the state of the reed before the experiment is conserved
throughout the whole run.
The relaxation time can be measured on the graph, considering that the relaxation
begins at 165 seconds verify in MatLab, and that the reed opening stabilises at the same
value as observed around 45 seconds, the reed opening reaches 60% of its final (e −1/2 )
value at 180 seconds, yielding a relaxation time (τ) of 30 seconds. The differences of the
relaxation times found in this graphic when compared to that of figure 3.9 are probably
related to different material used for the reed.
46 Chap. 3: Geometrical and Mechanical aspects of the reed

4 50 0.4

3.5 40 4 0.35
30

Pressure difference (kPa)


3 0.3
20 2

opening (mm )
2
2.5 10 0.25
0 0
2 0 100 200 0.2

1.5 0.15

1 0.1

0.5 Pressure 0.05


Opening
0 0
110 120 130 140 150 160 170 180 190 200
Time (s)

Figure 3.10.: Zoom on the time evolution of the reed opening for low displacements of
the reed relative to its rest position.

3.6. Partial conclusions


In this chapter we analysed some important details for a model of the double reed.
The internal cross-section profile was measured, showing a divergent duct throughout
most of the length of the reed. These measurements will be useful in the remaining of
this work.
It was demonstrated that the reed opening area S(z ) varies linearly with the distance
between the blades z, and during the oscillations of the reed no higher transverse modes
were observed.
When applying a pressure difference across a dry natural cane reed without flow,
the reed opening area varies non-linearly according to the applied pressure, showing a
higher stiffness when the reed is almost closed. For synthetic reeds and natural soaked
reeds a mostly linear relation is observed, allowing to attribute a stiffness constant to
the reed. These measurements unveiled a viscoelastic behavior of the reed.
By completely releasing a pressure applied during a sufficiently long period of time
and observing the opening area variations after the release, we were able to measure
the relaxation times of the reed material and the amplitude of the viscoelastic effects.
These are long-term effects that can be seen as a gradual shift of the opening area at
equilibrium (S0 ).
These results will be used in the interpretation of further experiments realised through-
out the remaining of this work, and provide important information for the tuning of the
model described in section 2.4.
4. The nonlinear characteristics of the reed
In this chapter we try to characterise the double-reed in static regimes. The non-
linear relation between the pressure drop ∆p in the double-reed and the volume flow
crossing it q is measured for slow variations of these variables. The volume flow is
determined from the pressure drop in a diaphragm, a technique used in the past for
similar measurements in clarinet mouthpieces [Dalmont et al., 2003]. The influence of
environment and experimental conditions on this relation is verified for double-reeds,
and the measurements are compared to other reed instrument exciters and to physical
models.

4.1. Introduction
4.1.1. Context
It is a common assumption that a self-sustained instrument model can be divided into
a resonator and an exciter block [MacIntire et al., 1983]. The two blocks are connected
through a set of coupling variables. In section 2.2.3 a minimal set was identified as the
pressure p and volume flow q at the input of the resonator.
The physics of resonators was extensively studied in the past, with increasing re-
finement. . In musical instruments, the resonator can be described with sufficient
accuracy using a one-dimensional distribution of pressures p and volume flows q along
it. In turn, the values of these variables at the bore entrance are well characterised by
a linear, frequency-dependent relation (see sect. 2.3).
From the reed side, a relation between p and q is also required to describe the behavior
of the complete instrument. The coupled solution of the two relations determines the
time evolution of variables p and q and of the complete instrument. This relation is
called the non-linear characteristic of the exciter. In fact, whereas we know that in
the bore a linear relation models reasonably well its behavior, in the case of the reed
this is obviously not true. In fact, a negative slope in the p/q relation is required to
maintain auto-oscillations of the instrument, and a non-linearity to limit the growth of
the oscillations [Maganza, 1985]. It was previously shown that even for a simple model
of a reed exciter (sect. 2.4) this relation is highly nonlinear.

Static characteristics Due to dynamic properties of the reed (for instance, its inertia and
damping, see sect. B.1) and of the flow (again, the inertia of the flow) the non-linear
relation between p and q is likely to be non-instantaneous .
In this chapter however, we will try to bypass this problem, and restrict the study
of the reed to slow variations of the coupling variables. By slow it is understood that
48 Chap. 4: The nonlinear characteristics of the reed

the time-scales of these variations are larger than the typical delays involved in the
relations between the two variables p and q in the reed. Measured in such conditions,
the relation will be called the quasi-static non-linear characteristic.
This restriction is imposed by the difficulty of measuring quickly varying volume flows,
and the complexity of a measurement which would take into account a great range of
time-scales. However, a great amount of information can be extracted from the quasi-
static regime, and it can be shown that an instantaneous relation is a good approxima-
tion to the actual non-linear characteristics when the playing frequency of the instru-
ment is low when compared to the resonance frequency of the reed [Wilson and S., 1974]
[Dalmont et al., 1995].

Double-reed quasi-static characteristics The nonlinear characteristics curve measured on


a double-reed can be compared to that of single-reed instruments. It can be checked
whether the nonlinear behavior of the reed in low-frequency regimes can be described by
a generic model such as described in section 2.4 and successfully used for single-reed
instruments, or if some particularities of the flow inside the reed imply a modification of
the generic model used for single-reeds, as described in chapter 2, and suggested based
in theoretical considerations about the flow inside the reed [Hirschberg, 1995].

4.2. Principles of measurement and practical issues


The characteristic curve requires the synchronised measurement of two quantities: the
pressure drop across the reed (∆p)r and the induced volume flow q.

4.2.1. Volume flow measurements

One of the main difficulties in the measurement of the reed characteristics lies in the
measurement of the volume flow. There are instruments which can accurately measure
the flow velocity in an isolated point (LDA, hot-wire probes) or in a region of a plane (PIV),
but it can be difficult to calculate the corresponding flow by integrating the velocity field.
In fact it is difficult to do a sampling of a complete cross-section of the reed because
a large number of points would have to be registered. Since the flow is found to be
approximately axisymmetric at the reed output (see sect. 5.6.3), the measurement along
a diameter of the reed would be sufficient, but regions close to the wall are inaccessible.
Another difficulty would be making sure that the profile is being measured along a
diameter.
On the other hand, commercial flow-meters considered for these experiments have
the disadvantage of requiring a direct reading, which would have been unpractical for a
complete characteristic measurement, since it requires a large number of reading in a
short time interval.
An indirect way of measuring the flow was then preferred to the above mentioned
methods. It consists in introducing a flow resistance in series with the reed, for which
the pressure can be accurately related to the flow running through it (see fig. 4.1).
4.2 Principles of measurement and practical issues 49

The diaphragm method, used successfully by S. Ollivier [Ollivier, 2002] to measure


the non-linear characteristic of single reeds, is based on this principle. The resistance
is simply a perforated metal disk which covers the reed output.

(∆p)s

PSfrag replacements (∆p)r  (∆p)d


  
  
pm p  patm
  
r


1  2

Figure 4.1.: Use of a diaphragm to measure flow and pressure difference in the reed.
Numbered rectangles correspond to the pressure probes used in the mea-
surement.

For such a resistance, and assuming laminar, viscous less flow, the pressure drop
(∆p)d = pr − patm across the diaphragm can be approximated by the Bernoulli law1 :
 2
1 q
(∆p)d = pr − patm = ρ (4.1)
2 Sdiaph

where q is the flow crossing the diaphragm, Sdiaph the cross section of the hole, and ρ
the density of air. In our experiment, pressure patm is the pressure downstream of the
diaphragm (usually the atmospheric pressure, because the flow opens directly into free
air). The volume flow q is then determined using a single pressure measurement p r .

4.2.2. Practical issues and solutions


Issues

The realisation of the characteristic measurement experiments encountered two main


problems:

Diaphragm reduces the range of (∆p)r for which the measurement is possible The ad-
dition of a resistance to the air flow circuit of the reed changes the overall nonlin-
ear characteristic of the reed plus diaphragm system (corresponding to (∆p) s and to
the dashed line in fig. 4.2). The solid line plots the flow against (∆p)r , the pressure
drop needed to plot the non-linear characteristics. When the resistance is increased,
the maximum value of the system’s characteristic is displaced towards high pressures
1
The flow velocity at the reed output is neglected when compared to the velocity inside the diaphragm
(Sdiaph  Soutput )
50 Chap. 4: The nonlinear characteristics of the reed

[Wijnands and Hirschberg, 1995], whereas the static beating pressure2 value does not
change (because when the reed closes there is no flow and the pressure drop in the
diaphragm ((∆p)d ) is zero).

Q (l/s)
0.3

0.2

0.1

10 20 30 40 50 60 p (kPa)

−0.1

Figure 4.2.: Comparison of the theoretical reed characteristics (solid line) with the reed
plus diaphragm system characteristics (dashed) — mathematical models,
based on the Bernoulli theorem [Wijnands and Hirschberg, 1995].

Therefore, if the diaphragm is too small (i. e., the resistance is too high), part of the
decreasing region of the system’s characteristics becomes vertical, or even multivalued,
so that there is a quick transition between two distant flow values, preventing the
measurement of the characteristic curve (see figure 4.2).

Reed auto-oscillations The restriction of the measurements to low-frequency regimes is


a simplification in terms of the amount of measured data because it is not necessary to
measure a different curve for each frequency, and because simpler methods can be used
to measure the flow in quasi-static conditions. Nevertheless, this restriction requires
that the pressure and flow vary smoothly throughout the whole range of measurement.
This can be difficult to achieve in practice, because of auto-oscillations, sought in a
normal utilisation of the instrument, but that have to be prevented here using artificial
procedures.
Reed auto-oscillations arise when the reed ceases to behave as a passive resistance
(a positive ∂p
∂q
, which absorbs energy from the standing wave inside the reed channel)
to become an active supply of energy ( ∂p ∂q
<0). As seen in section 2.3, all real acoustic
resonators are slightly resistive (the input admittance Yin has a positive real part). This
can compensate in part the negative resistance of the reed in its active region, but only
below a threshold pressure, where the slope of the characteristic curve is smaller than
the real part of Yin for the resonator.

2
Static beating pressure: The minimum pressure for which the reed channel is closed
[Dalmont et al., 2005]
4.2 Principles of measurement and practical issues 51

Solutions proposed to adress these issues

Size of the diaphragm The volume flow is determined from the pressure drop across the
diaphragm placed downstream of the reed. In practice there’s a tradeoff that determines
the ideal size of the diaphragm. If it is too wide, the pressure drop is too small to be
measured accurately, and reed oscillations likely to occur. If it is too small, the system-
wide characteristic can become too steep, making part of the ((∆p)r ) range inaccessible.
The ideal diaphragm cross-section is then found empirically, by trying out several
resistance values until one complete measurement can be done without oscillations or
sudden closings of the reed.
In practice, the experiment is first performed with a device replacing the diaphragm,
whose resistance can be adjusted continuously until the system behaves conveniently.
One possibility for this replacement, which we used, is a medical flow regulator used in
intravenous injections. Once an appropriate resistance is found, the replacement device
is calibrated using the method described below for the diaphragms, and an appropriate
diaphragm is sought whose characteristics approach that of the replacement device
using the data in figure 4.6.

Finer control of the mouth pressure pm During the attempts to find an optimal diaphragm,
it was found that sudden closures were correlated to sudden increases in the mouth
pressure. A part of the problem is that the mouth pressure is not controlled directly.
This control is done using a pressure reducer between the main compressed air source
and the experimental apparatus. The pressure in the mouth thus depends both on the
valve position and on the downstream resistance.
By introducing a leak between the pressure reducer and the experimental apparatus,
it is possible to have smaller pressure increases in the mouth while the reed is closing.
This improves measurements in the decreasing region of the characteristic when the
system-wide characteristic is hysteretic. The addition of a leak does not influence the
experiment, because it is placed upstream of the artificial mouth.

Increase the reed mass One other way to reduce the oscillations is thus to prevent the
appearance of instabilities, or to reduce their effects. An increase in the reed damping
would certainly be a good method of avoiding oscillations, because it cancels out the
active role of the reed (which can be seen as a negative damping) [Debut, 2004].
It is difficult to increase the damping of the reed without altering its opening or
stiffness properties. The simplest way found to prevent reed oscillations was thus an
increase in the reed mass.
This mass increase was implemented by attaching small masses of Blu-Tack 3 to
one or both blades of the reed. During measurements on previously soaked reeds it
was difficult to keep the masses attached to the reed, so that an additional portion
of Blu-Tack is used to connect the to masses together, wrapping around the reed. A

3
A plastic adherent material usually used to fixate paper to a wall
52 Chap. 4: The nonlinear characteristics of the reed

verification of the consequences of these added masses on the static properties of the
reed is investigated later in this chapter.

Figure 4.3.: Front view of the reed with attached masses, at left in dry conditions, at
right in soaked conditions (to prevent the masses from slipping).

4.2.3. Experimental set-up and calibrations

The experimental device is shown in figure 4.4. The artificial mouth described in ap-
pendix G.1 was used as an artificial blowing mechanism and support for the reed. The
window in front of the reed allows the capture of frontal pictures of the reed opening.
Artificial lips, allowing to adjust the initial opening area of the reed were not used
here, fearing that they would modify some of the elastic properties of the reed, yet
differently from what happens with real lips.

pm pr

Lens
Camera Reed Controllable
leak

Diaphragm

Manometer Compressed
air source
PSfrag replacements Artificial
mouth Humidifier

Figure 4.4.: Device used for characteristics measurements.

As stated before, the plot of the characteristic curve requires two coordinated mea-
surements: the pressure difference (∆p)r across the reed and the induced volume flow
q, determined from the pressure drop (∆p)d across a calibrated diaphragm (sect. 4.2.1).
In practice thus, the experiment requires two pressure measurements pm and pr , as
shown in figure 4.1.
4.2 Principles of measurement and practical issues 53

Pressure measurements

The pressure is measured in the mouth and in the reed using Honeywell SCX series,
silicon-membrane differential pressure sensors whose range is from −50 to 50 kPa.
These sensors are not mounted directly on the measurement points, one terminal of
each sensor is connected to the measurement point using a short flexible tube (about
20 cm in length). Therefore, one tube opens in the inside wall of the artificial mouth, 4
cm upstream from the reed, and the other tube crosses the rubber socket attaching the
diaphragm to the reed output. The use of these tubes does not influence the measured
pressures as long as their variations are slow.
The signal from these sensors is amplified before entering the digital acquisition card.
The gain is adjusted for each type of reed. The system consisting of the sensor connected
to the amplifier is calibrated as a whole in order to find the voltage at the amplifier output
corresponding to each pressure difference in the probe terminals: the stable pressure
drop applied to the probe is also measured using a digital manometer connected to
the same volumes, and compared to the probe tension read using a digital voltmeter.
Voltage is found to vary linearly with the applied pressure within the measuring range
of the sensor.

Diaphragm calibration

The curve relating volume flow q to the pressure difference through diaphragms (∆p )d
can be approximated by the Bernoulli theorem, because they are constructed so as to
minimize friction effects (by reducing the thickness of the diaphragm) and jet contraction
— the upstream edges are smoothed by chamfering at 45◦ (fig. 4.5). The chamfer height
(c) is approximately 0.5 mm. This dimension is not very precise, because the chamfers


are hand-made. The diaphragm channel is 3 mm long (L).
 

 


Flow 
 
   d
c
  
  
direction

Figure 4.5.: Detail of the diaphragm dimensions.

Nevertheless, this ideal characteristic is checked for each diaphragm (see fig 4.6). It
was found that the effective cross section is slightly smaller than the actual cross section
(about 10%), which is probably due to a Vena Contracta effect in the entrance of the
diaphragm (see appendix C.3). Moreover, above a given pressure drop the volume flow
increase is lower than what is predicted by Bernoulli’s theorem (corresponding to a lower
exponent than 1/2 predicted by Bernoulli). This difference is probably due to turbulence
54 Chap. 4: The nonlinear characteristics of the reed

1
10
1.0 mm (exp)
1.3 mm (exp)
1.8 mm (exp)
1.0 mm (theor)

Volume Flow (q) in l/s


0
10 1.3 mm (theor)
1.8 mm (theor)
Rec limit

−1
10

−2
10 0 1 2
10 10 10
Pressure drop (∆ p)d in kPa

Figure 4.6.: Calibration of diaphragms used in characteristic measurements (dots are


experimental data and lines are Bernoulli predictions using the measured
diaphragm diameters). The dashed black line represents the pressure/flow
relation corresponding to the expected transition between laminar and tur-
bulent flows (Rec = 2000), parameterised by the diaphragm diameter d.

generated for high Reynolds Numbers. In figure 4.6 the dashed line corresponding to
the critical value of the Reynolds number (Rec = udν
= 2000) is shown. It is calculated
using the following formulas for u (the average flow velocity in the diaphragm) and d
(the diaphragm diameter):
 1/2
2q
d = (4.2)
πu
 1/2
2(∆p)d
u = (4.3)
ρ
so that the constant Reynolds relation is given by:
 π 1/2  ρ 1/4
Q 1/2 ∆p1/4 = Rec ν (4.4)
2 2
where the right-hand side should be a constant based on the diaphragm geometry.
Since a suitable model was not found for this data, we chose to interpolate the ex-
perimental calibrations in order to find the flow corresponding to each pressure drop in
the diaphragm. Linear interpolation was used in the (p, q 2 ) space.

Typical run

Two different kinds of measurements can be distinguished:

Simple measurement Only the pressures pm and pr are measured, so that in the end only
the pressure / flow characteristic is available.
4.2 Principles of measurement and practical issues 55

Measurement including opening data The reed opening is recorded in video at the same
time as the pressure measurements. This allows further inspection of the cou-
pling by plotting the flow vs pressure characteristic as well as stiffness and other
measurements related to the reed opening.

Simple measurement This kind of measurement is usually performed in a first stage,


when trying to find the suitable experimental conditions for a certain kind of reed, in
particular the appropriate size of the diaphragm and the amount of mass to add to the
reed in order to avoid oscillations.
When both the pressure sensors and the diaphragm are calibrated, a run consists in
increasing the pressure from 0 until slightly above the pressure at which the reed closes.
The pressure is then decreased back to 0. A typical run lasts for about 3 minutes, and
is depicted in figure 4.7.

30
pm
25 pr

20
pressure (kPa)

15

10

0
PSfrag replacements
−5
0 50 100 150 200
time (s)

Figure 4.7.: Time variation of the mouth pressure (pm ) and the pressure inside the reed
(pr ) during a successful characteristic measurement.

Measurement including opening data A more complete characterization of the reed is


achieved by recording and identifying the reed opening at the same time as the pressure
measurements.
The main part of the measurement is similar to the one described above, but be-
fore this measurement actually begins, the pressure is forced to increase and decrease
slightly 2 or 3 times for a short period of time, inducing a simultaneous reduction of
the reed opening area. This procedure is necessary to obtain data synchronized in time,
56 Chap. 4: The nonlinear characteristics of the reed

because the image rate of the video (25/sec) is different from that of the pressure mea-
surements (usually 4000/sec) and both recordings start at different times. An example
of such a measurement is shown in figure 4.8.

anche4h_18_2masseslourdes_nonosc14.BIN
50
reed pressure
mouth pressure
40 opening area 4e−06

30 3e−06
pressure (kPa)

20 2e−06

10 1e−06

0 0

−10 −1e−06
0 50 100 150 200 250
time (s)

Figure 4.8.: Time variation of the mouth pressure (pm ), the pressure inside the reed
(pr ) and the opening area (S) of the reed during a successful characteristic
measurement.

These pressure peaks are supposed to be simultaneous to the reed displacement


peaks, because they are slow enough when compared to the natural period of oscillation
of the reed (see section 2.4.3), and fast enough when compared to the visco-elastic
relaxation times of the reed (sect. 3.5).
Once the images are analysed using the procedure described in appendix H, the
resulting geometrical data are oversampled (using linear interpolation) to the pressure
data sample rate and the delay between both measurements is determined by hand,
producing the red line in figure 4.8.

4.3. Results

4.3.1. Double-reeds used in this study and operating conditions

Among the great variety of double-reeds that are used in music, we chose as a first target
for these measurements a natural cane oboe-reed fabricated using standard procedures
(by Glotin), and sold to the oboist (usually a beginner oboist) as a final product (i.e., ready
to be played, although most musicians scrape the reed before using it, in order to tune
it to their preferences)
4.3 Results 57

The choice of a ready to use cane reed was mainly retained because it can be consid-
ered as an average reed. This avoids considering a particular scraping technique among
many used by musicians and reed-makers. Of course, this does not greatly facilitate the
task of the reed measurement, because natural reeds are very sensitive to environment
conditions, age or time of usage.
Other reeds were also tested, as a term of comparison with the natural reeds used
in the majority of the experiments. None of these reeds however was produced by a
professional oboist or reed maker, although it will be an interesting project to investigate
the variations in reeds produced by different professionals.
To conclude, the results presented in this section may depend to a certain extent on
the reed chosen for the experiments, and a larger sample of reeds embracing the big
diversity of scraping techniques needs to be tested before claiming for the generality of
the results that will be presented.
Another remark has to be made on the conditions during the experiments. The kind
of reeds used in most experiments are always blown with highly moisturised air. In fact,
in real life, most of the times, reeds are soaked before they are used, and constantly
maintained wet by saliva and water vapour condensation. These conditions were sought
throughout most of the experiments, although their sensitivity to environmental condi-
tions was also investigated.
In our measurements, humidification is achieved by letting the air flow through a
plastic bottle half-filled with hot water at 40◦ C (see fig. 4.4), recovering it from the
top. Air arriving in the artificial mouth has a lower temperature, because it arrives
at approximately 10◦ C from the compressed air source. As the experiment is carried
out, the temperature of the water and the air arriving at the reed decreases during
experiment and so does humidification.
All measurements were performed without artificial lips. The presence of lip force
over the reed provides an offset of the pressure applied to the reed, which shows out in
a reduction of the initial reed opening area S0 .

4.3.2. The pressure vs flow characteristics


Using the formula of eq. (4.1), and the calibrations carried out for the diaphragm used
in the measurement, the flow is determined from the pressure inside the reed (p r ). The
pressure drop in the reed corresponds to the difference between the mouth and reed
pressures ((∆p)r = pm − pr ). Flow (q) is then plotted against the pressure difference
((∆p)r ), yielding a curve shown in figure 4.9.
In this figure, the flow is seen to increase until a certain maximum value (at about 6
kPa). When the pressure is increased further, flow decreases due to the closing of the
reed. Instead of completely vanishing, the flow tends to stabilize at a certain minimum
value, indicating that it is very hard to completely close the reed. The flow stabilises
once a maximum pressure is attained (corresponding to the reed beating pressure p M
in models) rather than being completely blocked by a closed reed.
The fact that the flow does not decrease when the pressure is increased after the
reed shutting suggests that this remaining opening stays more or less constant inde-
58 Chap. 4: The nonlinear characteristics of the reed

0.35

0.3

0.25

flow (q) in l/s


0.2

0.15

0.1

0.05

0
−10 0 10 20 30 40 50
pressure difference (∆ p) in kPa
r

Figure 4.9.: A typical result for the measurement of the flow vs pressure characteristic
of a natural cane oboe reed.

pendently of the force applied to the reed. In fact, the flow tends to increase slightly,
as expected if the opening remains constant and the pressure difference is increased.
The rate of growth of the flow should be smaller than the one predicted by the Bernoulli
theorem, because these residual openings are probably very narrow and viscosity is
probably very important.
When reducing the pressure back to zero, the reed follows a different path in the
pressure/flow space than the path for increasing pressures. We will see further down
that this might be due to memory effects of the reed material.

4.3.3. Repeatability

Measurements such as the one shown in figure 4.9 were repeated several times for each
reed, in order to test the effects of non-controlled parameters. In figure 4.10, three
different measurements are shown for the natural cane reed shown in figure 4.9.
The ranges spanned by the pressure and flow on each of the curves are different. For
instance, in run 14, the pressure was increased further after the reed is closed, so that
the curve has a longer tail of residual flow, which corresponds to the remaining opening
area after the reed shuts.
The difference in scalings of the three curves may be due to differences in the reed
opening at rest. In fact, memory effects in the reed material might accumulate not only
from the increasing pressure phase to the decreasing pressure phase but also from one
run to the other.
4.3 Results 59

natural cane reed no. 4 (humidified)


0.35
run 13
0.3 run 14
run 17
0.25
flow (l/s)

0.2

0.15

0.1

0.05

0
0 10 20 30 40 50
pressure (kPa)

Figure 4.10.: Comparison of three different runs for a same reed in similar experimental
conditions.

natural cane reed no. 4 (humidified)


1.2
run 13
run 14
1 run 17

0.8
flow (normalised)

0.6

0.4

0.2

0
0 1 2 3 4 5 6 7
pressure (normalised)

Figure 4.11.: Same data as in figure 4.10 but normalized with respect to the maximum
of each curve in both axis.
60 Chap. 4: The nonlinear characteristics of the reed

If the pressures and flows of the data shown in figure 4.10 are divided by a reference
pressure for each run, the curves can be brought to a closer overlap, so that it is easier
to compare their shape. In figure 4.10, each curve was scaled vertically by the maximum
value of the volume flow qmax and horizontally by the value of the pressure for qmax .
The comparison of non-linear characteristics for a same reed shows that there is
some dependence on parameters that were not controlled during these measurements.
A better coincidence of measurements could probably be achieved by controlling the
reed opening before each run, so that the use of artificial lips is considered necessary
in further experiments.
Interpretation of the data shown in the following sections should thus take into ac-
count the fact that there are some parameters not perfectly controlled.

4.3.4. Effect of the added mass


As was explained above, in order to prevent the oscillation of the reed it is necessary
to attach extra mass to the reed blades, which increases their inertia. This was imple-
mented by sticking one Blue-Tack mass to each blade of the reed. It is expected that the
inertia increase only affects the dynamic behavior of the reed, not the static character-
istic that we are trying to measure. However, due to the non-negligible contact surface
between these masses and the reed blades, it is possible that the elastic behavior of the
reed is affected by the presence of the masses.
In order to check this, two consecutive measurements were done, with and without
mass, as depicted in figure 4.12 in non-normalised coordinates.

0.35
with added mass
without added mass
0.3

0.25
flow (q) in l/s

0.2

0.15

0.1

0.05

0
−5 0 5 10 15 20 25 30 35
pressure (∆ p)r in kPa

Figure 4.12.: Comparison of the characteristic curves for a same reed, with and without
added mass on the reed’s blades. These curves are not normalised.
4.3 Results 61

As expected, there are some oscillations of the reed in the decreasing portion of the
characteristic curve (they are not very pronounced due to the presence of the diaphragm
which reduces acoustic oscillations). 4
If we neglect the effects due to the oscillations of the reed, the two characteristic
curves in figure 4.12 look very similar. Since, as seen in the subsection above, there
are variations in measurements effectuated in the same experimental conditions, there
is no evidence in this comparison that the mass affects the measurement of the reed.
However, if the slight change in the maximum flow is in fact a consequence of the added
mass, this could mean that the contact between the mass and the reed slightly increases
its stiffness.
It is difficult however, to draw conclusions on an effect in stiffness alone. Adding and
removing masses requires some manipulation of the reed which can change the reed
opening at rest (S0 ). The fact that a scaling is only observed in the q-axis is an indication
that both k and SO change during this manipulation, because both these parameters
have effects in both axis.
The mass increase due to the addition of the Blu-Tack to each blade can be estimated
by comparing the frequencies of the oscillations that arise in some runs. Using the model
of a harmonic oscillator
q and neglecting the coupling with the flow, these frequencies are
given by f = 21π k
m
. Supposing that the stiffness k is not affected by the addition of
masses, the mass relation is estimated by m2 /m1 = (f1 /f2 )2 .
In these experiments, the reed is found to vibrate at 1000 to 1200 Hz without added
masses, while with the smaller masses the frequency drops to around 600 Hz. The
added masses are thus estimated to be 2.3 times heavier than the intrinsic mass of the
reed.
In most of our experiments, the elimination of auto-oscillations required heavier
masses, about twice in size as the masses used for the above estimations, which are
thus estimated to be 5 times heavier than the intrinsic mass of the reed.

4.3.5. Effect of humidity

The water contained in the reed has great effect on the elastic properties of the reed mate-
rial (for natural reeds). This is shown in several works by Casadonte [Casadonte, 1995]
or Obataya [Obataya and Norimoto, 1999], and referred to in section 3.5.
In terms of the quasi-static characteristic, the effect of humidity is striking: for some
dry reeds, it was not possible to measure a complete characteristic, because oscillations
would always overcome at some point during the measurement, whatever the diaphragm
and added mass configurations. As explained in section 4.3.1, most of the results shown
in this section required a moisturized air supply which was almost saturated in water.
However, the moisturised air running through the reed does not completely saturate
the reed material in water. It is usually possible to increase the reed content in water by
4
During the massless measurement, the read oscillated twice while increasing the pressure and twice
while decreasing it back to zero. The moments where the reed oscillates can be identified by the
discontinuities in the derivative of the characteristic curve.
62 Chap. 4: The nonlinear characteristics of the reed

soaking it for some minutes, a technique that is used by a few oboists and bassoonists.
In figure 4.13 we compare two successive measurements for a same reed, one of
which is done by supplying the artificial mouth with moisturized air but the reed was
dry before the beginning of the experiment, and for the other the reed was left soaking
in water for 5 minutes.
Because the reed surface is wet for the case of the soaked reed, the Blue-Tak masses
do not stick to it, so that it is necessary to create a complete ring of Blue-Tak that
encircles the two reed blades, as shown in the right-hand sketch of figure 4.3.

natural cane reed no. 5


0.35
moisturized air
0.3 soaked

0.25
flow (l/s)

0.2

0.15

0.1

0.05

0
−10 0 10 20 30 40 50
pressure (kPa)

Figure 4.13.: Comparison of the characteristic curves of a same oboe double reed with
two different levels of water content.

The scaling of the soaked reed characteristic (for the increasing pressure phase) is
about 2/3 of the humidified reed, and this can be related to the fact that its opening
at rest is smaller when the reed is soaked (more water content in the reed softens it,
reducing the natural curvature of the blades). It is also possible that the reed is more
closed in the wet case, because of the ring surrounding the reed (whereas Blue-Tak is
only stuck to a small surface when the reed is not soaked)
One other important fact is that the distance between increasing and decreasing
pressure paths is higher for the soaked reed case.

4.3.6. Variations between reeds

In addition to the study of the influence of different factors on the same reed (added
masses, humidity, etc.) it is interesting to compare two similar reeds (same instruments,
same mark, same model).
4.3 Results 63

comparison between two oboe reeds


0.35
reed 4
0.3 reed 5

0.25
flow (l/s)

0.2

0.15

0.1

0.05

0
−10 0 10 20 30 40 50
pressure (kPa)

Figure 4.14.: Comparison of the characteristic curves of two oboe reeds of the same
mark and model, both fed with moisturized air.

A comparison between the two reed’s characteristics (figure 4.14) shows that differ-
ences between two different reeds seem smaller than the variability that can be observed
within a same reed (check figure 4.10).

4.3.7. Differences between bassoon and oboe reeds

Since oboes are not the only double-reed instruments, it is interesting to compare the
non-linear characteristic curves from different instruments. The bassoon is a typical
case of another modern instrument that uses a different kind of double reed based on
the same principle but with dimensions that are very different from those of the oboe.
Figure 4.15 compares two characteristic curves for natural cane oboe and bassoon
reeds. Both were done trying to use similar experimental conditions. The reed was
introduced dry in the artificial mouth, but the supplied air is moisturized at nearly
100% humidity, and masses were added to both reeds to prevent auto-oscillations. The
diaphragms used in both measurements are different however, and this is because the
opening area of the reed at rest is much larger in the case of the bassoon, so that a
smaller resistance (a larger diaphragm) is needed to avoid that the reed closes suddenly
in the decreasing side of the characteristic curve (see section 4.2.3). This should not
have consequences in the measured characteristic curve.
In the flow axis, the bassoon reed reaches higher values, and this is probably a
consequence of its larger opening area at rest, although the surface stiffness is likely to
change as well from the oboe to the bassoon reed. In the pressure axis, the bassoon reed
64 Chap. 4: The nonlinear characteristics of the reed

Pressure flow characteristic


0.5
Clarinet
Oboe
0.4 Bassoon

0.3
Flow (l/s)

0.2

0.1

0
−5 0 5 10 15 20 25 30
Pressure difference (kPa)

Figure 4.15.: Comparison of the characteristic curves of different reed exciters for dif-
ferent instruments. Clarinet data was obtained by S. Ollivier and J.-P.
Dalmont [Dalmont et al., 2003] for a PlastiCover® reed. Oboe and bas-
soon reeds used moisturised air.

extends over a smaller range of pressures so that the reed beating pressure is about 17
kPa in the case of the bassoon reed whereas it is near 33 kPa for the oboe reed.
Apart from these scaling considerations, the shape of the curves are similar and this
can be better observed is flow and pressure are normalized using the maximum flow
point of each curve (figure 4.16).

4.3.8. Comparison with single-reeds

The excitation mechanism of clarinets and saxophones share the same principle of
functioning with double reeds. However, there are several geometric and mechanical
differences between single reeds and double reeds. For instance, flow in a clarinet
mouthpiece encounters an abrupt expansion after the first 2 or 3 millimeters as it
crosses the channel between the reed and the rigid mouthpiece, and the single reed is
subject to fewer mechanical constraints than any double reed. These differences suggest
that the characteristic curve of single-reed instruments might present some qualitative
differences with respect to the double reed [Vergez et al., 2003].
The non-linear characteristic curve of clarinet mouthpieces has been measured by S.
Ollivier an J.-P. Dalmont [Dalmont et al., 2003] using similar methods as the ones we
used for the double reed. A comparison between the curves for both kinds of exciters
(in figure 4.15) shows that the overall behavior of the excitation mechanism is similar
4.3 Results 65

Normalized pressure flow characteristic


1.2
Clarinet
Oboe
1 Basoon

0.8
Flow

0.6

0.4

0.2

0
0 1 2 3 4 5 6
Pressure difference

Figure 4.16.: Normalized data from figure 4.15

in both cases. Similarly to when comparing oboe to bassoon reeds, the scalings of the
characteristic curves of single-reeds are different from those of oboe reeds, although
closer to those of the bassoon. This is probably a question of the dimensions of the
opening area.
A different issue is the relation between reference pressure values in the curve (shown
in the adimensionalized representation of figure 4.16). As predicted by a simple theo-
retical Bernoulli based model and a linearly elastic reed (see chapter 2), in the single
reed the pressure at maximum flow is about 1/3 of the beating pressure of the reed,
whereas in double-reed measurements, the relation seems to be closer to 1/4.
Figure 4.16 also shows that in the clarinet mouthpiece used by S. Ollivier and
J. P. Dalmont the hysteresis is relatively less important than in both kinds of double-
reeds. In fact, whereas the measurements for double-reeds were effectuated in wet
conditions, the PlastiCover® reed used for the clarinet was especially chosen because
of its smaller sensitivity to environment conditions.

4.3.9. Comparison with synthetic reeds

Some measurements and experiments related to the double reed were effectuated in
synthetic material reeds, because of the greater simplicity of use (for instance, they are
not as sensitive to humidity conditions as natural cane reeds). Characteristic curve
measurements were also done on these reeds in order to establish a term of comparison
with other measurements in this document. One of such measurements, for a plastic
66 Chap. 4: The nonlinear characteristics of the reed

reed, is represented in figure 4.17, as compared to a natural cane reed.

comparison between reeds of different materials


0.35
cane
0.3 plastic
fibercane
0.25
flow (l/s)

0.2

0.15

0.1

0.05

0
0 10 20 30 40 50
pressure (kPa)

Figure 4.17.: Comparison between characteristic curves of oboe reeds in natural cane,
plastic and Selmer Fibercane® (synthetic fibrous material)

The plastic reed characteristic curve is a reduced version of the natural cane reed.
This can be mainly due to the fact that these reeds are usually sold with a smaller
opening area at rest, so that they are easier to blow, but provide less control over the
instrument.

4.4. Partial conclusions

In this chapter we presented an experimental procedure to characterise the quasi-static


behavior of double-reeds. The results are given as a nonlinear relation between the
pressure difference across the reed (staple included) and the induced volume flow. While
mainly focused on oboe reeds, an example of a bassoon reed was also presented.

4.4.1. Differences relative to the elementary model

In section 2.4 a simple model was presented for the behavior of a reed instrument
exciter. It had been previously shown [Dalmont et al., 2003] that this model correctly
describes the exciter of a clarinet mouthpiece throughout most of the range of operating
pressures that are spanned by the instrument when a note is played in the clarinet.
Some suppositions made for the case of double reeds [Hirschberg, 1995] indicated
that this model would have to be adapted for the case of double-reeds. In particular,
4.4 Partial conclusions 67

a significant head-loss in the reed would induce abrupt transitions between open and
closed states, even if the pressure driving the reed varies smoothly.
The measurements provided in this chapter show that this is not the case in double-
reeds. The predicted displacement of the maximum of the characteristic curve towards
the reed beating pressure is not observed. Moreover, the displacement seems to be in
the opposite direction, implying a smoother transition between closed and open states,
if the bore pressure driving the reed movement is also smooth.
Relative to the model of section 2.4, the measured characteristic curves for double-
reeds show a relation of 1/4 between the pressure corresponding to the maximum of
the curve PT and the beating pressure PM , rather than the 1/3 predicted by the model.
This can be an effect of the geometry of the staple, which can be assimilated to a conical
diffuser.

4.4.2. Other remarks


The variety of reeds measured in this chapter, and also the variability of the measure-
ments within a single reed indicate that the generic shape of the characteristic curve,
and the displacement of PT relative to what is predicted by the elementary model is a
constant for all the tested double-reeds.
Nevertheless, the scalings of the curve along the pressure and volume flow axis show
an important variability. The fact that this variability is observed in a same reed mea-
sured in constant (as far as possible) experimental conditions is an indication that these
scaling variations are most probably induced by the variations in the S0 parameter (the
reed opening area at rest), since it is not likely that the stiffness k varies widely from one
measurement to another performed a few minutes later (as is the case in figure 4.10,
for example).
In fact, according to equations (2.28-2.29), an increase in S0 should enlarge the curve
scaling both in the p and q axis. The scaling on the q axis should be more increased
(proportionally to S03/2 ) than on the p axis (proportionally to S0 ), and this seems to be
the trend by observing the position of the maximum of the curves in figure 4.10.
In the actual state of this work it is still not possible to provide quantitative arguments
for these conclusions. Further investigation of the variation of the curve scalings using
a precise control and determination of the reed opening can help in determining the
source of the variability of the curves, in a similar manner as what was done for the
clarinet mouthpiece.
Finally, comparing the characteristics curve of a natural cane reed to that of a syn-
thetic cane reed (more closed at rest, in general) suggests that the different scales of both
curves are mostly due to the change in reed opening. Synthetic reeds, which are mostly
intended for beginners, are probably fabricated with narrower openings to reduce the
amount of pressure needed to play a note.
The observations of this chapter are analysed in chapter 5, in the light of a more
detailed investigation of the flow and the proposal of a model for the static flow in the
reed.
68 Chap. 4: The nonlinear characteristics of the reed
5. Details of the flow inside the reed

The flow output from the reed is analysed in detail when uncoupled to the instrument
resonator, in a static regime. Reed oscillations are completely avoided by attaching it to
the artificial mouth walls. The flow output is analysed first through direct observation of
Schlieren photographs, and then through a hot-wire scan of the velocity profiles near the
reed output. These observations are used to improve the description of the flow done in
the previous chapter.

5.1. Introduction

In the preceding chapter, the overall characteristic curve of the double-reed was mea-
sured, providing the volume flow for a typical range of pressure differences between
the mouth and the reed output. These measurements state the relation between pres-
sure and volume flow in quasi-static regimes, an important result because the coupled
behavior of the reed with the resonator will be determined by this p/q relation.
These results however do not give any details about the flow distribution inside the
reed or at its output. In particular, the deviation of the measured nonlinear charac-
teristic curves from the standard model used for reed exciters (section 2.4) is yet to be
explained. A hint is given by the geometry of the double-reed duct, which is different
from the geometries of clarinet or saxophone mouthpieces. However, a more in-depth
analysis of the flow inside the reed and at its output can provide some important knowl-
edge about the functioning of this kind of exciters.
Moreover, a different method of measuring the flow inside the reed can serve as a
term of comparison with the method used in chap. 4, in order to test for the influence
of extraneous devices that had to be added to the double reed such as the diaphragm
and the added masses on the reed blades.
Another motivation for studying the details of the reed flow is evaluating the relevance
of the hypothesis of unidimensional variation of the acoustic properties (pressure and
flow) in the bore, including at the region of coupling to the reed (see section 2.3).

5.2. Overview

The chapter starts by describing the problems implied in measuring complete flow ve-
locity fields inside and near the reed (sect. 5.2).
Section 5.4 shows some photographs of visualisations of the flow at the reed output,
drawing some generic conclusions on the flow regime inside the reed.
70 Chap. 5: Details of the flow inside the reed

In section 5.5 we describe the experimental setup that is used to collect the velocity
profiles of section 5.6. In the latter section, some comparisons are made between dia-
metrical profiles measured for different values of the mouth pressure and reed opening.
These profiles are discussed in section 5.7.
Section 5.8 uses these results to calculate the volume flows in the reed and compares
these to volume flows measured in chapter 4.
Section 5.9 relates the average flow at the reed input to the pressure, using data from
chapter 4.
Finally, section 5.10.4 proposes some models to describe flow and pressure variations
inside the reed, and uses data from this and the preceding chapter to test these models.

5.3. Difficulties in measuring a complete velocity field


Despite the interest in measuring a complete flow velocity field inside the reed, sev-
eral practical constraints made this objective unreachable, at least given the technical
conditions available in our and collaborating laboratories.
Another difficulty arises when trying to maintain the measurement conditions as
close as possible to those in which the reed is used. It would be desirable to be able
to measure the flow inside the reed when coupled to the resonator, because at the
uncoupled reed output the boundary conditions are different from the case of a coupled
reed, which inevitably changes the distribution of the flow not only at the reed output
but inside the reed as well.
In general, measurements or observations of the flow require a relative accessibility
to the flow. In the case of the double reed this becomes a limiting factor, because of the
small dimensions and geometry of the reed (in particular, the oboe double-reed).
For example, several attempts were made to directly visualize and measure the flow
field inside the reed: in particular, Particle Image Velocimetry (PIV) and Laser Doppler
Anemometry (LDA) were addressed and evaluated in collaboration with P. Gougat and F.
Lusseyran, specialists in fluid measurement and visualisation at the LIMSI 1 . However,
these methods have the inconvenient of requiring direct visual access to the flow. Some
attempts were made to produce a transparent double reed in order to have visual access
below the blades, but the curvature and the small dimensions of the reed had several
difficulties which could not be overcome. Moreover, in cases where the auto-oscillations
of the reed are not prevented, different materials would cause a change in the mechanical
properties of the reed.
Given these practical constraints, hot-wire anemometry prove to have a greater flexi-
bility in terms of accessibility to the interior of the reed. Its main drawback is the fact
that it can interfere with the flow, whereas PIV and LDA do not require the insertion of
a material probe in the flow.
The length of the hot-wire used in our measurements is about 1.2 mm and its width
less than a hundredth of a millimetre, so that most of the ending part of the staple
of the reed can be accessed (cf. figure 3.1). Near the walls however, measurements
1
Laboratoire d’Informatique pour la Mécanique et les Sciences de l’Ingénieur, (CNRS – UPR3251) in Orsay
5.4 Visualisation of the flow at the reed output 71

are impossible with the hot-wire probe, and a security distance always has to be kept
between the probe and the reed walls, because of the fragility of the hot-wire. Due to
these limitations, measurements along a diametral line were taken outside the reed, at
a safe distance from the reed walls (approximately 0.5 mm).
Measuring a complete velocity field over a cross-section of the flow is also difficult
in practice, so that some key points have to be selected which are expected to pro-
vide information about the whole field. Nevertheless, a general idea of the flow, for
example, its laminar or turbulent characteristics was preliminarily obtained by flow
visualisation, using the Schlieren technique. This method has been used successfully
to observe both static and variable flows in music instruments, such as in organ flutes
[Verge et al., 1994].

5.4. Visualisation of the flow at the reed output

The Schlieren technique takes advantage of the phase variation of light beams when
crossing a medium with a different refractive index. A Schlieren snapshot shows the
gradient of the refractive index as intensity levels [Merzkirch, 1974]. It was used to
observe the flow at the output of the reed, by blowing the reed in the artificial mouth
with carbon dioxide. These experiments were carried out using the Schlieren setup
developed by B. Fabre [Fabre, 1992] [Fabre et al., 1996] at the LAM2
Figure 5.1 shows Schlieren images of the flow at the reed output for three different
mouth pressures pm . The free jet far from the reed output turns from laminar (at very low
pressures, not used in playing conditions) to turbulent (in normal playing conditions).
Turbulence is observed above 3 mBar in CO2 . In air this value should be corrected
because of the differences in viscosity (νC02 /νair )2 = 0.28, so that this critical pressure
would be ' 1 mBar, for a same critical Reynolds number in CO2 ).
Although the jet boundaries are straight when leaving the reed output, this does not
mean that the flow is not turbulent at this point. In fact, because the duct walls are
smooth until the reed exit, some time is needed for the turbulence to start mixing the
carbon dioxide to the air. During this time, the jet can travel a short distance during
which its walls remain fairly straight. Given the Reynolds numbers involved in the two
bottom snapshots, it is fair to conclude that the flow is turbulent along most of the
length of the reed for the range of pressures used while playing the instrument.
In the top image of figure 5.1, the bending of the jet relative to the reed axis is mostly
due to weak air currents in the room, as could be verified by the original sequence of
images that showed fluctuations in the jet direction. In fact, the flow velocity at the reed
exit calculated by Bernoulli is about 0.5 m/s, a value that is comparable to currents
generated by the motion of people and objects inside the room. Two other effects may
account for this bending, the Coanda effect (attachment of a flow to the solid boundaries
due to viscous entrainment [Hofmans, 1998]) and gravity. The latter can be calculated
with an estimation of the flow velocity at the reed output (see appendix E)

2
Laboratoire d’Acoustique Musicale in Paris
72 Chap. 5: Details of the flow inside the reed

pm = 15 Pa
Re = 700

pm = 300 Pa
Re = 3400

pm = 15 kPa
Re = 19700

Figure 5.1.: Flow at the reed output at different blowing pressures. The reed opening is
kept fixed.
q Reynolds Numbers are calculated based on the mouth pressure
2pm
(u = ρ
) and the reed opening height (d = h ' 10−3 m)
5.5 Hot-wire measurements 73

Another interesting fact is that at the output the jet is limited exactly by the solid
boundaries which constitute the reed channel. This is important because the channel
is slightly divergent (about 5.2◦ between walls, 2.6◦ relative to the cone axis, calculated
using data from fig. 3.1). It could be possible for the flow to detach from the reed walls,
but we see that if this happens inside the reed, the jet reattaches before arriving at the
reed output. This agrees with the known result [Ouziaux, 2003] that flow detachment
occurs in conical diffusers with a tapering angle wider than 8◦ .

5.5. Hot-wire measurements


These measurements were performed at LAM, with the help and assistance of B. Fabre.

5.5.1. Experimental device


The setup used for hot-wire anemometry measurements is represented schematically
in figure 5.2. It is similar to the setup used below for anemometry in dynamic regimes
(fig. 6.3), although some of the instruments shown in this photograph were only used
in section 6.4.
Artificial
mouth

Manometer
Compressed air
CO2

Camera Reed
Hot−wire

Lens

Figure 5.2.: Experimental device

Most of the experimental device is mounted on an optics rail. The artificial mouth is
kept fixed in the same position throughout the experiments, although for some experi-
ments it may be turned 90◦ around the axis of the reed output.
In the reed axis, in front of it, a camera allows the recording and measurement of
the reed opening during the experiment. Between the artificial mouth and the camera,
a convergent lens magnifies the reed opening area in order to maximize the digital
resolution of the image that will be analysed.
74 Chap. 5: Details of the flow inside the reed

Each blade of the reed is kept fixed relative to the walls of the artificial mouth using a
miniature screw attached to a ball-and-socket joint device (fig. 5.3) conceived by Alain
Terrier at the IRCAM mechanics studio. The device is attached to the reed at about
2 mm from the tip, so that a small length of the reed would still be free to vibrate.
It is interesting to notice that in fact it does not, confirming that the action of the
embouchure cannot be to completely prevent oscillations of the reed along the surface
of contact between the lips and the reed as stated in section 1.4.2.
The reed opening cannot be precisely determined with the screws, so that the reed
opening recordings are used during the exploration of the data. One reed opening
image is recorded for each set of velocity measurements, in order to precisely track
slight opening changes due to alterations in the mouth pressure.

Artificial
mouth
walls

Positioning
Scew

Pinhead
Epoxy
Joint

Reed blade

Figure 5.3.: Reed fixation device, used to adjust the reed opening. The position of each
blade is adjusted relative to the artificial mouth walls.

The hot wire probe is placed in the continuation of the reed, that is, along the flow
leaving the reed. It is mounted on a three-axis precision stage with micrometric actua-
tors, so that the probe position can be adjusted with a precision of 0.01 mm.
The points of measurement are situated on a diameter of the reed (x or y in figure
5.4), 0.5 mm away from the reed output (distance d in figure 5.4, which is not adjusted
precisely before the experiment, but is more precisely measured on a photograph).

Figure 5.4.: The measurement plane and definition of the diametral directions
5.6 Velocity profiles 75

5.5.2. Calibration of the hot wire


The hot wire probe has to be calibrated very carefully using a well-known velocity source,
such as a carefully designed flue exit used for example for the study of flutes and
other flue instruments [Verge et al., 1994]. Calibration consists in relating the tension
output from the hot-wire system with the velocity calculated with the Bernoulli formula
for the corresponding pressure difference across the slit. Since the device response
is non-linear, several velocities are required for the calibration curve, in a range that
includes the velocities to be measured. Because of temperature fluctuations inside the
laboratory, the hot wire probe was calibrated every two hours.

5.5.3. Double-reeds used in these measurements


Although the experimental device described above (sect. 5.5.1) can be used in principle
with most kinds of double-reeds, only one double-reed was used in the measurements
shown below. This fact is mainly due to the difficulty involved in changing reeds: they
are attached to the immobilizing device (fig. 5.3) with cyanolite, so that changing the
reed is a lengthy process and destroys the reed currently mounted in the device.
We chose a synthetic reed (Selmer Fibercane), easier to glue with cyanolite than a
natural cane reed. In fact, we expect that the results are not much influenced by
the material, since in this chapter the elastic properties of the reed are not important
(because the reed is kept fixed as much as possible).

5.6. Velocity profiles


Velocity profiles are measured on a diameter of the backbore exit (4.8mm long), as
close as possible to the output (closer than d = 0.5mm). The main diameter used for
measurements is parallel to the largest dimension of the reed opening (x in figure 5.4),
although some investigation was made on the perpendicular diameter.

5.6.1. Variation of diametrical profiles with the mouth pressure


A first investigation consisted in measuring the velocity profile along the diameter x
(fig. 5.4) of the reed output. The pressure was changed while keeping the reed opening
fixed. Figure 5.5 displays the results for three different blowing pressures.
Figure 5.6 displays the velocity profiles once normalized relative to their maximum
velocity. It reveals a very similar structure of the flow. This structure is conserved
for the range of pressures used in normal playing of the instrument, at least for high
Reynolds numbers.
This has an important consequence on the calculus of the volume flow since it means
that for a given blowing pressure, the volume flow is the product of a reference volume
ref
flow (calculated by integration of a reference velocity profile) by the ratio V max /Vmax ,
ref
where Vmax is the velocity measured at the center of the output section, and Vmax is the
corresponding velocity for the reference case (see sect. 5.8 for an application).
76 Chap. 5: Details of the flow inside the reed

50
15 mBar
130 mBar
40 65 mBar

Flow velocity (m/s) 30

20

10

0
4 6 8 10 12 14 16
Probe position (mm)

Figure 5.5.: Comparison of diametral velocity profiles at approximately 0.5 mm down-


stream of the reed output, for three different mouth pressures. Dashed
vertical lines indicate the position of the reed walls, and the dash-dotted
line the midpoint between the walls.

1
15 mBar
130 mBar
Normalised flow velocity

0.8 65 mBar

0.6

0.4

0.2

0
4 6 8 10 12 14 16
Probe position (mm)

Figure 5.6.: Same data as figure 5.5, but normalized so that maximum points equal 1
5.6 Velocity profiles 77

Note that velocity measurements correspond to time-averaged values. As seen in


section 5.4, flow is turbulent for the pressures used in these measurements, so that
fluctuations are observed around these averaged values. Unfortunately, the time evolu-
tion of the profiles was not recorded during these experiments in order to have an idea
of the range of these fluctuations.
As an indication, inside the reed the fluctuations in pressure range from 7% to 10%
of the actual value of the measured pressure for the lowest pressures in the turbulent
regime (cf. fig. 5.21).

5.6.2. Variation of diametrical profiles with the reed opening


Given the invariance of the velocity profiles against the mouth pressure pm , a similar
question can be asked about the influence of the reed opening on the structure of the
flow output. Indeed, in addition to a change in the input cross-section of the reed (which
affects the total flow that enters the reed), the reed opening has consequences on the
geometry of the reed duct (which unfortunately were not measured).
The line over which the profiles were measured is similar to the one used in the above
section (along the x-axis). Neither the probe position (along y and z axis) was changed,
nor the artificial mouth position. The pressure was kept constant at 65 mBar, and the
reed opening was changed using the device described in fig. 5.3.
The three normalized profiles are presented in figure 5.7. Again, similar profiles are
found. The differences between the profiles, and in particular the slight asymmetry of
the profiles can be a consequence of the asymmetry or displacement of the reed opening
relative to the reed main axis (each blade of the reed is pushed or pulled independently
using the micrometric screw mechanism). One other reason may be the Coanda effect,
an attachment of the flow to a wall. This is observed in similar experiments on glottal
flow [Hofmans, 1998].

5.6.3. Comparison of two perpendicular diameters


Although the downstream part of the reed (towards the butt) is axisymmetric, the en-
trance channel (at the reed tip) is highly asymmetric, due to the oblong shape of the
reed opening (see fig. 3.1).
In the previous section, results tended to show that changes in the initial geometry
of the reed (reed opening) were erased along the reed length. However, an effect of the
geometry concerning the position of the reed opening relative to the reed axis seemed to
be noticeable in the profiles of fig. 5.7, although this was not sought on purpose.
A further test of the effects of the initial geometry on the output flow can be inves-
tigated by comparing the profiles measured in two perpendicular directions, for which
the asymmetry in the upstream part of the reed is more evident (these are the x and y
directions in fig. 5.4).
The profile measured along x shows higher measured velocities around the axis,
suggesting that in fact the flow keeps some traces of the initial asymmetry, despite the
turbulent character of the flow. Even if flow velocity data is normalised according to the
78 Chap. 5: Details of the flow inside the reed

1
open
half

Normalised flow velocity


0.8 small

0.6

0.4

0.2

0
0 2 4 6 8 10 12
Probe position (mm)

Figure 5.7.: Comparison of diametral velocity profiles for three different reed openings.
Normalized with respect to the maximum value of the pressure

1
parallel
perpendicular
Normalised flow velocity

0.8

0.6

0.4

0.2

0
0 5 10 15
Probe position (mm)

Figure 5.8.: Comparison of velocity profiles for two perpendicular diameter directions
(parallel – x – and perpendicular – y – to the reed opening). Velocities are
normalised according to the maximum flow velocity
5.7 General discussion on the flow profiles 79

maximum value of velocity (as in fig. 5.8), the discrepancy in the profile is greater than
normalised profiles on the same diameter for varying pressures or reed openings (check
previous sections).

5.7. General discussion on the flow profiles


Turbulent character of the flow Direct observations of the flow output from the reed
(sect. 5.4) show that the flow at the reed output is turbulent for most of the range of
pressures used when playing the instrument. This conclusion seems to be valid for
the upstream part of the reed duct and, given the Reynolds numbers of these flows,
turbulence is likely to develop almost as soon as air enters the reed.
All the mouth pressures used in the profile measurements shown in the previous
sections are in the turbulent range, as observed directly through Schlieren snapshots
in section 5.4, and is suggested by a calculation of the Reynolds number. Additionally,
pressure measurements inside the reed indicate that the flow is turbulent in this region
for mouth pressures (pm ) above 500 Pa (cf. fig. 5.21).

Cancelation of the initial asymmetry The turbulent character of the flow seems to cancel
out most of the upstream asymmetry that the reed may induce on the flow. The close
resemblance of the profiles measured in two perpendicular diameters confirms this
conclusion. Two observations however may indicate an opposite conclusion:

Asymmetry of measured profiles The downstream reed duct does not share its axis of
symmetry with the measured profiles. The maximum velocity seems systematically
displaced about 0.25 mm aside from the reed axis (cf. fig. 5.7). This displacement
cannot be due to uncertainties in the position of the probe, because the maximum
value is not only displaced from the reed axis, but also relative to the boundary layers
of the flow. In fact, in fig. 5.6, the core of the flow seems to reach lower velocities (about
50% of the maximum velocity) in the right-hand side of the graphic than in the left-hand
side (about 60 % of the maximum velocity).
There may be several causes for this displacement:

• The first would be that the reed opening changes position relative to the down-
stream channel, however the profile is measured in the same direction as the reed
opening major axis, so that the reed position should not have major consequences
in the measured profile.

• Coanda effect: due to the tapering of the downstream duct, the flow may attach
partially to a wall, deviating from the symmetric position. It is unlikely, however
that the deviation does not change with the flow intensity.

• The direction of measurement may be deviated from a reed output diameter and
have an angle with the hot-wire itself. One of the edges would then leave the flow
first through smaller velocities whereas in the other edge it will leave first through
80 Chap. 5: Details of the flow inside the reed

higher velocities. In other words, in one edge the hot-wire can be parallel to the
boundary layers at some point, whereas in the opposite edge it is tilted relative to
the jet’s walls (fig. 5.9).

measurement
direction

hot−wire measured
profile
real profile

Figure 5.9.: Differences between the measured profile and the real profile

In the first 4 centimeters of the z-axis, the reed geometry has an opposite tapering
in each of the directions x and y (cf. fig. 3.1). Along the x axis, the width of the duct
decreases in the downstream direction. A convergent duct is usually used at the input
of wind tunnels [White, 2001] to flatten the velocity profile of the flow. In the literature,
analysis of velocity profiles along the diffuser are usually based on flat profiles at the
diffuser inlet [Azad, 1996].
In the y-direction of the reed, the duct is mostly divergent throughout the first 2.5
cm of the reed (cf. fig. 3.1), so that at the diffuser inlet the flow profile is expected to be
decreasing from the axis to the edges.
In reality, this separate discussion about the flow according to the direction is prob-
ably too simplistic, because the turbulent character of the flow is likely to induce a
coupling between the xz and yz planes. However, it gives some insight on the differ-
ences observed in both directions, in particular the steeper profiles around the axis
observed in the y-direction.

Comparison with flow in a straight tube Considering the small tapering of the reed duct
(check fig. 3.1), and the absence of flow detachment in the conical diffuser, the flow in
the downstream part of the reed can be compared to a flow in a straight pipe (fig. 5.10),
if in a first approach we forget the slight tapering. It is pointless to consider laminar
flow solutions, because the flow is probably turbulent all the way from the reed tip.
The measured profiles show similar characteristics to developed turbulent flow solu-
tions in straight smooth pipes [White, 2001], in particular, the steep boundary layers
that surround the core flow. However, in a straight pipe, the core flow is relatively flat
and surrounded by a steep transition region to zero velocities. The bulging of the central
region of the measured profiles can be due to the conical diffuser effect: central flow
is ‘‘pushed’’ harder than surrounding regions by the incoming flow, which is directed
along the axis.
5.8 Volume flow calculations 81

1.4
1.2
1

uUavg
0.8
0.6
0.4
0.2

-1 -0.5 0 0.5 1
rR

Figure 5.10.: Approximate velocity profile of a turbulent flow in a cylindrical pipe with
smooth walls

Flow detachment It is also interesting to recall that as mentioned in sect. 5.4, in diffusers
with larger tapering angles the flow can detach from the diffuser’s walls due to a highly
adverse pressure gradient. It is a well-known result [Ouziaux, 2003] that this effect
occurs in general for tapering angles larger than 8◦ , whereas the reed staple tapering
angle is 5.2◦

Widening of the jet In all the measured profiles (cf. for example fig. 5.5) the jet extends
beyond the reed walls. This fact cannot be explained in terms of viscous or turbulent
diffusion of the jet, because the widening of the jet of original diameter D is known to
start at a distance z0 ∼ D [Bailly and Comte-Bellot, 2003] from the output. The profiles
were measured at a much smaller distance from the output 0.5 mm, versus a diameter
of 4.8 mm.
A possible is that the jet walls continue the tapering of the conical diffuser. Based on
a tapering angle of 2.4◦ , as seen in section 5.4, the diameter difference between the jet
and the reed output would be 0.02 mm, too small to be noticed in the profiles.
A more plausible explanation is that the hot wire is not perfectly perpendicular to
the diameter along which the profiles are supposed to be measured, or that there is a
systematic deviation of the hot-wire from the diameter of the jet. Given that the observed
widening is approximately 0.2 mm, and the hot-wire length is 1.2 mm, slight deviations
could explain this difference.
A symmetric widening could be explained by figure 5.9 either if the measurement line
is a diametrical line or if the hot-wire is perpendicular to the measurement line. Based
only on an inclination of the hot-wire, a widening of 0.2 mm would require an angle of
20◦ , which should have been noticed.

5.8. Volume flow calculations


5.8.1. Volume flow deduced from the maximum of the velocity profile
As seen in sections 5.6.1 and 5.6.2 , normalized velocity profiles can be considered
independent of the blowing pressure and of the reed opening to a good approximation.
82 Chap. 5: Details of the flow inside the reed

As a consequence, volume flows can be deduced from a reference volume flow q ref
through a simple multiplication by a constant coefficient, depending on the velocity
measured at a single point (for example the maximum value) :

Vmax
q= ref
qref (5.1)
Vmax

The reference flow q ref is calculated as follows. The volume flow is written as the flux of
the velocity, neglecting the slight asymmetry that was observed in section 5.6.3. Velocity
is thus considered constant in an annulus of radius r around the axis of the reed.
Z Z rmax
q= udS = 2πrudr (5.2)
0

where rmax should be the output radius of the reed staple. However, in all measurements
it can be observed on the velocity profiles that the jet extends further away from the
output walls of the duct. This has been discussed in section 5.7.
In practice, an empirical choice of the integration limits had to be made. Since there
is an observed background flow underneath 5% of the maximum flow velocity of the
profiles, an appropriate choice was to limit the integration at the position x0.1 where the
velocity reaches 10% of the maximum flow velocity.
The volume flow is then the integral of the velocity profile according to:
Z x0.1
q= 2π |x − xc |u (x )dx (5.3)
0

with xc , 0.5 (x0.1+ + x0.1− ) represents the axis of symmetry of the profile. Finally, the
volume flow is found to fit the following law :

q = 1.27 × 10−5 umax (5.4)

This principle is confirmed in figure 5.11 where four volume flows have been calcu-
lated by the integration of the velocity profile on a circular annulus. As expected, these
points lie on the slope given by equation (5.4).

Deviation from axial symmetry One important factor in the calculation of the volume
flow is whether the profile is independent of the choice of the diametrical direction on
which it is measured. Only a slight difference is observed for the two most asymmetric
directions in terms of the reed geometry (fig. 5.8).
The error produced by supposing that the flow is axisymmetric can be estimated by
calculating the total flow by integration of the two profiles shown in figure 5.8. The
flow based on the y-diameter profile (the direction perpendicular to the profiles used for
fig. 5.11 and 5.12) is 11% smaller than a flow calculated using an x-diameter profile.
For an arbitrary tilted relative to the two measured perpendicular diameters, the
profile should have an intermediate shape, so that the actual volume flow is expected
to be 5.5% smaller in average than the flow calculated using the x-diameter profile.
5.8 Volume flow calculations 83

−4
x 10
8

6
Volume flow (m /s)

q=1.270e−05*vmax+−6.746e−06
3

variable pressure
0 variable opening
fit
0 10 20 30 40 50
Maximum flow velocity (m/s)

Figure 5.11.: Variation of volume flow as a function of maximum flow velocity

5.8.2. Variation of flow with pressure

According to the method presented in section 5.8.1, it is possible to estimate the volume
flow through a measurement of the air velocity at a single point. This is done for five
different reed opening areas, the reed position being controlled as in section 5.6.2. The
hot-wire probe is placed at the center of the air flow. This is practically achieved by using
the micrometric screws to adjust the probe position until maximum output tension is
reached. For each opening area, the volume flow is estimated for several increasing
blowing pressures from 0 to nearly 15000 Pa. Results are shown in figure 5.12.

A mean flow velocity vm at the reed entrance is estimated by dividing the volume
flow q by the reed opening area. The reed opening area is estimated through an image
processing technique developed by the authors ([Almeida et al., 2006]). An illustration
is displayed in figure 5.13. Although only one reed opening area is shown per curve,
this variable was measured for each point in the graphic, in order to account for small
variations of the reed opening. In fact, even if the reed blades are blocked by the device
shown in figure 5.3, there is a small length of the reed tip that is free and can be affected
by the mouth pressure. Area shown in the legend of figures 5.12 and 5.13 is the average
reed opening area for all points in each line.
84 Chap. 5: Details of the flow inside the reed

−4
x 10
6
0.3
5 0.9
1.4
4 1.9

volume flow (m /s)


2.6

3
3

−1

−2
0 5000 10000 15000
pressure (Pa)

Figure 5.12.: Volume flow plotted against mouth pressure for different reed openings (in
mm2 )

5
x 10
2
0.3
0.9
1.4
1.5 1.9
2.6
(q/s) (SI)

1
2

0.5

0
0 5000 10000 15000
pressure (Pa)

Figure 5.13.: Average flow velocity at the reed tip as a function of the mouth pressure, for
different reed openings (corresponding to data of fig. 5.12). The dotted line
corresponds to the flow velocity calculated using the Bernoulli theorem
5.8 Volume flow calculations 85

Figure 5.14.: Picture of the reed opening (left) and automatically detected area super-
imposed on original image (right)

Comparison with results from chapter 4

Volume flow calculations shown in figure 5.12 can be compared with the results shown
in chapter 4. Each of the runs used to plot the runs in this figure correspond to a single
opening area of the reed, shown in the legend.
Some of the characteristic curves measured in chapter 4 were complemented with
opening area measurements, which were later synchronised with flow and pressure
measurements (see sect. 4.2.3). For run 14 in figure 4.10, for instance, the reed opening
area corresponding to each of the points of the characteristic curve was recorded. It
is possible thus, to know the volume flow corresponding to each opening area used in
figure 5.12. With this information, one can compare the volume flows in the free reed
(chap. 4) to those determined in the blocked reed (chap. 5).
For the non-linear characteristics measurements, the reed was free to position itself
for each pressure difference, so that a particular opening area is observed only twice
(once for increasing pressures, and once for the decreasing pressures). Figure 5.15
shows for each reed opening area used in figure 5.12, both the volume flows measured
for a fixated reed (lines) and the corresponding volume flows for the free reed (• increas-
ing pressures, 4 decreasing pressures), using same colors for same opening areas.
Figure 5.15 shows that the volume flows determined in this chapter are in general
twice as large as those from the preceding chapter.

Opening area uncertainty There are several possible reasons for this discrepancy, one
of them is the relative uncertainty in the opening area measurements, however it is not
likely that this uncertainty is as big as 100%, at least for large openings.

Flow uncertainty Another possible explanation comes from the uncertainty of the vol-
ume flows calculated in this chapter using the flow velocities measured by hot-wire
anemometry. Flow velocities should not be affected by big uncertainties, because the
hot-wire probe was carefully calibrated and has been used in several other experiments
with success.
86 Chap. 5: Details of the flow inside the reed

−4
x 10
6
2
0.3 mm
5
2
0.9 mm
4
1.4 mm2
3 1.9 mm2

flow (m /s)
3
2 2.6 mm2

−1

−2
0 1 2 3 4
pressure (Pa) 4
x 10

Figure 5.15.: Comparison of volume flows measured in the fixated reed with similar
opening area volume-flows extracted from the non-linear characteristics
measurements (run 14 in figure )

Velocity fluctuations There is however one detail that can introduce some error in the
velocities measured at one point. As discussed before, these are average values: no
recording was made of the time-evolution of the flow velocity, so that there is no clue
about the range of the velocity fluctuations.
The velocity vector fluctuates not only in the main flow direction but also in perpen-
dicular directions. Due to the symmetry of the hot-wire, it can measure the projection
of the flow velocity on a plane rather than on a single direction. The module of the
projected vector is measured by the probe.
If the fluctuations were only in the same direction of the flow, the average value of the
measured velocity u 0 would still be the same as the average velocity U , but if there are
fluctuations perpendicular to the main flow direction, the average value of u 0 is greater
than U , because whatever the sense of the fluctuations, they cause an increase in the
velocity vector u.

Flow asymmetry Even if considering that fluctuations are not negligible in the mean
value of the flow, the integration of the flow profile has some problems, the main one
being that the asymmetry was verified (sect. 5.6.3) only for one particular configuration
of the reed opening and mouth pressure, and taken into account in the calculation of the
volume flow somewhat empirically. Deviation from axial symmetry can vary according
to the Reynolds number and is probably accentuated when the reed is more closed.

Area correction For the opening area data shown in the next section (5.9.1), it was found
that a residual opening area that cannot be detected by the image analysis technique
5.9 Average flows deduced from characteristics measurements 87

had to be added to the measured opening areas.


The data of chapter 5 can be corrected using a similar principle, considering that
there is also a residual area that cannot be detected by the image analysis. This is done
empirically, by trying several different area corrections.

−4
x 10
4
0.3 mm2
2
0.9 mm
3 2
1.4 mm
2
flow (m /s)

1.9 mm
3

2
2 2.6 mm

0
0 1 2 3 4
pressure (Pa) 4
x 10

Figure 5.16.: Data of figure 5.15, but where the main area measurements (correspond-
ing to the lines) have been corrected by an area offset Sres = 1.1 mm2 .

The best superposition between the two data sets can be obtained by adding a residual
area Sres = 1.1 mm2 to the areas measured with the image analysis technique, for the
data measured with the hot-wire.
Again, this area correction can have several justifications, not only regarding the
opening area measurements themselves, but also the flow measurements. It seems
daring to propose that this residual area is in fact a physical opening not detected
by the image analysis technique, because an area of 1 mm2 is represented by several
thousand pixels in a typical reed opening image. However, lateral openings are not
observable in frontal pictures of the reed.
Flow data from this chapter should thus be used with caution.

5.9. Average flows deduced from characteristics measurements

Some of the characteristics measurements of chapter 4 have associated reed opening


area measurements that can be used to determine mean flow velocities at the reed
entrance, for example.
88 Chap. 5: Details of the flow inside the reed

5.9.1. Small openings


As explained in appendix H, the application of the image analysis algorithm to small
openings is more problematic than for large ones. In particular, a detailed analysis of
the results of a run reveals that the reed appears as closed even if the measured flow
is never 0 (figure 4.10, for example). In fact, flow is never completely stopped, which
means that very narrow gaps remain along the junction between the two reeds (in the
reed tip when the reed closes, or along the reed rails) that are very difficult to close.
By plotting the volume flow against the reed opening area (see figure 5.17), it is
possible to see that before the reed actually closes the curve has a recognizable curvature
tendency.

anche4h_18_2masseslourdes_nonosc14.BIN
0.35

0.3

0.25

0.2
Flow (l/s)

0.15

0.1

0.05

0
−0.5 0 0.5 1 1.5 2 2.5 3 3.5 4 4.5
2 −6
Opening area (m ) x 10

Figure 5.17.: Plot of the volume flow against the opening area of the reed, and extrap-
olation of a residual opening area (red extrapolated from the lowest area
part of the increasing pressure branch, and green from the decreasing
pressure branch)

We use the leftmost part of this curve (except for the points corresponding to an area
very close to zero) to extrapolate a value for the effective closing area. This value (about
-0.35 mm2 in fig. 5.17) is negative and is considered as a constant that is subtracted to
the whole measurement of the opening area. The symmetric of this value (Sres = 0.35
mm2 ) can be seen as a residual opening area, which is probably distributed by several
tiny openings along the reed edges, that cannot be observed in the reed images.
Nevertheless, data for small reed openings should be used with care because it may
or may not be constant when pressure increases beyond that for which the detected
area is zero.
The residual opening area can also be estimated using data from the tail of the char-
5.9 Average flows deduced from characteristics measurements 89

acteristics measurement (fig. 4.9). As a rough analysis, if we consider the Bernoulli


theorem to be valid for these small openings, the volume flow q would be
s
2(∆p)r
q = Sres (5.5)
ρ

Using this relation, and for example for (∆p)r = 33 kPa, and a corresponding vol-
ume flow of q = 0.022 l/s (run 14, decreasing pressure), one finds Sres = 0.095 mm2 .
This value is probably under-estimated because, for very small openings the Bernoulli
formula is no longer correct as the viscous losses become important (see appendix D).
Thus for a same pressure drop, and opening area, the flow would be weaker.

5.9.2. Mean flow velocity estimations


Using the corrected area measurements from the previous section, it is possible to
estimate a mean flow velocity calculated as uavg = Sq at the reed entrance as a function
of the total pressure drop (∆p)r in the reed. This is shown in figure 5.18.

anche4h_18_2masseslourdes_nonosc14.BIN
600
Measured
Bernoulli

500

400
velocity (m/s)

300

200

100

0
−10 0 10 20 30 40 50
pressure difference (kPa)

Figure 5.18.: Mean flow velocity at the reed entrance estimated using data from run
14 in fig. 4.10 and simultaneous reed opening area measurements. The
expected velocity using a Bernoulli model is plotted as a red line.

The flow velocity seems well estimated by a Bernoulli model throughout most of the
pressure range, using a small corrective factor for the opening area, probably to be
related with a Vena Contracta effect or other changes in the flow along the reed (see
below).
For high pressures though, the Bernoulli model seems to be no longer valid. This
region, where the blue curve suddenly drops far below the Bernoulli model, can be
90 Chap. 5: Details of the flow inside the reed

identified in figure 5.17 as the residual flow region, where the reed opening recordings
show that the reed is already closed.

5.10. Static flow model


With the data gathered from the experiments in this and the previous chapter, we are
able to propose a stationary model for the flow along the reed. This model will be
based on certain propositions previously stated in the literature, in particular for the jet
contraction at the reed entrance [Hirschberg, 1995] [Gokhshtein, 1981].
The model introduced in this section serves as an interpretation for the data in the
previous chapters, but care should be taken when transporting this model to dynamic
regimes.

1 2 3
pm

S1 pr
Stagnant
region

PSfrag replacements pm

(∆p )1

(∆p )2

Figure 5.19.: Definition of flow regions along the reed: (1 – Vena contracta, 2 – Flow
reattachment (singular loss), 3 – staple input. In the axial pressure dis-
tribution graphic, the solid line represents a model where the diffuser is
supposed to cover the whole length of the reed, and the dashed line a
diffuser starting at the staple input.

5.10.1. Vena contracta


Because of the inertia of the flow when entering the double reed, it is usual to see the
formation of a jet at the entrance of a pipe with thin walls. In fact, a tight bending
of the streamlines around the tube walls would require the action of infinite forces.
5.10 Static flow model 91

This fact is known in the literature of fluid engineering as the Vena Contracta effect
[Hirschberg, 1995]. A calculation of the discharge coefficient CD is provided in appendix
C.3, based on the momentum and mass conservation equations and the application of
the Bernoulli theorem between the mouth and the reed entrance region.
The extreme case of the Vena Contracta effect is found when the tube wall thickness
is infinitesimal, and the reservoir walls are sufficiently distant from the tube entrance
so that the flow can arrive at the tube entrance at a direction of 180◦ relative to the
main flow inside the tube. In this configuration, known as Borda tube the contraction
coefficient is found to be CD = 1/2.
In a real scenario, the value of CD can vary between 1/2 and 1, depending on the
distance between the tube entrance and the reservoir walls, and the thickness of the
tube walls. Empirical values of CD according to these parameters can be found in
engineering literature [White, 2001] [Parsons, 1997]

5.10.2. Singular losses


When crossing a discontinuity such as the Borda tube, it is known that there are lo-
calised losses in the flow caused by the generation of turbulence and dissipation in
stagnant regions of the flow (for example, the fluid remaining between the jet and the
tube walls).
In a Borda tube configuration for example, the jet is known to reattach to the tube
walls at a small distance from the entrance, leaving a small stagnant region at the
entrance. The expansion of the jet to the full cross-section induces some pressure
recovery on the flow. This recovery cannot be calculated using the Bernoulli theorem
because some energy is lost through turbulence and viscous dissipation in the stagnant
region.
Nevertheless, mass and momentum conservation can still be applied to determine the
pressure difference before (1) and after (2) the expansion (see appendix C.4):
 2
q
p2 − p 1 = ρ CD (1 − CD ) (5.6)
Sj
where Sj is the jet cross-section, before the expansion and CD the contraction coefficient
from the previous section. Tapering has been neglected in this formula because flow
detachment occurs over a short length [van Zon et al., 1990]. This means that S = S2 .
The initial pressure drop is calculated using the Bernoulli theorem between the mouth
(where flow velocity is neglected) and region 1:
 2
1 u
pm − p 1 = ρ (5.7)
2 Sj
and Sj = CD S2 is the jet cross-section affected by the Vena contracta.
According to this model, the pressure recovery fraction at the reed entrance C r =
p2 −p1
pm −p1
can be plotted as a function of the contraction coefficient (fig. 5.20):

p2 − p 1
Cr = = 2CD (1 − CD ) (5.8)
pm − p 1
92 Chap. 5: Details of the flow inside the reed

0.5

0.4

0.3
Cr
0.2

PSfrag replacements 0.1

0
0.5 0.6 0.7 0.8 0.9 1
CD

p2 −p1
Figure 5.20.: Pressure recovery at the reed entrance (Cr = pm −p1
) as a function of the
contraction coefficient (Vena Contracta)

For the extreme case of a Borda Tube, where the jet contraction is maximum (C D = 0.5),
the recovered pressure (p2 − p1 ) is maximum, and equals half of the pressure difference
between the mouth and the jet pm − p1 . For larger values of CD (expected when the
reed blades are thickened or rounded [Gokhshtein, 1981]) the pressure recovery drops
so that the pressure after the singularity (p2 ) approaches the pressure in the jet (p1 ).
One consequence of the difference between p2 and p1 is that the force at the reed tip
can vary for a same total pressure drop (∆p)r in the reed. This can be compared to the
case without contraction using the ratio of the force in both cases:
pm − p 1 1
Cpd = = (5.9)
1/2ρu22 2
CD
For a same volume flow q, the pressure drop in the reed entrance (pm − p1 ) for a Borda
tube case (CD = 0.5) is four times the case without contraction.

5.10.3. Conical diffuser


Figure 3.1 shows that, after the reed entrance the cross-section of the reed increases
continuously throughout most of the reed length, with a small tapering angle (less than
the downstream tapering of 5.2◦ in the staple).
Due to the similarity between the observed flow profiles at the reed output (sect. 5.7)
and profiles in well-studied flows in conical diffusers [Okwuobi and Azad, 1973], results
found in literature can be used to model the staple (region 3 till the exit) part of the reed
as a smoothly divergent turbulent flow, in particular in terms of the pressure recovery.
Pressure recovery is usually quantified in terms of a recovery coefficient C P stating
the relation between the pressure difference between both ends of the diffuser and the
ideal pressure recovery which would be achieved if the flow was stopped without losses:
pout − pin
CP = 1 2
(5.10)
2 ρuin

CP values range from 0 (no recovery) to 1 (complete recovery, never achieved in practice).
Distributed losses along the reed are neglected. Appendix D provides a discussion on
the magnitude of the distributed losses.
5.10 Static flow model 93

Experimental data for the diffuser In engineering literature, CP is found to depend mostly
on the ratio between output and input cross-sections (AR = Sout /Sin ) and the diffuser
length to initial diameter ratio (L/din ) [White, 2001]. The tapering angle θ influences
the growth of the boundary layers, so that above a critical angle (θ = 8◦ ) the flow is
known to detach from the diffuser walls considerably lowering the recovered pressure.
An in-depth study of turbulent flow in conical diffusers can be found in the literature
[Azad, 1996] usually for diffusers with much larger dimensions than the ones found in
the double reed.
The geometry of the conical diffuser studied in [Azad, 1996] can be compared to the
one studied in our work: the cross-section is circular and the tapering angle θ = 3.94 ◦
is not very far from the tapering angle of the reed staple θ = 5.2◦ (in particular, both are
situated in regions of similar flow regimes [Kilne and Abbott, 1962] as a function of the
already mentioned AR and L/din ). Reynolds numbers of his flows (Re = 6.9 × 104 ) are
also close to the maximum ones found at the staple input (Re ' 104 ).
The length to input diameter ratio of the reed staple L/d = 20 is bigger than that
found in [Azad, 1996], however the pressure recovery coefficient can be extrapolated
from his data to find the value CP ' 0.8 (fig. 2 in [Azad, 1996]).

Diffuser start It can be discussed whether the first part of the duct (between the blades)
also behaves as a diffuser, implying pressure recovery. The cross-section increases
towards the exit in this part as well as in the staple, however the width and height
have opposite evolutions: while the height grows continuously till the staple, the width
decreases.
Since no references were found for such a complex geometry, two limit cases can
be studied, one where diffuser is supposed to begin at region (2), so that the pressure
recovery is highly dependent on the reed opening because the flow velocity at the input
of the diffuser varies with reed opening (u1 S1 = u3 S3 , S3 being constant). The other case
supposes that there is no pressure recovery between (2) and (3), the pressure recovery
being calculated using the flow velocity at the staple input. CP will be the same for
both models, because CP does not seem to increase greatly for L/d ratios above 15
[White, 2001].

Experimental measurements of the pressure recovery

In order to check for the recovery model used for the reed staple, the pressure inside
the reed (' p2 ) was measured simultaneously to the applied mouth pressure. These
measurements were performed in a plastic reed and the reed pressure was measured at
a distance of 1 cm from the reed tip, which is far enough from the tip for the reattachment
to take place upstream.
No diaphragm was used for these measurements, so that pr = 0 for all mouth pres-
sures (pm ). The pressure recovery is then simply −p2 . However, due to the absence of
diaphragm, the reed is free to oscillate, so that only a small range of pressures can be
measured in a static regime (see section 4.2.2). Beyond the maximum mouth pressure
shown in figure 5.21 the reed oscillated.
94 Chap. 5: Details of the flow inside the reed

0.2

0.1

Reed (2) pressure (kPa)


0

−0.1

−0.2

−0.3

−0.4

−0.5
−1 0 1 2 3 4
Mouth pressure (kPa)

Figure 5.21.: Measured reed pressure in region 2 (p2 ) as a function of the applied mouth
pressure (pm ). Data is acquired twice, for increasing and decreasing pm

Figure 5.21 shows the observed reed pressures plotted against the applied mouth
pressures. The standard deviation of the measured reed pressure is plotted as a vertical
error bar, allowing to recognize the transition between laminar and turbulent regimes.
p2 is negative throughout the whole range of pm , indicating a pressure recovery in the
staple. A transition of flow regime is found at pm ' 0.4 kPa. Simultaneously, the trend
of the relation between pm and p2 is seen to change. The approximately linear relation
between pm and p2 is lost, until at about pm = 1 kPa, a new trend starts with smaller
slope.
Although neither the flow velocity nor the volume flow were directly measured, the
volume flow can be determined in the reed tip (region 1) using the pressure drop (p m − p2)
and correcting it for Vena Contracta effects using equation
 2
1 q
pm − p 2 = ρ [1 − 2CD (1 − CD )] (5.11)
2 CD S 2

Using this volume flow, we can determine the average flow velocity at the beginning of
the diffuser, and compute the pressure recovery in the staple using eq. (5.10). These
computed values are compared to the measured data in figure 5.22.
In the literature, little is discussed about the variation of the recovery pressure for
different input flow velocities, however this is an important factor in the double-reed
because the flow velocity can vary from 0 to a few tens of m/s. In figure 5.23 we
show the evolution of CP according to the Reynolds number of the flow at the staple
input (Re = u2νd2 , where d2 is the diameter of the staple input calculated using the
cross-section area S2 = π (d2 /2)2 )
5.10 Static flow model 95

0.1
measured
0 without VC
Recovery pressure (kPa) with VC
−0.1

−0.2

−0.3

−0.4

−0.5

−0.6
−1 0 1 2 3 4
Mouth pressure (kPa)

Figure 5.22.: Pressure recovery computed using flow models compared to the measured
pressure recovery. Two extreme CP values are shown and the diffuser is
assimilated to the staple using CP = 0.8

1.5
Recovery coefficient (C )
P

0.5

−0.5

−1
0 1000 2000 3000 4000 5000
Reynolds number

Figure 5.23.: Pressure recovery coefficient (Cp = 1p/22−ρupr2 ) in the double reed as a function
2
of the Reynolds number calculated at the beginning of the conical diffuser
(Re2 = u2νd2 ).
96 Chap. 5: Details of the flow inside the reed

Obviously, CP cannot be greater than 1, so that there is a small factor affecting our
calculations of volume flows, especially for the strongest pm s. However, in the turbulent
region, CP seems to have a value near 0.7. For higher values of pm , the problem is
probably the missing correction for the opening area data (see sect. 5.9.1).

5.10.4. Complete model

The Vena contracta effect can be put together with the data for the conical diffuser, to
build a complete model of the static flow between region 1 and the mouth.
In both models there is a region of the reed where the Backus model (sect. 2.4) can
be applied: the reed tip. The volume flow is once again calculated using the Bernoulli
theorem in region (1):
 2
1 q
pm − p 1 = ρ (5.12)
2 CD S
and the reed opening is determined by the pressure difference between both sides of the
reed tip:
(∆p )1 = pm − p1 = ks (S0 − S) (5.13)

Combining both equations, the expression for q(∆p)1 (the reed tip characteristics is
obtained:
(∆p)1 2
 
2CD
q2 = (∆p)1 S0 − (5.14)
ρ ks

The effect of CD on the reed tip characteristics is simply to scale the curve on the
q-axis (fig. 5.24).

0.00025
0.0002
0.00015
0.0001
0.00005

5000 10000 15000 20000

Figure 5.24.: Non-linear characteristics at the reed tip (q(∆p )1 ) according to equation
(5.14) (blue – CD = 1; red – CD = 0.5)

Diffuser assimilated to the whole reed

The total pressure drop in the reed requires the addition of the pressure recoveries due
to the jet expansion from (1) to (2), and in the diffuser, from region (2) until the reed
exit.
5.10 Static flow model 97

The complete pressure drop is obtained using equations (5.11), (5.10, with pin = p2
and pout = pr ) and the conservation of mass:
 2
1 q
2CD2 − 2CD + 1 − CD2 CP

pm − p r = ρ (5.15)
2 CD S
which is none other than the multiplication of a constant value by the reed tip pressure
drop (∆p )1 . In this model, pressure recoveries thus scale the non-linear characteristics
along the p-axis when compared to the reed tip characteristics (fig. 5.25).

0.00025
0.0002
0.00015
0.0001
0.00005

5000 10000 15000 20000

Figure 5.25.: Non-linear characteristics (q(∆p )r ) considering the full reed as the diffuser
with a pressure recovery coefficient CP = 0.8 (blue: CD = 1; red: CD = 0.5;
dashed curves show q(∆p )1 for comparison)

Diffuser assimilated to the staple

In this model, the pressure recovery due to the jet reattachment is the same. The
formula used for the pressure recovery in the conical diffuser is also basically the same
as in the previous model (with a same value for CP ), but the velocity used to calculate
it (u3 ) is now determined at the staple entrance. This allows the pressure recovery
to be considerably reduced when the reed is almost closed: even if flow velocities are
large in the reed tip, the mean flow velocity is small at the diffuser input, because the
cross-section ratio between (2) and (3) becomes very small. In this case p in = p3 and
pout = pr :
S2
 
1  q 2 2 2
pm − p r = ρ 2CD − 2CD + 1 − CD CP 2 (5.16)
2 S S3
It is worth noting that in this case the coefficient relating (∆p)1 to (∆p)r is no longer
constant. The diffuser effect is considerably reduced when the reed is nearly closed.
This fact is useful in explaining the displacement of the non-linear characteristics
maximum value towards lower pressures, as observed in the measurements of chapter
4.

Reed pressure in the beginning of the staple

Using the models shown above, the recovered pressure can be calculated based on
measured volume flows shown in figure 4.9, either considering the diffuser as the reed
98 Chap. 5: Details of the flow inside the reed

0.00025
0.0002
0.00015
0.0001
0.00005

5000 10000 15000 20000

Figure 5.26.: Non-linear characteristics (q(∆p )r ) considering the staple as the diffuser
(blue: CD = 1; red: CD = 0.5; dashed curves show q(∆p )1 for comparison)

staple:
 2
CP q
prec = ρ (5.17)
2 S3
or considering it to extend from the reed entrance until the exit (fig. 5.27):
 2
CP q
prec = ρ (5.18)
2 S2

In the last case, the reed cross-section area S2 is considered to be the measured opening
area of the reed (also called S), which varies along the characteristic curve.

Pressure recovery
20
staple recovery
reed recovery

15
pressure recovery (kPa)

10

0
−10 0 10 20 30 40 50
pressure difference (kPa)

Figure 5.27.: Pressure recovery in the reed, based on data from figure 4.9 comparing
recovery in the staple and full-reed recovery

The pressure recovery determined using the experimental data can be added to the
total pressure drop in the reed to find the pressure drop between the mouth and the
5.10 Static flow model 99

reed entrance (pm − p2 = pm − pr + prec neglecting Vena Contracta for now (CD = 1). This
corresponds to what should be measured if the reed pressure probe pr was placed on
the reed. In figure 5.28 we thus show the measured reed characteristics as a function
of the total pressured drop (∆p)r in blue, and two extrapolated reed pressure drops
(∆p )2 = pm − p2 , determined using eq. (5.18) (in red) or eq. (5.17)(in green). This plot
should be considered opposite to the plot of figure 5.26, where the original curve is the
reed pressure drop (∆p )2 , and the overall pressure drop (∆p)r is compared to it.
If the staple is considered as the diffuser, a shift of the maximum-flow pressure
(pT , the maximum of the characteristic curve) is observed towards the high pressures,
whereas the beating pressure (pM , the pressure at which the flow stops decreasing)
remains almost constant. If the whole reed is considered as the diffuser, a similar shift
is observed for pT , but pM undergoes an exaggerated increase.
Assimilating the conical diffuser to the reed is thus appropriate to explain the shifting
of pT observed for the non-linear characteristic curves measured in double-reeds when
compared to a standard Backus-like model.

Pressure recovery
0.35
measured
staple recovery
0.3 reed recovery

0.25

0.2
flow (l/s)

0.15

0.1

0.05

0
−10 0 10 20 30 40 50 60
pressure difference (kPa)

Figure 5.28.: Non-linear characteristic data from figure 4.9 compared to the reed pres-
sure calculated using recovery in the staple and recovery in the full reed

Another interesting analysis is to use the calculated reed pressure to extrapolate the
real stiffness of the reed. In figure 5.29 we show plots of the reed opening area as a
function:

• of the total pressure drop in the whole reed (pm − pr )

• of the pressure drop between the mouth and underneath the reed blades:

3 if pressure is recovered only in the staple (eq. (5.17))


100 Chap. 5: Details of the flow inside the reed

3 if the pressure is recovered along the whole reed (eq. (5.18))

If considering the pressure drop only at the reed entrance, the curves should approach
the stiffness curve measured with blocked flow for a cane reed in appendix F.

−6 Pressure recovery
x 10
5

Opening area (mm2)


3

2 measured
staple recovery
reed recovery
1
no flow

0
−10 0 10 20 30 40 50 60
pressure difference (kPa)

Figure 5.29.: Stiffness measurements compared: reed opening as a function of the full-
reed pressure drop, reed-pressure drop considering a staple diffuser and
reed-pressure drop considering a full-reed diffuser. Stiffness measure-
ment with blocked flow is shown for comparison

This comparison is delicate however, because the experimental conditions are highly
variable between the blocked-flow elasticity measurement and the non-linear charac-
teristic measurement. For instance, in the latter case the conditions inside the artificial
mouth guaranteed a saturated environment in humidity whereas the blocked-flow mea-
surement was effectuated in dry air for a dry reed.

5.11. Conclusion on the static flow model

Figures shown in the preceding section serve as a test for the models of the flow inside
the reed. It is clear from figure 5.21 that some recovery is present in the conical part of
the reed duct (the staple).
The inaccuracy of the measurements of the reed opening make it difficult to propose
a value for CP based only on the measurements of mouth pressure pm and pressure
before the staple p2 . However, based on results from other sources [Azad, 1996] and on
estimations on our measurements (fig. 5.23), a value of CP ' 0.8 seems reasonable.
As for the extent of the conical diffuser, comparison of the reed opening deformation
against static pressure in the absence of flow with the deformation during the character-
5.11 Conclusion on the static flow model 101

istics measurements (fig. 5.29) indicate a better agreement if the diffuser is considered
as the whole reed.
Since this data is highly dependent on the environment conditions, and the humidity
conditions are probably very different, it is difficult to rely on this comparison. However,
another indicator is the comparison of the corrected characteristic curves to the basic
model of the reed. Figure 5.28 provides some clues about which part of the reed acts
as a diffuser. In fact, the curve corresponding to the staple assimilated to a diffuser
brings the characteristics nearer to the classical model described in chapter 2. In
particular, the maximum of the characteristic is attained at close to 1/3 of the reed
beating pressure.
On the other hand, figure 5.28 reveals that considering not only the staple but also
the reed acting as a diffuser has too dramatic consequences on high pressures. It is
therefore sensible to propose that in fact the simultaneous converging and diverging
character of the upstream reed prevents the pressure recovery in this part.
Another important aspect is that the experiment results fit better to a model where
the Vena Contracta coefficient (CD ) in near 1. There are thus no corresponding singular
losses in the reed, although some contraction can exist, followed by pressure recovery.
This agrees with the fact that a change in the sharpness of the blades’ tips does not
have significant consequences in their behavior
It is important to notice that if the pressure recovery would be complete, the total
pressure drop inside the reed (∆p)r would equal zero, which would mean that the reed
would not be able to control the flow into the instrument (in a quasi-static regime). This
is known as d’Alembert’s paradox. In such a case the oscillations would have to be
controlled by dynamic effects in the reed. The double-reed is nearer to this limit than
the clarinet, (in which there is no pressure recovery), so that dynamic effects may be
more important than in the clarinet.

Role of the diffuser in the exciter It was already stated in the beginning of this work
(sect. 2.2) that it can be difficult to establish a boundary between the exciter and the
resonator, in particular for double-reeds.
A general view of the two last chapters reveals this ambiguity. The double-reed was
studied as the detachable part of the instrument, including both the oscillating reed
blades but also part of what can be considered the conical resonator. The remaining
resonator is almost a perfect continuation of the reed staple3 , so that the flow expansion
continues throughout the resonator.
It is thus not clear whether the diffuser should be included in the quasi-static de-
scription of the exciter (where no dynamic effects of the flow were so long taken into
account) or in the acoustical description of the resonator.
Information about the boundaries between the two elements of the instrument must
be sought in the properties of the flow along the whole instrument. It is clear that in
the tip of the reed the usual description of an acoustical waveguide is not valid: the
3
Tapering decreases from 5.2◦ in the staple to 2.9◦ in the resonator, but the discontinuity in cross-section
is minimal [Nederveen, 1998]
102 Chap. 5: Details of the flow inside the reed

duct cross-section undergoes wide variations through time and flow velocities become
too important for the linear approximation of acoustics to be valid.
The issue of flow velocities can be addressed by plotting the distribution of flow veloc-
ities inside the reed (fig. 5.30): when these become comparable to the velocity of sound
non-linear effects need to be taken into account for instance, shock waves arise along
the propagation. Figure 5.30 shows that non-linear effects are important at the tip of
the reed, especially if we take into account that the cross-section data was measured
for a fixed opening of the reed whereas for the volume flow used in this plot the reed
can be almost half-closed, increasing the flow velocity at the tip to around 200 m/s (see
fig. 5.18).

0.4

0.35

0.3
Mach nbr.

0.25

0.2

0.15

0.1

0.05
0 10 20 30 40 50 60 70
axial distance (mm)

Figure 5.30.: Distribution of the Mach number of the flow along the reed (tip is to the
left), calculated from the maximum volume flow observed at the reed out-
put (qout = 0.3 l/s, fig. 4.9), and the cross-section profile S(z ) (fig. 3.1)
qout
Ma (z ) = cS (z )

In the remaining path the flow velocity is never above 0.25c, and this value is reached
for a short length (compared to the wave length). The linear acoustic approximation
should thus be applicable throughout the whole staple.
One additional condition for the resonator acoustic equations to be valid is that the
flow must be homogeneous on a wavefront (a spherical surface in the case of a conical
resonator). The symmetry of the flow at the staple output suggests that this condition
is fulfilled at some point in the staple, at least in the static regime.
Should these conditions be met also for oscillating flows, the reed staple, or at least
part of it, could be considered as part of the resonator. Nevertheless, time variations in
the flow introduce many differences that will be investigated in the next chapter.
In the clarinet, an ambiguity also exists as to the position of the boundary between
exciter and resonator, because the turbulent mixing region is not well-defined. Yet, a
mistake in the coupling point (between the exciter and the resonator) can be corrected
by adding a length correction to the resonator, because the resonator is straight. In
the case of a conical resonator, in addition to the length correction, the mathematical
formulation of the flow expansion in a conical duct leads to different results if the
acoustical propagation is used rather than the quasi-static flow model.
5.11 Conclusion on the static flow model 103

In conclusion, the effect of the conical diffuser was useful in this discussion to explain
the measured characteristic curves, and allowed to show that the upstream part of the
reed behaves as a classical reed exciter. The effect of its inclusion in the quasi-static
model of the reed will be tested in sound synthesis, but it should be kept in mind that
in dynamic conditions the role of the diffuser may not be modified by the flow dynamics.
104 Chap. 5: Details of the flow inside the reed
6. Auto-oscillations in the double-reed
In this section, some aspects of the behavior of the double-reed in dynamic conditions
are introduced. In a brief discussion, we try to identify the more relevant changes that
need to be introduced in a static model of the reed. After some remarks on observed
time-evolutions of the reed motion, measurements of the flow profiles during oscillations
are presented and discussed.

6.1. Introduction
Most of the issues discussed in the preceding chapters were studied in the framework
of quasi-static conditions, where time-derivatives of the properties being studied are
negligible.
One particular aspect that changes this assumption was already discussed in section
B.1: the mass and amping of the reed material, prevent it to adapt instantaneously to
a change in the force applied to it.
In the flow there are similar aspects that may need to be taken into account. No such
effects were observable in chapters 4 and 5 because all experiments were performed in
such a way that time-variations of the properties could be neglected.
However, when time variations of the flow properties become important, there are
several aspects that are changed. In particular, some terms in the equations of fluid
motion (in particular Navier-Stokes and mass conservation) that could be neglected
until now may play an important role when describing the functioning of the instrument
during auto-oscillations, because there are several properties that change very rapidly
in time.
Some features that have been left aside for now and that require further investigation
are discussed in the next section. We will then present a closer look into the double-reed
and its properties during oscillations in order to discuss some of these dynamic issues.

6.1.1. Overview of the chapter


The chapter starts (sect. 6.2) with a discussion about the dynamic effects that were
neglected in the basic model of the double-reed presented in chapter 2 and that were
avoided in the experiments in the previous chapters.
In section 6.3 we present some observations of the periodic reed motion when play-
ing an oboe, and discuss some features present in these observations in the light of
observations from other authors.
Finally, section 6.4 extends the measurements of the velocity profiles of chapter 5 to
dynamic regimes: the time evolution of the flow profiles is observed and discussed.
106 Chap. 6: Auto-oscillations in the double-reed

6.2. Discussion on dynamic flow effects


Given the impossibility to fully extend the study of the double reed to account for
dynamic effects in the reed and flow, this section overviews some of the most important
effects that were neglected in the previous presentations and experimental studies of
the double-reed model.
The presentation will run through the terms of the fluid motion equations that were
not considered in previous discussion, trying to present some hints about their mag-
nitudes. Some of this discussion was also presented in more theoretical grounds in a
previous article [Vergez et al., 2003].

6.2.1. Effects of inertia in the air flow

The most obvious effect (deliberately) forgotten in previous discussions, is the inertia of
the flow. Similarly to what was seen for the reed, the air also presents some resistance
to change its motion due to variations in internal and external forces.
This fact is represented in the Navier-Stokes equation as the only term involving time-
derivatives. The importance of inertial effects in a dynamic flow is usually evaluated
through an non-dimensional number, the Strouhal number St = ωl u
, comparing the
relative magnitude of the time variations of the flow ( ∂∂tu ) to the advective term ((u · ∇)u)
in the NS equation.
In the clarinet, an evaluation is made based on the length of the reed and the velocity
at the reed input [Hirschberg et al., 1990]. The obtained magnitude St ' 0.1 suggests
that stationary effects can be neglected in this case.
In the double-reed, a similar analysis can be made using data from previous sections.
An estimation for the oscillating flow velocity is obtained from figure 6.6 (which will be
introduced later in sect. 6.4) umax ∼ 200 m/s (using the mass-conservation equation
to extrapolate a value for the velocity at the reed input u1 = SSout 1
uout ). These values are
confirmed from the static case in figure 5.18. However, these values should be used
with caution, because incompressibility may not be valid in oscillating regimes.
l = 10−2 cm is taken as the oscillating length of the reed, and an average playing
frequency ω = 2πf = 3000 Hz is used. These values yield an upper-bound for the
Strouhal number St = 1.5. This value indicates that the time-derivative term in the
Navier-Stokes equation is very likely to be important in the description of the dynamic
flow in the double-reed.

6.2.2. Compressibility

In previous chapters, the air inside the reed was supposed to always have a constant
density. In static flows, this hypothesis is valid if the flow velocities are small compared
to the velocity of sound, which is expressed by another non-dimensional number, the
Mach number M = uc .
In the static case, the maximum flow velocities found inside the reed approach 200
m/s at the reed tip for low reed openings (cf. fig. 5.18). The Mach number found at
6.2 Discussion on dynamic flow effects 107

these velocities (M ' 0.6) indicates that compressible effects may need to be taken into
account for small reed openings (before viscous effects become important).
In dynamic regimes, it is clear that variations of density in the oscillating flow are
important in the bore, in fact they are necessary to describe acoustic wave propagation.
The question of knowing if compressibility effects should be included in the reed model
is thus connected to the ambiguity in the separation between the exciter and resonator.

6.2.3. Fluid-structure coupling

The model described in chapter 2 supposes that the valve controlling the flow is punc-
tual, and that its motion does not induce an air flow inside the reed. In the real reed this
clearly cannot be the case because an important area of the reed moves during normal
oscillation, pushing or aspirating part of the air underneath it.
Usually, this issue is addressed by adding to the flow expressed by equation (2.25)
a term corresponding to the flow ‘‘pumped’’ by the reed [Barjau and Agulló, 1989]. In
other works this flow is included in the model as a correction in the resonator impedance
(a cavity in parallel with the main bore) [Thompson, 1979].

Estimation of the pumped flow

An estimation of the pumped flow can be obtained simply by using data about the reed
opening (fig. 6.1) to calculate volume variations inside the reed. The present estimation
will be restricted to the time-intervals where the ‘‘pumped flow’’ should be maximum,
that is when the reed transits from closed to open.
It is seen in section 6.3.3 that the reed velocity is maximum at this instant: the
reed opening area changes from 0 to 2 × 10−6 m2 in an interval of about 10−4 seconds,
corresponding to a cross-section variation at the tip of 2 × 10−2 m2 /s.
If we suppose the reed to bend along a length of 1 cm, which is roughly the length
of the scraped area at the tip (see sect. 1.3) the pumped flow can achieve a maximum
value of 2 × 10−4 m3 /s. This is roughly equivalent to maximum flows observed in static
regimes when measuring the reed characteristics (fig. 4.9).

Model of fluid-structure coupling

Flow induced by the reed movement is usually considered to correspond to the volume
swept by the reed during its closing or opening motion. However, the effect of the air
inertia and incompressibility is usually not taken into account, so that the reed motion
is not affected by the presence of the air inside the reed. This can be particularly
problematic when the reed is about to close, because the air between the two blades
cannot escape instantaneously from between the two reed blades.
This fact was taken into account in different physical models involving the fluid-
structure coupling, such as the glottis [Deverge et al., 2003] and previously for a similar
coupling in industrial machinery [Antunes and Piteau, 2001].
108 Chap. 6: Auto-oscillations in the double-reed

6.3. Periodic motion of the reed


In section 3.3 the shape of the reed entrance was observed for a full range of reed open-
ings. It was already mentioned that the observed shapes were approximately maintained
during the auto-oscillations of the reed.
Using a stroboscope, the images of the reed can be recorded in slow motion. Using a
recording of the sound produced by the instrument, it is easy to know the period of the
reed motion, the same as the period of the sound.
The images can be analysed automatically using the procedures described in appendix
H, and the dimensions of the reed opening extracted for each image. This allows for
instance to plot the time-evolution of the reed opening area during auto-oscillations.

Temps (s)
−6
0x 10 1 2 3 4 5 6
3.5 7

3 6

2.5 5

Largeur (m)
2 4
Aire (m )
2

1.5 3

1 2

0.5 1

−4
x 10
0 0
0 1 2 3 4 5 x 106
−3
Temps (s) −3
x 10

Figure 6.1.: Time evolution of the reed opening area, determined by image analysis of a
video recording. Red plots the area and blue the distance between blades h

A simultaneous plot of the opening area and the distance between blades is shown in
figure 6.1 for a natural cane oboe reed coupled to the bore is shown without lips or lip
replacement.

6.3.1. Two-state movement


In all observations of the reed oscillations, the reed spends a part of its motion in
a completely closed state, in which the tips of both reed blades touch all along the
opening length. This behavior is different from that of single reed instruments where,
at least for piano regimes, the reed can oscillate in a non-beating regime (this is true
both in mathematical models and in the real instrument [Gilbert, 1991], although in
the clarinet the shape of the mouthpiece lay introduces a further obstacle to the reed
closure, so that the real reed is more different from the linear oscillator model).
Piano notes are difficult to obtain with a double-reed instrument played with the
artificial mouth. This probably explains in part why double-reeds have a narrower
dynamic range than other reed instruments (see sect. 1.4). In fact, if the reed can only
6.3 Periodic motion of the reed 109

oscillate in beating regimes, its amplitude of oscillation cannot be as small as if the


reed would be able to oscillate between two different opening states, and this is maybe
a reason for the advices given by experts to increase the lip pressure and approach the
lips to the reed tip when playing in a piano dynamics, as described in section 1.4.3.

6.3.2. Time relations between open and closed states


The duration of the closed state can be a small fraction of the whole period. Again,
this is unlike what happens in beating regimes of the clarinet, where the closed state is
usually very close to one half of the period. Moreover, neither the duration of the closed
state nor its ratio to the period of the motion remain constant when the breath pressure
changes, an observation also made by [Gokhshtein, 1979].
This fact seems to be more associated to the conical resonator (which was used in
our experiments) than to the double-reed, if we relate this fact to the ideal Helmholtz
motion in conical waveguides. To convince ourselves of the role of the conical resonator
in this fact, it useful to remind that this asymmetry in the time-evolution of the reed
opening (and internal pressure wave) is observed also in saxophones.
[Benade, 1988] proposes that the conical resonator can be in many ways assimilated
to a cylindrical resonator where the exciter (for example the reed) is connected not to
one of its ends, but to a derivation of the tube perforated along the tube. In order that
the similarities are maximized, this derivation should be situated at a distance L tr (the
truncation length of the cone) from one end, and L (the conical resonator length) from
the other one.
In ideal conditions, where visco-thermal and radiation losses can be neglected, a
single-reed coupled to such a resonator can be shown to describe a Helmholtz motion
where the open state and the closed state have different durations. On the other hand,
a cylindricl resonator where the reed is coupled to one of the ends always has a perfectly
symmetrical Helmholtz motion.
As for the variations in the closed time-interval in the reed motion, a slight change
can be justified by the action of visco-thermal losses in the resonator, which smooth
the square pressure waveform of the ideal Helmholtz motion. The beating pressure is
situated somewhere between the two pressure states of the Helmholtz motion, and since
the transitions between these two states are no longer instantaneous, a slight change
should occur as the minimum pressure state approaches the beating pressure.
However this can only account to small changes in the variation of the duration of
the closed-state.

6.3.3. Closed to open transition


The transition time between closed and open states can be estimated as the time in-
terval between the beginning of the reed motion and the first peak (fig. 6.1). This is
approximately 0.1 ms.
There are two main reasons why the transition between both states is not instanta-
neous:
110 Chap. 6: Auto-oscillations in the double-reed

• The driving force does not change instantaneously from one value to another.
In fact as discussed above, this would only be true in an ideal Helmholtz mo-
tion, which is destroyed by the visco-thermal losses in wave propagation in the
resonator.

• The inertia of the reed, implying that a change in the force is not followed instan-
taneously by a change in the reed position.

If we neglect the first effect and suppose the pressure wave to be equivalent to that
of a Helmholtz motion, the force acting on the reed at the transition is a step force. A
damped linear harmonic oscillator responds to such a force oscillating at its resonance
frequency. The time between the end of the closed-state and the first peak is thus 1/2
of a period.
In the case of the oboe reed, from sect. B.1 it can be estimated as 3200 Hz, so that
the time corresponding to half a period is 21fr = 0.16 ms. This is close to the estimated
transition time based on the figure, 0.1 ms.
Obviously, this effect must be combined with the non-instantaneous transition in the
pressure wave, but the reed inertia is seen to play an important role in this transition
time.

6.3.4. Open to closed transition


Transitions from open to closed state seem slower than their counterparts from closed
to open state. Gokshtein [Gokhshtein, 1979] reports that the double reed sightly opens
before shutting closed. This seems to be confirmed in several observations of the oscil-
lation of the oboe reed. Brown [Brown, 1990] relates this opening to the arrival of the
incoming pressure wave (p − ).

6.3.5. Oscillations during the open state


During the open state, the reed oscillates several times around an average value without
completely closing (see figure 6.1). There are two possible reasons for these oscillations:
the fact that the reed is itself an oscillator upon which a sudden force is applied when
the pressure wave returns from the resonator and the oscillations of the pressure wave
itself (because the resonator is not ideal and some pressure reflections might arrive
before the main pressure wave reflected at the end of the bore).

6.3.6. Effect of biting


Although our experiments didn’t cover a large enough sample of reeds in order to gen-
eralize our conclusions, we did study the effect of the lips (or its replacement in the
artificial mouth) on the oscillations of the reed (fig. 6.2).
Lips (or their replacements) prevent large openings of the reed. In fact measurements
on a reed played with lips see the height of the initial peak (as the reed opens) decreased,
but further oscillations still happen although their shape is slightly different.
6.4 Dynamic flow characterization 111

−6
x 10 −6
0x 10
3 2.5

2.5
2

2
1.5
Area (m2)

Area (m2)
1.5

1
1

0.5
0.5

0 0
0 1 2 3 4 5 6 0 1 2 3 4 5 6 7 8
Time (s) −3 Time (s) −3
x 10
x 10

Figure 6.2.: Time variation of the reed opening area without and with lips

Moreover, the first valley in the reed opening graphic after the transition is similarly
deep in the both cases, which can be an indication that the oscillations during the
open-state are not an effect of the reed’s mechanics. In fact, if the motion is attenuated
while the reed is opening, it should not have enough energy to decrease its opening as
much as when it is not attenuated.
Of course, this analysis is different if the material used to replace the lips is rigid
and not shock-absorbing. In this case, the reed is prevented to open but its energy is
recovered when moving out of lip contact. However, this does not seem to be the case
when manipulating the artificial lip, which seems to be absorbing.

6.4. Dynamic flow characterization


Velocity profiles presented in chapter 5 were effectuated while reed vibrations were
artificially prevented, by rigidly fixating them to adjustable screws avoiding the auto-
oscillating effects induced by flow instabilities. However, these auto-oscillations are part
of the normal functioning of the reed, when it is coupled to the instrument’s bore.
In this section the static condition of the reed is released, and the time-evolution of
the profiles will be analysed and related to the profiles measured in the previous section.
There are several difficulties in trying to measure the profiles at the reed output when
it is coupled to the instrument resonator. First of all, the resonator is a rather long
structure. An extension of the hot-wire probe could be considered, but given the length
of the resonator, it would be difficult to ensure that this extension is rigid enough to
keep the probe fixed at a constant position during each measurement. The solution
would be to replace the real instrument resonator by a simplified model that can be
perforated in order to place the probe inside of it. This however was not tried out.
The second difficulty is that the probe dimensions (the hot wire is about 1 mm long)
are comparable to the duct dimensions (the reed output is about 5 mm in diameter) so
that it would only be possible to measure a part of the diametrical length of the staple
output, and special care would be necessary when arriving at the walls, due to the
fragility of the hot-wire probe.
112 Chap. 6: Auto-oscillations in the double-reed

All the profiles were thus measured using an isolated reed (uncoupled to any res-
onator), at the immediate vicinity of the reed output, except for the longitudinal profile,
for which a part of the measurements were taken inside the reed.
Note that the reed oscillations are of much higher frequency than these at which the
instrument is played in normal regimes, so that unstationary effects in the flow are
exaggerated when compared to a normal operating regime. Also, the fact that the flow
terminates in open air introduces several modifications with respect to a bounded case,
such as dragging some of the air in the surroundings of the reed output.

Figure 6.3.: The experimental setup used for flow measurements output from the reed,
and a detailed view of the measurement region: (1) Reed, mounted on the
artificial mouth (2) Electrostatic microphone (3) hot-wire probe (4) Temper-
ature probe (5) Stroboscopic lighting (6) Camera (7) Micro-metric actuators
for the probe position.

6.4.1. Procedure

The hot-wire probe measures the absolute value of the velocity in a precise region. The
phase lag between measurements at two different positions is thus unknown. However,
since the flow is periodic, a simultaneous measurement of a variable that remains the
same throughout the measurements can be used to re-synchronise the signals taken at
each position (described in appendix I).
This is why, simultaneously to the velocity measurements a pressure recording is
taken always at the same position. This signal is used to calculate the time lag to be
introduced in each signal so that they are in phase. The delay between two different ve-
locity measurements is calculated using the correlation between the respective pressure
recordings.
For each run (a series of velocity measurements at different positions with a same
parameter configuration – pressure and reed biting), a video recording of the reed open-
ing is done along with a recording of the reference pressure signal. This recording is
6.4 Dynamic flow characterization 113

performed with a high-speed camera synchronised with stroboscopic lighting. The fre-
quency of acquisition of the images in this video recording is not related to the frequency
of the reed motion. However, simultaneously to each image, an impulse is recorded in
another channel of the acquisition card. This sequence of impulses is used during the
analysis to rearrange the images in a periodic signal (see appendix I).
Due to a problem in the synchronisation between some of the channels in the acqui-
sition card that were discovered after the experiments, it is not possible to assure that
these impulses are in-phase with the velocity and pressure measurements.

6.4.2. Experimental conditions


In this experiment a synthetic reed in FiberCane (from Glottin) was used, uncoupled to
any resonator.
The mouth pressure was fixated at 14 kPa, which is situated nearer to the extinction
pressure (15 kPa) than to the threshold of oscillation (about 9 kPa). In these conditions,
the reed squeaks at strong sound levels.
Two measurements were made, one without lips and another with a lip replacement
consisting of a synthetic foam pressed against the reed.

6.4.3. Obtained data


The configuration of the hot-wire probe allows to measure the absolute value of the
velocity, but it does not give any information about the direction of the flow. This
means for instance that incoming and outgoing flow with the same velocity value will
be measured as a same value.
In some cases the flow direction can be deduced from other characteristics of the
flow. For example, when the direction of the flow changes, this can usually be noticed
by a tight bend of the graphic near the horizontal axis (near v = 0). This bend is not
exactly a discontinuity in the derivative because it is possible for the flow to turn (change
direction instead of simply the sense), which is reflected as a minimum value slightly
above 0.

Frequency of oscillation

Playing the reed without the bore is a test also performed by the musicians when fine-
tuning the scraping of the reed. This is called a ‘‘tuned squeak’’ [Agulló, 2001].
The oscillations have a frequency of 1208 Hz for the experiment without lips and 1336
HZ with the lips replacement. In other experiments with different mouth pressures and
without lips which are not shown here, the frequency is always situated between 1200
and 1300 Hz.
These frequencies are very different from the mechanical resonance frequencies of the
double-reed (around 3200 Hz, see sect. B.1).
This difference can be understood by noticing that the reed channel (including the
staple) is also a small, 7 cm long resonator (check figure 3.1). The frequency of the
first mode can be determined using an algorithm that splits the reed duct into conical
114 Chap. 6: Auto-oscillations in the double-reed

segments and calculates the impedance curve. The first mode can be found at 1530 Hz,
slightly above the frequency of the tuned squeak (see figure).

9 Impedance modulus (linear)


x 10
2

Modulus (linear)
1.5

0.5

0
0 1000 2000 3000 4000 5000 6000
Frequency (Hz)

Impedance phase
2

1
Phase (rad)

−1

−2
0 1000 2000 3000 4000 5000 6000
Frequency (Hz)

Figure 6.4.: Impedance curves calculated for the reed duct (reed and staple)

The frequency of the ‘‘tuned squeak’’ does not only depend on the mechanical fre-
quency, nor on the acoustical frequency of the reed channel. However both factors
influence the squeak: shortening the staple raises the playing frequency and a stronger
lip pressure has the same effect.

Evolution of diametrical profiles

Measurements of the time-evolution of the profile are shown in figure 6.5. They show
ten evenly spaced instants within one period of the reed motion. Several periods of
velocity measurements were used to calculate average profiles and standard deviation
from the average profile.
Due to the fact that only velocity modules are measured, it can be difficult to dis-
tinguish the flow leaving the reed from the flow that enters it in figure 6.5. There are
however some indications on the image that may help in distinguishing them:

• In the last 4 frames, there is a plateau of velocities formed between the reed walls.
This can be related to the jet observed in static conditions in section 5.4. Since
the velocity is measured at a short distance from the reed output, it is in fact
likely that the jet limits are very well defined. The jet profile is different from
those observed in section 5.6.1, which is not surprising, because there are now
additional dynamic effects.

• In the same frames, the plateau is affected by stronger fluctuations of the velocity.
Again, in section 5.4 it was observed that the jet is turbulent for practically all
6.4 Dynamic flow characterization 115

t/T= 0.01 t/T= 0.11

20 20

10 10

0 0
0 5 t/T= 0.21 10 15 0 5 t/T= 0.31 10 15

20 20

10 10

0 0
0 5 t/T= 0.41 10 15 0 5 t/T= 0.51 10 15

20 20

10 10

0 0
0 5 t/T= 0.6 10 15 0 5 t/T= 0.7 10 15

20 20

10 10

0 0
0 5 t/T= 0.8 10 15 0 5 t/T= 0.9 10 15

20 20

10 10

0 0
0 5 10 15 0 5 10 15

Figure 6.5.: Diametrical velocity profile evolution throughout one period (average value
shown as the central blue line and standard deviation shown as the red
area around the average profile). Position in mm (x-axis, reed boundaries
shown as dash-dotted vertical lines); flow velocity in m/s (y-axis)
116 Chap. 6: Auto-oscillations in the double-reed

blowing pressures used in the static regime. The same is true for the dynamic
regime, where the pressure drop in the reed (∆p)r rises up to 2pM .

• The first four frames show that the flow is distributed around the staple walls,
inside but also outside the staple. Such a distribution is not possible when the
air is leaving the duct, because it would form a jet rather than diffusing quickly
into the environment. This suggests that airflow is negative (directed towards
the reed output) in the first 4 frames. It is worth recalling that, at least in the
Helmoltz oscillations of the reed there should be no negative velocities (which
would correspond to larger pressures inside the reed than in the mouth), so that
the inversion of the flow is probably due to unstationary effects in the flow or flow
induced by the reed motion.

• In frames 5 and 6, a strong peak of velocity rises near the reed walls, before the
formation of the jet. The sense of the flow in this peak cannot be deduced at this
point.

Evolution of axial profiles

Following the axis of revolution of the reed, it is possible to acquire the velocity values at
some points inside the reed (until the staple becomes too narrow). Figure 6.6 shows the
time evolution of the flow velocity measured on the axis of the reed, simultaneously to the
evolution of the reed opening area (due to synchronisation problems in the acquisition
card it is not possible to assure that the opening area signal is well-synchronised with
the velocity measurements).
The flow velocity is always positive outside the reed (for x < 4mm), and its oscillations
are weak.
Within the reed the velocity shows much stronger oscillations. The space-time surface
of the flow evolution shows a kink (between 0.5 and 0.8 seconds) which, due to the fact
that it is surrounded by a valley of low values, corresponds very likely to negative-valued
velocities. This agrees with the negative velocities in the first frames of figure 6.5.

Influence of damping

It was seen above (sect. 6.3.6) that the lips avoid wide oscillations of the reed, in partic-
ular they tend to decrease the overshooting of the reed during the opening phase.
By using a similar arrangement for the lips, it is possible to check if the reduction in
the reed amplitude of oscillation has or not noticeable effects on the flow evolution, in
which particular moments, and in which regions of the flow output.
The measurement of section 6.4.3 was repeated with the introduction of a soft and
damping material (sponge material used in earplugs) encircling a region close to the tip
of the reed.
The oscillating regime for the uncoupled reed is different from that shown in figure
6.2. In particular, the oscillating frequencies are much higher in the present experiment,
which is likely to increase the importance of dynamic effects. . However, the transitions
6.4 Dynamic flow characterization 117

−2

60 0
velocity (m/s)

2
40
4
20
6
0
8
1.6
1.4 10
1.2
1
0.8 12
0.6 position (mm)
0.4
0.2 14
0
time (msec)

Figure 6.6.: Time variation of the average axial profile, with standard deviation shown
as color scale — from blue (low) to red (high). The reed tip is represented
towards the observer and the staple output is placed at 6 mm. The time-
variation of the reed opening area is represented in the back wall of the
graphic.

between open and closed states have comparable durations (about 15 × 10 −5 sec for
the closed to open transition in figure 6.2 against 7.5 × 10−5 for figure 6.7), and this
fact may be used to draw some general conclusions about the flow induced by these
transitions.
It can be observed (fig. 6.9) that the artificial lips have the expected effect of reducing
the oscillation amplitude of the reed from 2.3 mm2 to 0.9 mm2 (peak values of the reed
opening area without and with lips). The period of time during which the reed remains
closed is also slightly increased.
The general profile evolution is not substantially changed by the addition of lips. If
the profiles are synchronised using the time-evolution of the flow velocity on the axis,
the localised velocity peaks that were found in sect. 6.4.3, appear slightly later in the
case of the reed played with lips. However the peak value is approximately the same.
In section 6.2.3, the pumped flow in the oscillations of a reed without the action of lips
was estimated to peak at 2 × 10−4 m3 /s. These values should remain the same in the
present case. In fact, even though the reed oscillates at a much higher frequency than
in the case of figure 6.1, the durations of the transitions remain basically the same, as
can be verified by comparing the former figure with figure 6.9.
If this flow is distributed over the reed output cross-section (2 × 10−5 m2 ), the average
flow velocity induced during transitions would thus be 10 m/s.
In the case with lips, this value should be reduced to less than 1/2, because the reed
velocity is reduced of this value.There should thus be a difference of at least 5 m/s at
118 Chap. 6: Auto-oscillations in the double-reed

t/T= 0.01 t/T= 0.11

20 20

10 10

0 0
0 2 4 6 8 10 12 14 0 2 4 6 8 10 12 14
t/T= 0.21 t/T= 0.31

20 20

10 10

0 0
0 2 4 6 8 10 12 14 0 2 4 6 8 10 12 14
t/T= 0.41 t/T= 0.51

20 20

10 10

0 0
0 2 4 6 8 10 12 14 0 2 4 6 8 10 12 14
t/T= 0.6 t/T= 0.7

20 20

10 10

0 0
0 2 4 6 8 10 12 14 0 2 4 6 8 10 12 14
t/T= 0.8 t/T= 0.9

20 20

10 10

0 0
0 2 4 6 8 10 12 14 0 2 4 6 8 10 12 14

Figure 6.7.: Diametrical velocity profile snapshots at the output of an auto-oscillating


reed: in blue, without lips (data from figure 6.5) an in green, with lips.
6.4 Dynamic flow characterization 119

15 25
without lips without lips
with lips with lips

20

10
Flow velocity (m/s)

Flow velocity (m/s)


15

10

0 0
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1 0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1
Time (fraction of period) Time (fraction of period)

Figure 6.8.: Comparison of the time evolution of the flow at two key-points in the dia-
metrical profile. On the left, the flow velocity in the axis (data from both
runs has been synchronised by hand in order to obtain similar velocity evo-
lutions in the axis of revolution of the reed), and on the right, flow velocity
near the reed walls, using the same synchronisation as for the left-hand
figure.

some point (at 2 different times) in the flow output. This value would be even more
important if the pumped flow were localised at the axis or borders of the reed output
instead of being distributed over the output cross-section.
Such variations in the flow velocity should be observed in the analysis of the velocity
profiles, however differences between the time-variations of the flow velocity are consid-
erably underneath the expected variations in the pumped flow. We thus conclude that
no trace of the pumped flow is found at the output of the reed.

6.4.4. Sketch of the flow evolution


The deductions presented in the former sections allow us to sketch a possible flow
evolution at the reed output, as shown in figure 6.10.
The flow can be divided into two phases, an injection phase, where air is flowing from
the outside into the reed at the output section, and a ejection phase where flow is mainly
directed outwards.
In the injection phase, the flow can be related to that of a sink. It is being sucked
into the reed and thus is arriving from several directions. We recall that based only
on a quasi-static analysis, there should never be a flow entering the reed, because the
pressure drop in the reed (∆p)r oscillates roughly between 0 and 2pm . One possible
reason for these negative flows can be the flow ‘‘pumped’’ by the reed motion during the
transition from closed to the open phase.
The main difference from a circular sink flow is that the velocity is higher near the
walls than at the center. This is probably due to fluid inertia, which causes some of the
outgoing flow from the previous phase to keep trying to escape while part of the flow is
120 Chap. 6: Auto-oscillations in the double-reed

2.5

Undamped
Damped
2

Reed opening area (mm )


2
1.5

0.5

0
−0.5 0 0.5 1 1.5 2
rescaled time (fraction of period)

Figure 6.9.: Comparison of the opening area evolution between the case without lips
and with lips

Figure 6.10.: Sketch of the flow at the reed output for three phases of the reed oscillation.
From left to right, frame indications relative to the profiles in figure 6.5:
incoming (1-4), edge flow (5-6), and outgoing jet (7-10).
6.4 Dynamic flow characterization 121

already being forced to enter by the change in the conditions at the reed tip. Similar
behavior has been observed in acoustic flow at the termination of pipes [Atig, 2004],
where for strong amplitudes there can be the formation of vortexes in the vicinity of the
pipe walls.
The outgoing flow starts with very pronounced peaks of outgoing flow near the reed
boundaries. Again this might be related to the inertia of some incoming flow that may
still be blocking the axis of the duct. The fact that outgoing velocity peaks are sharper
and more pronounced than the incoming peaks is probably a consequence of the reed
walls, which force the flow to come from a privileged direction, whereas in the case of
the incoming flow, the flow may be distributed along several directions.

6.4.5. Conjectures on the flow in a coupled reed


The observations made in this section refer to an uncoupled reed. Some boundary effects
would not be observed in a reed that is coupled to the resonator, and the frequency of
vibration would be considerably lower.
The issue of different playing frequencies can change the influence of unstationary
flow effects. For example, if transitions between incoming and outgoing flows is slower,
it is likely that flow output and input is more evenly distributed over the output cross-
section of the reed.
However, it was observed in section 6.3.1 that the transitions between open and
closed state are effectuated over a short period of time, when compared to a whole
period of motion. In fact the transition time of 0.15 ms estimated for the reed opening
observations of figure 6.1 is similar to the transition times (0.2 ms, approximately half
of the total opening time) when the reed oscillates uncoupled from the resonator, such
as in figure 6.6.
There is not however, an equally fast succession of opening and closing transitions
as there exists in the uncoupled case. It is possible that the flow has time to stabilise
(become homogeneous over the reed output cross-section) between each transition, but
the same flow inertia problems will probably be observed at lower frequencies as well.
The prolongated boundary conditions in the coupled case also affect the distribution
of the flow: for instance, during the injection period, flow is constrained to the reed
walls, so that a similar effect to that observed during the transition between incoming
and outgoing flow (frames 5-6 in fig. 6.5) could take place.
It is difficult to say if these effects are purely boundary effects or if there is a layer
of flow near the walls travelling in the opposite direction of the flow throughout the
resonator, during the transitions between incoming and outgoing flows. Simulations
by M. Atig [Atig, 2004] seem to show that the formation of vortexes is in fact an effect
induced by the termination of the resonator, and that layered flows are not observed
inside the resonator.
122 Chap. 6: Auto-oscillations in the double-reed
7. Retrospective on the experimental results

This chapter resumes the most important results of the experimental part of the thesis. For
each topic a list of values for important parameters is given. Finally, we conclude by
describing the retained experimental facts to be used in a double-reed model.

7.1. Summary of part II


Throughout the four chapters of this part, several aspects of the reed were characterised
experimentally. The aim of this chapter is to collect the main obtained results and to
group them in thematic sections. For each result the appropriate section is referred
where the extended discussion can be found.
A list of parameters identified and measured within the section is given after the
discussion, along with the associated values or ranges of values.
Most of the results are available for a few double-reed samples, although it is not
always possible to guarantee the musical usability of the reed.
The presented results apply primarily to oboe reeds. In certain cases they were also
verified in the bassoon reed.

7.1.1. Geometrical aspects of the reed opening


The double-reed opening or aperture has an oblong shape of an eye. This general
shape is maintained as the distance between blades changes, and the length of the reed
opening does not vary considerably (sect. 3.3). Some reeds may present an extra opening
at rest because they do not touch along the full length of the rails before pressure is
applied (sect. 3.3.2).
A single variable can be used to parametrise the reed shape at all instants in models
intended for musical regimes (sect. 3.3.5).
The reed opening area is approximately proportional to the distance between blades
(sect. 3.3.3). This allows to propose an effective length as the proportionality constant.
These results should be checked for different fabrication procedures, in order to test
different reed opening shapes.
The shape remains symmetric (sect. 3.3.4) throughout the range of reed openings. In
some reeds an asymmetry is observed but given the indications of musicians (sect. 1.3)
this should be considered as a defective reed.
During reed oscillations, the shape and the symmetry of the reed is maintained (sect.
3.3 and 3.3.4). The two blades are seen to oscillate in phase (against each other) when
playing notes (coupled to the instrument) and also in tuned squeaks (observed in images
from section 6.4).
124 Chap. 7: Retrospective

Useful parameters

Effective length leff ' (5.2 ± 0.2) × 10−3 m

7.1.2. Elastic properties

The reed opening is found to be mostly proportional to the pressure difference applied
between the inside and the outside of a synthetic reed (sect. 3.4). For these reeds, a
stiffness of 6 × 109 Pa m−2 is found.
In natural reeds the elasticity curve is slightly non-linear, showing a stronger stiffness
for smaller openings (this can be seen in the graphic of figure 5.29). In wet reeds, the
measured stiffness varied between 4 × 109 Pa m−2 for large openings and 14 × 109 Pa
m−2 for small openings.
Evidence of visco-elastic effects was found on the reed material (sect. 3.5). When
increasing the pressure until the closure of the reed and decreasing back to 0, the
return curve is displaced from the original one. The final reed opening is smaller than
the initial one. After the pressure is released, an exponential return to the original
opening is observed, with characteristic times of 900 seconds in the synthetic reed and
30 seconds in the natural reed. Moreover a faster relaxation is observed in the synthetic
reed (τ = 2.5 s) but not in the natural reed.
These effects are supposed to be responsible for the long-term variations of the sound
quality in experiments with the artificial mouth (oscillation frequency, intensity, timbre),
but are liable to be balanced by the gestures of the musician in musical performances

Useful parameters

Stiffness Found to be constant in a synthetic reed, with a value ks = 6 × 109 Pa m−2 ,


variable in a natural reed between 4 × 109 Pa m−2 for large openings and 14 × 109
Pa m−2 for small openings

7.1.3. Vibrational properties

The mechanical response of the reed is complex, revealing several modes of vibration
(sect. B.5.1).
The first modes are found to have a resonance frequency between 1300 and 1800 Hz
followed by a stronger resonance at 2400-2700 Hz in the bassoon reed. Wetting the
reed reduces the resonance frequency.
The quality factors of the resonances are not easy to measure in the response curves.
Wet reeds show higher quality factors than dry reeds (fig. B.3). A synthetic reed shows
a much higher quality resonance than the natural reeds (fig. B.4), which reveals a less
damped material.
In the oboe, resonances are found 3500 Hz and 4200 Hz.
7.1 Summary of part II 125

Useful parameters

Mechanical resonance frequency For the oboe ωr ' 3500 Hz, for the bassoon ωr ' 1300−
1500 Hz.

7.1.4. Non-linear characteristics


Measured non-linear characteristics curves in the double reed show that the behavior of
the double-reed is qualitatively similar to what is expected from the Backus reed model
(sect. 4.3.2).
However some quantitative differences are observed in the ratio between the maximum-
flow pressure (PT ) and the beating pressure (PM ). In the Backus model, this ratio is
expected to be ( PPMT = 3) whereas in the double-reed this value varies between 4 and 5
(section 4.4). These differences are later explained by the pressure recovery in the staple
(sect. 5.10.4).
The characteristic curves measured when decreasing the pressure is displaced from
that measured for increasing pressures. This hysteresis is considered to be related
to the viscoelastic properties of the reed found during the measurement of the elastic
properties of the reed.
For a same reed used in similar experimental conditions, the characteristic curves are
found to have different scalings (sect. 4.3.3). The shape of the curves in different mea-
surements does not vary considerably, so that the spread in the scalings is considered
to be due to uncontrolled parameter variations such as the reed opening at rest.

7.1.5. Flow inside the reed


For most of the range of mouth pressures (above at most 0.3 kPa) used in experiments
and in playing conditions, the flow at the reed output is found to be turbulent (sect.
5.4). Similarly, inside the reed, turbulence is found above 0.5 kPa (fig. 5.21).
Measured velocity profiles at the reed output without reed oscillations show a constant
shape for variable mouth pressures pm and reed openings h, at least in turbulent
regimes (sect. 5.6).
These velocity profiles are almost axially symmetric (sect. 5.6.3). When comparing two
perpendicular diameters (parallel and perpendicular to the reed opening), differences in
the value of the velocities are smaller than 11%, at least for large reed openings. We
conclude that turbulence is able to erase the initial asymmetry in the flow caused by
the geometry of the reed duct.
Comparing the profiles to well-known pipe flows shows that the reed flow is more
closely related to a cylindrical pipe than to a conical diffuser (sect. 5.7). However,
measurements of the pressure inside the reed (sect. 5.10.3) show that there is a non-
negligible pressure recovery that is supposed to be induced by the conical geometry of
the staple.
In section 5.10.4 we use the pressure recovery model used in engineering for coni-
cal diffusers to explain the shifting of the maximum-flow pressure PT observed in the
measurements of the characteristic curves. This allows us to use a Backus-like model
126 Chap. 7: Retrospective

for the reed static behavior, if the staple is not considered to be part of the exciter.
Nevertheless, due to the continuity between the staple and the resonator there is an
ambiguity as to where the separation between exciter and resonator should be placed.
The Vena Contracta at the reed entrance has an effect of scaling the characteristic
curves along the flow (q) axis. The contraction coefficient CD is negligible (close to 1) for
a good agreement with experimental results.

Useful parameters

Pressure recovery CP ' 0.8

Vena contracta CD ' 1

7.1.6. Dynamic regimes

During its motion when coupled to a resonator, the reed observes a two-state movement,
oscillating between a completely closed state and an open state (sect. 6.3.1). This is
observed for several playing frequencies (determined by the resonator length or fingering)
and mouth pressures, in opposition to the oscillation of the clarinet reed (which can
oscillate in beating or non-beating regimes depending on the mouth pressure).
The open-state lasts longer than the closed-state (sect. 6.3.2). The duration of the
closed-state is conserved when the playing frequency changes. The open-state is thus
shorter when the playing frequency increases. Some variation of the duration of the
closed-state is observed for different mouth pressures.
The transition times between open and closed states are short when compared to
the period length (sect. 6.3.3). The open-to-closed transition is slightly longer than
the closed-to-open transition (sect. 6.3.4), but transition times are not affected by the
presence or absence of the resonator (compare figure 6.1 with fig. 6.9). The duration
of the closed-to-open transition has the same magnitude as the half-period of a reed
resonance.

7.1.7. Dynamic flow

In section 6.4 the velocity profiles at the reed output were measured during tuned
squeaks of an oboe double-reed (periodic oscillating regime without resonator).
The flow alternates between an outgoing jet and an incoming sink flow near the walls
(fig. 6.10). The formation of the jet is accompanied by a strong velocity region near
the staple walls, for which the direction is not clear, and cannot be explained with the
present data.
Some of these aspects are likely to change when the reed is coupled to the resonator.
However, since the transitions between open and closed regimes have similar durations
in both coupled and uncoupled cases, similar effects should be found with the resonator,
like incoming flow through the walls simultaneously to some outgoing flow in the center,
or the strong peaks observed during the formation of the jet.
7.2 Retained facts for a double-reed model 127

Comparing two oscillations showing different reed motion velocities, we conclude that
no trace of the flow ‘‘pumped’’ by the reed is found in the velocity profiles (sect. 6.4.3).
Results for dynamic regimes (both reed motion and flow evolution) require further
investigation in order to propose explanations for most of the described features. More-
over the flow dynamics results should be compared to the resonator-coupled case so
that boundary effects do not affect the measurements and the playing frequencies are
brought closer to usual musical regimes.
A complete model of the reed should take into account both unstationary effects in
the flow and the fluid-structure coupling in the tip of the reed.

7.2. Retained facts for a double-reed model


The experimental study presented in this part of the thesis allows us to extract some
facts for a double-reed model. When applicable, relevant parameters were listed that
serve as a starting point for the tuning of the model. Since most of these parameters
are affected by the interaction with the musician, they need to be tuned to arrive to a
perceptively good numerical model of the instrument. In some cases it is expected that
the time-evolutions of these parameters are as important as their static magnitudes.
This opens a perspective for further investigations on the double-reed.
The results obtained in characteristics measurements suggest that the Backus model
used for other reed instruments can be used for double-reeds with some precautions.
The ambiguity yet unsolved relative to the boundaries between the exciter and the res-
onator opens some questions as to whether the pressure recovery should be considered
in the exciter. In practice this effect can be included in the instrument model, with a
tunable parameter CP controlling the magnitude of this effect.
Similarly, the effect of Vena Contracta and subsequent reattachment can be included
in the model leaving the contraction coefficient CD as a tunable parameter, although
its effects did not prove to be necessary in the description of the static model of the
double-reed.
128 Chap. 7: Retrospective
Part III.

Towards a numerical model of a double


reed instrument
8. Synthesis of double reed instruments sound

In this chapter we introduce a synthesis algorithm for the simulation of double-reed


instruments. The methods used to build the digital components of the model are
described, and the physical parameters are linked to the measurements of the previous
part. A short demonstration of the results of the simulation is given, as well as a comparison
between flow models proposed in chapter 5

8.1. Introduction
A physical description of the instrument through mathematical equations allows to
predict its behavior through time. The model presented in chapter 2 already provides a
set of equations than can be used to describe a reed instrument. In particular, using
correct discretisation methods for this set of equations opens the way to simulations of
the behavior of the instrument and the sound produced by it.
This approach has been used since more than two decades now to build a new kind of
synthesis method which is known as ‘‘Physical modeling synthesis’’. This kind of syn-
thesis presents some advantages with respect to other synthesis methods, for instance,
it can predict changes in the timbre due to parameter variations. The timber evolves
naturally as different dynamic ranges are used, or different notes are played. Moreover,
the input parameters to the model are closely related to the control exerted on the real
instrument, rather than the spectral parameters used for example in FM or additive
sound synthesis.
A physical model does not need to be very complicated in order to reproduce the
general behavior and timbre evolution of an instrument. However, the finer details of its
sound can only be precisely modelled through a deep investigation of all the factors that
can be involved in sound production. This is the main disadvantage of Physical Modeling
Synthesis when compared to a database of recorded sounds (Wavetable synthesis). In
the other hand, the timbre possibilities of an instrument are infinite, so that a sound
database can never mimic all kinds of sounds produced by an instrument.

8.1.1. Historical background


The first works on physical modelling synthesis [MacIntire et al., 1983] proposed a time-
domain simulation of the resonator coupled to a generic non-linearity relating the two
variables of acoustic propagation. This relation was initially not based on the physics of
the instrument. However some of the effects inherent to a physical production of sound
were already present, such as the spectral enrichment or the period doubling associated
to non-linear system dynamics.
132 Chap. 8: Synthesis

Other developments have followed that allowed the evolution of physics-based sound
synthesis. Digital waveguides, on which the following synthesis model is based, were in-
troduced in sound synthesis by [Smith, 1986], and evolved under the works of [Hirschman, 1991]
or [Välimäki, 1995]. Later, [Scavone, 1997b] successfully applied this technique for con-
ical waveguides.

8.1.2. A digital model of a reed instrument

The digital instrument model that will be presented in this chapter is based on the
generic reed model presented in section 2. It is built in a modular way, so that it can
evolve according to improvements brought to the physical model. Another advantage
is that it is suitable to be applied to different reed instruments by adapting the exciter
parameters or the resonator models. In fact, this model is a basis to the work developed
by the author for a saxophone model in the framework of the partnership between
IRCAM and Arturia.
In what concerns this thesis, the main interest of a physical model is to test the
perceptive relevance of the physical phenomena discussed in the previous section.
Eventually, the model can be enriched with effects not yet taken into account, and
may be able to mimic the behavior and timbre of a real double-reed instrument. With
proper tuning of the parameters and a complex work to map the input parameters of
the model to a small set of parameters used by the user, this model will be suitable for
musical applications.

8.1.3. Overview of the chapter

The chapter starts with the detailed description of the methods used to discretise the
equations of the double-reed model (sect. 8.2).
In section 8.3 we discuss the application of the parameters discovered using the
experimental data of part II. The control of the time-evolution of these parameters is
briefly discussed in section 8.4.
Finally we present some results of the simulations of an oboe and the effect of some
of the physical parameters of the model

8.2. Digital model

In this section, the model proposed in chapter 2 is transformed into a set of discrete-
time formulas suitable to simulate a reed instrument in a computer. Some additions
are made to the model of the exciter proposed in section 2.4, taking into account the
static flow model proposed in section 5.10.
8.2 Digital model 133

8.2.1. Implementation of cylindrical resonators


It was seen in section 2.3 that the propagation of an axial sinusoidal plane wave along
an acoustical wave-guide can be described by:

p± (z, t ) = e (∓k (ω)L ) p± (z0 , t ) (8.1)

with L = z − z0 .This is a consequence of equation (2.1). In general, and due to the


visco-thermal losses during the propagation, k (ω) depends in a non-linear way on the
frequency. The proposed model for these losses is given by equation (2.7), which for
now can be written as:
ω
k (ω) = − V (ω) (8.2)
c
This allows us to rewrite equation (8.1) as
ω
p± (z, t ) = e (∓V (ω)L ) e (∓ c L ) p± (z0 , t ) (8.3)
ω
The factor e (∓ c (z −z0 )) states simply that the wave at position z is a delayed copy of the
wave at position z0 , the delay interval being L/c:

L
p± (z, t ) = e (∓V (ω)L ) p± (z0 , t ∓ ) (8.4)
c

The remaining factor e (∓V (ω)L ) is more difficult to translate into a time-domain descrip-
tion.

Diffusion and dispersion of the waves

The above term describing visco-thermal losses is expressed in the frequency domain
as:
i 3/2
ηcω1/2 L )
Vf (ω) = e (∓V (ω)L ) = e (± 2 (8.5)

It is the non-linear dependence in ω that makes an exact time-domain description of


Vf (ω) impossible.
In practice, a direct discrete time filter with 3-poles and 3-zeros is approximated using
the least-square approximation algorithm provided by the MatLab function invfreqz.
Different approaches based on series expansions of the non-linearity or Padé approxi-
mations may allow to build a parametric model, but these were not attempted.

Resonator termination and radiation

For a simple resonator termination as described in section 2.3.4 it is also possible to


approximate formula (2.10) by the first terms of the polynomial expansion (because
ka ∼ 1).
However, a simple resonator termination does not usually provide a good approxima-
tion for the radiated wave, so that the whole bell should be taken into account together
with the termination impedance.
134 Chap. 8: Synthesis

Since bells are usually of small dimension in woodwinds, a time-domain description


of the bell together with the radiation impedance does not require too many coefficients
for a discrete-time filter, so that we judged appropriate to build a single filter for the
whole system. The coefficients are determined using the same approximation algorithm
used for the visco-thermal loss filter.

8.2.2. Implementation of conical resonators


The models shown in section 2.3.6 for conical resonators are based in spherical travel-
ling waves. The propagation of these waves can be described in a similar way as was
done in section 8.2.1 for the plane waves (eq. (8.4)), with the necessary correction for
the expansion of the spherical wavefront:
z L
p± (z, t ) = e (∓V (ω)L ) p± (z0 , t ∓ ) (8.6)
z0 c
A filter for visco-thermal losses is built upon the same formula as was used for the
plane waves (eq. (8.5)), using an optimised radius corresponding to the radius at 1/3 of
the conical resonator length. The method used to discretise the filter corresponding to
visco-thermal losses remains the same as described in section 8.2.1.
The difference between cones and cylinders is thus localised at the input and output
of the bore, where pressure p and flow q need to be transformed into spherical waves
using the relations:
    
P (ω, z ) 1 1 P − (ω, z )
 =   (8.7)
1
Q (ω, z ) Zc (ω,z )
− Z ∗ (1ω,z) P + (ω, z )
c

where Zc (ω, z ) is the characteristic impedance of the spherical wave at a distance z from
the center of the spherical wavefronts (eq. 2.20).
The ω-dependence of Zc and Zc∗ introduces the need of filters to transform (p + , p− )
into (p, q) in the time domain. Th implementation of these filters will be seen below
(sect. 8.2.5).

8.2.3. Other elements of the bore


The two different kinds of resonators described in sections 8.2.1 and 2.3.6 are too
simple to model a complete realistic resonator such as that of an oboe or bassoon. One
way to approximate to the real resonator is to assemble pieces of conical or cylindrical
resonators with different angles and/or diameters end to end. In the next section we
will describe how to describe the junction between to pieces of resonator where the
travelling wave description changes.

Junctions between resonator elements

In this section we present a generic method for joining two resonator elements (conical
or cylindrical) which may or not have discontinuities in their cross-sections or tapering
8.2 Digital model 135

angles. By introducing linear relations between the acoustic variables on each side of
the discontinuity (equation (8.12)), other elements can be introduced between the two
parts of the resonator, such as tone holes or cavities.
p1− p2−
T1

R1 R2

T2

p1+ p2+

Figure 8.1.: Scattering junction describing the variable names used for the reflection
(Ri ) and transmission (Ti ) coefficients.

The main assumption used to calculate the scattering matrix is the conservation of
flow (Q1 = −Q2 ), where a positive sign means that the air is arriving at the junction.
The pressure is also conserved (P1 = P2 ).
The scattering matrix S relates the outputs of the lumped system to its inputs. The
outputs and inputs are expressed as travelling waves, whose characteristic impedances
do not need to be the same.
      
P1+ R1 T2 P1− P1−
 =   , S  (8.8)
P2− T1 R2 P2+ P2+

The travelling waves are related to the oscillating pressure P and flow Q through the
Travelling Wave Matrix W:
   
P P−
  = W  (8.9)
Q P+

In the case of spherical waves, the characteristic impedance of the travelling waves
is:
P − (z ) ρc 1
Zc (z ) = − = 1 (8.10)
Q (z ) S(z ) 1 + ikz

which depends on the distance z to the center of radiation.


The travelling wave matrix W becomes:
    
P 1 1 P−
 =   (8.11)
1 1
Q Zc
− Z∗ P+
c
136 Chap. 8: Synthesis

The scattering junction is usually described in terms of the relations between the
acoustical variables pressure P and flow Q on each of its ends:
      
P1 A B P2 P2
 =   , Z  (8.12)
Q1 C D Q2 Q2

In order to calculate the scattering matrix S we write first the relationships between
the travelling waves on each end of the scattering junction which is:
   
P1− P2−
  = W−1 ZW   (8.13)
P1+ P2+

The scattering matrix is then calculated by solving equation 8.13 for P1− and P2+ .
For spherical waves on both ends of the scattering junction, we can write the elements
of the scattering junction matrix defined in equation (8.8) as:

Zc∗1 Zc∗2 (B − DZc 1 + (A − CZc 1 )Zc 2 )


R1 = (8.14)

Zc 2 Zc∗2 (Zc∗1 + Zc 1 )
T1 = (8.15)

Zc 1 Zc 2 (B − AZc∗2 + Zc∗1 (D − CZc∗2 ))
R2 = (8.16)

Zc 1 Zc∗1 (AD − BC)(Zc∗2 + Zc 2 )
T2 = (8.17)

with
∆ = Zc1 Zc∗2 (B + AZc2 + Zc∗1 (D + CZc2 )) (8.18)
and Zc 1 = Zc (z1 ) and Zc 2 = Zc (z2 ) the characteristic impedances on each side of the
scattering junction.

Discontinuities in the resonator

In our description of the resonators we studied two particular cases: cylinders (sect. 8.2.1)
and cones (sect. 2.3.6). A real instrument usually has an almost cylindrical (e.g. the
clarinet) or conical (e.g. the oboe) shape throughout most of its length. However, they
usually contain some parts which diverge from this regular trend, such as the bell or the
mouthpiece. Arbitrary shapes can be approximated by joining together small segments
of regular shape, for which the propagation is well known.
Simple junctions of resonator elements (i.e. where there are no other acoustic ele-
ments in between) can be described using a particular form of the scattering matrix
(sect. 8.2.3), where the Impedance Matrix (Z) of equation (8.12) has the form:
 
1 0
Z=  (8.19)
0 1
8.2 Digital model 137

which means that the flow an the pressure are the same on the extremities of the res-
onators in the junction. This relation is also true in first approximation for diameter
discontinuities, although there are some models [Karal, 1953] which propose an addi-
tional impedance due to the formation of higher order modes near the discontinuity.
This correction tends to zero as the ratio of the diameters approaches 1.
The scattering coefficients of equations (8.14-8.17) can then be written:

Zc∗1 Zc∗2 (−Zc 1 + Zc 2 )


R1 = (8.20)
Zc 1 Zc∗2 (Zc 2 + Zc∗1 )
Zc 2 Zc∗2 (Zc∗1 + Zc 1 )
T1 = (8.21)
Zc 1 Zc∗2 (Zc 2 + Zc∗1 )
Zc 1 Zc 2 (−Zc∗2 + Zc∗1 )
R2 = (8.22)
Zc 1 Zc∗2 (Zc 2 + Zc∗1 )
Zc 1 Zc∗1 (Zc∗2 + Zc 2 )
T2 = (8.23)
Zc 1 Zc∗2 (Zc 2 + Zc∗1 )

In order to implement a time-domain model of the scattering junction, we rewrite the


scattering junction coefficients in terms of their Laplace transform:

s/c (1 − σ ) + z11 − σ
z2
R1 = (8.24)
s/c (1 + σ ) − z11 + σ
z2
2s/c
T1 = 1 σ
(8.25)
s/c (1 + σ ) − z1
+ z2

s/c (−1 + σ ) + z11 − σ


z2
R1 = (8.26)
s/c (1 + σ ) − z11 + σ
z2
2s/cσ
T2 = 1 σ
(8.27)
s/c (1 + σ ) − z1
+ z2

where σ = r22 /r12 is the ratio between the resonator sections at the junction and s the
Laplace variable. 1
It is interesting to notice that the cone distances to the apex (zi ) always appear in
denominator for the coefficients of each filter. If we notice that zi−1 = tanri θi , this formula
can also be used for cylindrical resonator segments, for which θi = 0.
Stability of a filter is assured if the poles of its Laplace transform (equations (8.24–
8.27)are situated
 in the left half of the Laplace plane. All the filters have one pole at
σ 1 c
s = z2
− z1 σ +1
. The second factor in this equation is always positive, so that the
condition of stability resumes to σ
z2
< z11 . Written it in terms of the conicity angles,
tan θ1 < rr12 tan θ2 .
1
One can check that as expected these formulas can be inverted, that is T2 and R2 can be obtained
2
r2
respectively from T1 and R1 by replacing z1 → −z2 , z2 → z1 and σ = 2
r1
→ σ −1 .
138 Chap. 8: Synthesis

Discretisation of these equations is done using the bilinear transform s = iω →


1−z −1 1−z −1
2Fs 1+ z −1
= α 1+ z −1
. α is the bilinear transform coefficient which we choose to be twice
the sampling frequency (2Fs ).
The bilinear transform method [Oppenheim and Schaffer, 1989] has the advantage
of conserving the stability of the continuous system written in equations (8.24–8.27).
Coefficients for the discrete system corresponding to scattering junctions is given in
table 8.1.

R1 T1 R2 T2

A(1−σ )+1/z1 −σ/z2 (1−σ )−1/z1 +σ/z2


b0 A(1+σ )−1/z1 +σ/z2
2A
A(1+σ )−1/z1 +σ/z2
− AA(1+σ )−1/z1 +σ/z2
2A
A(1+σ )−1/z1 +σ/z2

b1 − AA(1−σ )−1/z1 +σ/z2


(1+σ )−1/z1 +σ/z2
2A
A(1+σ )−1/z1 +σ/z2
A(1−σ )+1/z1 −σ/z2
A(1+σ )−1/z1 +σ/z2
2A
A(1+σ )−1/z1 +σ/z2

a1 − AA(1+σ )+1/z1 −σ/z2


(1+σ )−1/z1 +σ/z2
(1+σ )+1/z1 −σ/z2
− AA(1+σ )−1/z1 +σ/z2
(1+σ )+1/z1 −σ/z2
− AA(1+σ )−1/z1 +σ/z2
(1+σ )+1/z1 −σ/z2
− AA(1+σ )−1/z1 +σ/z2

Table 8.1.: Coefficients of the recursive discrete-time filters used for resonator disconti-
nuities

The resonator discontinuity is used in our synthesis model to couple a short cylindri-
cal segment between the reed and the conical bore. This has two advantages: correcting
the inharmonicity inherent to a truncated cone (see sect. 2.3.6), and providing an easier
coupling between the exciter and the resonator (see sect. 8.2.5).

Tone-holes

In chapter 2, it was explained that the frequency of the sound produced by the kind
of instruments which are the subject of this thesis (oboes and bassoons) is mainly
determined by the length of the bore.
In real life, changing the length of the resonator is unpractical, especially for conical
resonators. Instrument makers found a practical way to overcome this difficulty by
opening holes in the walls of the bore, which in a first approach work have an effect
of shortening the bore length approximately to the length between the top and the first
open hole.
In practice, a complete description of the effect of the tone-holes is more complicated
than this. A discussion on this issue can be found for example in [Benade, 1976] or
[Keefe, 1982b]. Implementation of tone-holes in a discrete-time model of a wind instru-
ment has also been the issue of a number of works [van Walstijn and Campbell, 2003]
[Scavone, 1997a].
In the present work, the instrument model was not taken so far as to model tone-
holes, because the main emphasis was on the exciter. In practice, the full range of
notes in one octave is implemented by shortening the delay-line corresponding to the
propagation of the acoustic wave in the bore. For higher octaves, the register key (sect.
8.2 Digital model 139

8.2.3) helps to increase the impedance of the bore for a higher mode of vibration of the
air-column, which is supposed to be harmonic.
An important consequence of the tone-holes is that they work as an additional filter to
the propagation of the sound-wave in the bore [Benade, 1976]. The regular distribution
of the holes can be modelled approximately as a tone-hole lattice, whose consequence
on the acoustic wave inside the bore is to cut frequencies higher than a certain cutoff
frequency. The tone-hole lattice is also important in the radiation of the sound.
Nevertheless, these aspects were not implemented in the present synthesis model, so
that the final instrument is closer to a double-reed coupled to a single conical resonator
without holes.

Register holes

The method used in this synthesis model to change the playing frequency of the in-
strument (to ‘‘cut’’ the bore) provides acceptable results for the lower keys in the real
instrument.
However the whole pitch range of a double-reed instrument is about 3 octaves, which
would require for the highest notes a bore length 4 times shorter than the real length.
In a real instrument the musician can select the operating mode of the resonator, either
by changing its embouchure or mouth pressure, or by using special register keys (in
some cases, special fingerings in the tone holes have the same effect) which attenuate
the lower modes of the resonator.
While opening a tone hole greatly attenuates the standing wave in the part of the
resonator between the first open tone hole and the bell, the effect of the register hole is
to impose a pressure node at the position of the hole.
Because they are smaller in diameter, register holes are easier to implement than
tone holes: radiation can be neglected and only the inertia of the mass of air contained
within the whole and its resistance by friction against the walls need to be take into
account.
In fact, from Keefe’s tone hole model [Keefe, 1990], the series impedance is neglected.
His experiments [Keefe, 1982a] show that the series impedance (which is simply an
inertance) corresponds to a main bore length correction whose effects are masked by
the shunt impedance. The shunt impedance is then:
 2
ρc rB
Zrh = (ıkte + ξe ) (8.28)
πrB2 rh

In this formula, c represents the sound speed and ρ the density f the air. rB is the
radius of the main bore at the position of the register hole, rH the hole radius, te the
inertance parameter (proportional to the mass of the air contained in the hole) and ξe
the resistance parameter. The two last parameters are usually tuned by hand in order
to produce a correct sound for higher registers.
The register hole (as would happen with a tone hole) can be included into the model
as a special junction (see section 8.2.3) where the flow is not conserved in the two bore
parts. In fact three different flows have to be considered (fig. 8.2), the flows in the main
140 Chap. 8: Synthesis

bore (Q1 and Q2 ) and the flow leaving the instrument through the hole (QH ). Considering
the arrows in figure 8.2 as the sense of positive flows we have:

Q1 = Q 2 + Q H (8.29)

Pressure is conserved on both sides of the register hole junction.


dH

hH

QH
Q1 Q2

Figure 8.2.: Scattering junction used for a tone hole

In practice, radiation from the register hole is usually neglected (because it is very
small compared to the radiation from tone holes and bell), so that we only need to
consider the wave variables inside the bore. These can be related using the following
impedance matrix:
      
P1 1 0 P2 P2
 =   = ZT   (8.30)
1
Q1 Zrh
1 Q2 Q2

Zrh = QP2H = QP1H represents the register hole shunt impedance (the series impedance is
neglected).
Using a similar method as for a normal scattering junction, we derive the scattering
coefficients (same notations as in figure 8.1):

1
R1 (s) = R2 (s) = − (8.31)
πr 2
2Zrh ρcB +1
πrB2
2Zrh ρc
T1 (s) = T2 (s) = 1 + R1 (s) = (8.32)
πrB2
2Zrh ρc
+1

where s is the Laplace variable.


Both formulas of equations (8.31–8.32) assure the stability of the filters because the
real part of the poles is always negative.
8.2 Digital model 141

1−z 1−z −1 −1
Bilinear transform is used to discretise these filters, replacing s → 2F s 1+ z −1
= α 1+ z −1
:
c
ζ +αψ
(1 + z −1 )
R1 (z ) = R2 (z ) = − (8.33)
1 + ζζ +
−αψ −1
αψ
z
T1 (z ) = T2 (z ) = 1 + R1 (z ) (8.34)

r2 πrB2
with ψ = 2 rB2 t and ζ = 2 ρ
ξ + c. The stability of the discrete time filters is assured by
h
the stable continuous time filters and by the use of the bilinear transform.

8.2.4. Exciter model


The exciter is based on the model proposed by Backus for single-reed instruments
[Backus, 1961], and that was presented in section 2.4. Corrections to this model are
added to take into account some of the effects discussed in part II.
The mechanical part of the reed is modelled as a single harmonic oscillator, as de-
scribed by equation (2.31):

(∆p)1 (t ) ∂(∆x ) ∂2 (∆x )


= ωr2 ∆x (t ) + 2α + (8.35)
ms ∂t ∂t 2
The mechanical parameters ms , ωr2 and α are suggested by the experimental data of
chapter 3, but are left free to be adjusted by the user that can tune them in real-time to
account for the biting of the reed, for example. The contact with lips can in fact change
the effective mass and damping of the reed, and to a lesser extent, its stiffness.
The flow is described by a quasi-static model, where the flow velocity entering the
reed is proportional to the instantaneous value of the pressure drop across the reed
entrance:

1
(∆p)1 (t ) = ρu1 (t )2 (8.36)
2
The volume flow is calculated from u1 (t ) by multiplying by the jet cross-section q(t ) =
CD S(t )u1 (t ). S(t ) is, as usual, the reed opening area, and the factor CD is introduced
to account for the Vena Contracta effect. This way, the effect of thinning out the reed
blades can be tested in the synthesis model.
We thus arrive to the equation relating the volume flow q(t ) to the pressure drop
across the reed, from eq. (5.12):
 2
1 q(t )
(∆p)1 (t ) = pm − p1 (t ) = ρ (8.37)
2 CD S(t )

The pressure transmitted to the resonator is calculated from the mouth pressure pm
and the pressure drop (eq. (5.16)) in the whole reed, after the flow reattachment:
2 
S2
 
1 q(t ) 2 2
(∆p)r (t ) = pm − p(t ) = ρ 2CD − 2CD + 1 − CD CP (8.38)
2 CD S(t ) S32
142 Chap. 8: Synthesis

This equation supposes that the staple can act as conical diffuser, however the pressure
recovery can be regulated by the parameter CP , in order to test for the perceptive effects
of the pressure recovery. The total pressure drop in the reed (∆p)r (t ) is related to the
pressure drop in the tip (∆p)1 (t ) by the relation:

S2
 
2 2
(∆p)r (t ) = 2CD − 2CD + 1 − CD CP (∆p)1 (t ) (8.39)
S32

Discrete-time model of the exciter

A numerical simulation of the whole instrument requires the discretisation of the exciter
model as well as the resonator (seen in sect. 8.1.2). For equations (8.37) and (8.38) this
consists simply in replacing t by tn = nT , where T = f1s is the discretisation interval.
Because it describes a non-instantaneous relation, the discrete resolution of equation
(8.35) requires the choice of a discretising scheme, so that it can be re-written as a finite
difference equation:
N
X M
X
ai xn −i = bj pn −j (8.40)
i =0 j=0

where xn = (∆x )(tn ) and pn = (∆p )1 (tn ).


The impulse invariance method was chosen because it conserves the resonance fre-
quencies of the oscillator [van Walstijn, 2002], whereas the bilinear transform is known
to shift them as well as the Euler scheme, which moreover has the problem of increased
gains in the resonance peaks. It also has the advantage of providing xn explicitly as a
function of the past state of the reed, which simplifies the resolution of the system of
equations describing the coupling.
The coefficients ai and bi are listed in terms of the physical parameters in table.

a0 1
a1 −2e −αT cos ωr T
a2 e −2αT

b0 0
 
Te −αT
b1 ms ωr
sin ωr T

Table 8.2.: Coefficients of the discrete-time harmonic oscillator, calculated using the
impulse invariance method

The pressure drop in the tip (∆p )1 (tn ) cannot be calculated directly. The only pressure
values that are known at each sample are the mouth pressure pm and the pressure at the
input of the bore pn . It is thus necessary to use equation (8.39) to calculate (∆p )1 (tn −1 )
from (∆p )r (tn −1 ). The advantage of the impulse invariance method is thus verified,
because only the values at tn −1 are needed to calculate the reed displacement.
8.2 Digital model 143

Knowing the value of the reed opening at time tn , it is possible to calculate the opening
area (Sn ), which depends exclusively on xn .
The resolution of the system of equations corresponding to the whole instrument is
then reduced to solving equation (8.38) together with the equations of the resonator
relating p to q.

8.2.5. Coupling the exciter to the resonator


The exciters models introduced in section 5.10.4 can be described under the form of the
relation F(p, q) = 0 where F is usually a differential operator. On the other hand, the
resonator descriptions shown above are usually written under the form of G(p − , p+ ) =
0, because of the simplicity of implementation of G . The interface between the two
components requires the transformation of p and q in p + and p− which is done using
the travelling wave matrix introduced in equation (8.9).

Plane waves at resonator input

In a case where the resonator begins with a cylindrical segment, the relation between
variables (p, q) and (p+ , p− ) is instantaneous (none of the variables at time tn depends
on past values of another variable).
It is thus easy to replace one of the variables (for instance p) in equation (8.38),
according to the definition of the plane travelling waves:
pn = 2pn+ − Zc qn (8.41)
so that equation (8.38) can be solved for qn :
s
Zc Γ|pm − 2pn |
qn = sign (pm − pn ) ± 1+4 (8.42)
2Γ Zc2
In practice, this is the usual case applied to instruments, because as mentioned
in sect. 2.3.6 there is an added cylindrical segment to a conical resonator in order to
correct the inharmonicity of truncated cones.

Spherical waves at resonator input

If the resonator begins with a tapering segment, equation (8.41) is substituted by a finite
difference equation involving past values of q (because Zc is replaced by a recursive filter:
PM
j=0 bj qn −j
Zin q = yn = PM (8.43)
i =1 ai yn −i
In order to solve for qn using a similar relation as eq. (8.42), eq. (8.41) has to be
rewritten as:
M
X M
X
pn = 2pn+ − b0 qn − ( bj qn −j − ai yn −i ) = 2pn+ − b0 qn + Λ (8.44)
j=1 i =1

where Λ only depends on values determined during the calculation of previous samples.
144 Chap. 8: Synthesis

8.3. Application of experimental data


At this stage, the details of the implementation of the basic double-reed instrument are
complete. Some of the experimental data gathered throughout part II of this thesis can
be used to correctly dimension the instrument.
The value of ks was determined in section 3.4 to be ks = 6.15 × 109 Pa/m2 .
The surface mass of the reed ms can be determined using data from the mechanical
response of the reed (fig. B.5). Since the first resonance frequency is situated at about
3200 Hz, and using the surface stiffness of the reed:
r
1 ks
f = ωr = (8.45)
2π ms

The reed mass is found to be ms ' 0.08 kg/m2 .


Other parameters can be more difficult to determine from experiments. For example,
the damping coefficient α is determined through trial and error until the oscillations
induced by the reed resonance observed after the reed opens resemble what is observed
in figure 6.2 with the effect of the lips. In practice only one or one half of an oscillation
should be distinguishable.
When the damping is to small, reed squeaks overcome too easily at the attack of
a note, so that this is also taken in consideration when adjusting the reed damping.
Nevertheless the damping cannot be too high otherwise the range of pm for which auto-
oscillations exist is too narrow. In addition, for high dampings the instrument sounds
too dull and dark: perception is also taken in consideration when adjusting these
parameters.
The parameters used in section 5.10.4 to explain the measured reed characteristics
were found to be CP ' 0.8 and CD close to one. However, the role of these parameters is
not clear in the a complete model of the instrument. For this reason, they were tested
along the tuning of the model. They are discussed in the following section, along with a
short presentation of some results of the simulation.
The instrument model is implemented in a real-time software platform, which allows
to scan the infinity of combinations of parameter configurations more easily than with
a model implemented in MatLab, for example.

8.4. Control of the model


The synthesis model was implemented in a real-time synthesis environment 2 which
allows to control each parameter individually in real-time, at the same time the sound is
produced or eventually link several parameters through a set of higher-level parameters.
This is useful when using an external controller such as a Yamaha WX5 midi controller
or a keyboard with a breath controller.
For a scientific testing purpose however, it is convenient to be able to accurately
controlling each parameter in a predictable and reproducible manner.
2
PureData (http://puredata.info/)
8.5 Results of the simulation 145

One of the problems with respect to reproducibility is that for example, to start the
auto-oscillation of the instrument, usually it is not sufficient to increase the mouth
pressure from 0 to a certain pressure that is known to be able to sustain these auto-
oscillations. Similarly to what was seen about the control of real double-reed instru-
ments (section 1.4), it is often necessary to increase the mouth pressure beyond a certain
level in order to engage the oscillations, decreasing it afterwards to the sustained level
corresponding to the desired note and dynamics. This is why a reproducible envelope
system was introduced in the model, where it is possible to control the attack and decay
times as well as the overshooting of the pressure.
In the following presentation of simulation results, these envelopes were only applied
to the mouth pressure. Other parameters remain constant throughout the note, unless
stated otherwise, for instance when studying the behavior of the model for variations of
a single parameter.

8.5. Results of the simulation


Once the parameters of the model are fine-tuned, it is possible to simulate a note played
by the instrument, and to plot the time-evolution of some of its variables.
Figure 8.3 shows a plot of the evolution of the pressure at the bore input and the
reed opening, for two different values of the pressure recovery coefficient (C P ), during
the attack transitory. The permanent regime, in the sustained part of the note is shown
in figure 8.4.
Adimensioned pressure

0.5

−0.5
0 500 1000 1500 2000
time (samples)
Adimensioned reed position

C =0
1 P

CP=0.8

0.5

0
0 500 1000 1500 2000
time (samples)

Figure 8.3.: Pressure at the reed input and reed position plotted along the time, during
the attack of a note, for different values of CP

The change in CP causes a change in the range of working pressures of the instrument.
When this parameter is increased, it is necessary to increase the blowing pressure so
146 Chap. 8: Synthesis

that the oscillation can start. In the two cases, the sustained mouth pressure was
fixated just above the threshold of oscillation. The attack parameters were kept constant
(attack for 200 samples, decay for 800 samples and an overshoot of 40%).

Adimensioned pressure
0.5
CP=0
CP=0.8
0

−0.5
0 100 200 300 400 500
time (samples)
Adimensioned reed position

0.4
C =0
P
0.3 CP=0.8
0.2

0.1

0
0 100 200 300 400 500
time (samples)

Figure 8.4.: Pressure at the reed input and reed position plotted along the time, for a
sustained note, for different values of CP

The time-evolutions of pressure and reed opening look similar for both cases of the
pressure recovery coefficient. A plot of their spectrum (fig. 8.5) shows a reduction of
high frequencies, and in fact to hearing, a high value of CP 3 the sound is mellower and
closer to that of an oboe. It is also difficult to base a perceptive analysis on the pressure
wave inside the instrument, because the radiation filters the low frequency components.

60
C =0
P
C =0.8
40 P
Power spectrum density (dB)

20

−20

−40

−60

−80
0 1000 2000 3000 4000 5000 6000 7000 8000 9000 10000
Frequency (Hz)

Figure 8.5.: Spectrum of the pressure inside the bore for different values of CP
3
We remind that CP is limited to a range [0, 1], and in practice CP > 0.8 are never observed in conical
diffusers
8.5 Results of the simulation 147

A full analysis of the instrument model must be based on more than a few plots. The
behavior of the instrument corresponds to its response for the whole set of physical
parameters, and to likely transitions between these sets. From a musical point of view,
it is the experience of using these models that can determine their suitability. From a
physical point of view, ideally all possible ranges of parameters and possible transitions
would have to be tested.

8.5.1. General remarks on the time-evolution curves


The time variation of the simulated reed position can be compared for instance with the
observations of figure 6.2. The asymmetry noticed in the observations in the fractions
of periods the reed remains closed or open is also clearly visible in the simulations. As
discussed in section 2.5 this is a consequence of the acoustical response of the cone.
As discussed by [Gokhshtein, 1979], after the reed opens, the internal pressure de-
cays exponentially in time during the first part of the open-state, due to the tail seen in
the reflection function. This decay is stopped by the arrival of the incoming pressure
wave p− which reopens the reed before the transition to the closed-state.
The variation of the duration of the closed-state with the mouth pressure pm is not
observed in the simulations. Figure 8.6 shows the pressure and reed-opening time-
evolutions for several values of pm from the threshold of oscillation until the first register
transition. The duration of the closed-state is seen to increase slightly between the
threshold and slightly higher pressures, but until the register transition it does not vary
significantly.
In the first results shown in figure 8.4 it is not possible to observe the mechanical
oscillations of the reed because these are highly damped. Figure 8.7 shows a plot of the
internal pressure and reed opening for different values of the parameter α corresponding
to the damping of the reed oscillator. For low values, several high frequency oscillations
can bee seen in the reed opening plot. These are hardly observable in the pressure
plot but they have significant perceptive consequences when hearing the sound, which
becomes brighter.
148 Chap. 8: Synthesis

Adim. pressure Adim. reed position


1
0.5
0

−1 0
0 200 400 600 0 200 400 600
1
0.5
0

−1 0
0 200 400 600 0 200 400 600
1
0.5
0

−1 0
0 200 400 600 0 200 400 600
1
0.5
0

−1 0
0 200 400 600 0 200 400 600
1
0.5
0

−1 0
0 200 400 600 0 200 400 600
time (samples) time (samples)

Figure 8.6.: Simulation results for variable mouth pressures (pm ). pm increases from
top to bottom, from a pressure just above the threshold of oscillation (here
2.6 kPa) until the transition to the second register (at 4 kPa).

Adim. pressure Adim. reed position


1
0.5
0

−1 0
0 200 400 600 0 200 400 600
1
0.5
0

−1 0
0 200 400 600 0 200 400 600
1
0.5
0

−1 0
0 200 400 600 0 200 400 600
time (samples) time (samples)

Figure 8.7.: Simulation results for variable dampings (α)


9. Conclusion

Reed instruments, and in particular double-reeds are complex systems that involve the
interaction between fluid mechanics phenomena and elastic solids. Although several
aspects of their behavior are quite well mimicked by rather simple models, the details
of the oscillating flow inside the instrument (mainly in the upstream part) and the
interaction between the moving reed and the flow are challenging problems. This is
especially true in oboes and bassoons, where the geometry of the duct introduces further
complications with respect to a clarinet mouthpiece, for instance.
The aim of this work was to propose a model of the double-reed functioning that could
be used in simulations of a complete instrument. An important goal was to propose
a more refined model (than the one available in the literature) intended specifically for
double reeds, or to explain why the generic model is appropriate for double-reeds as
well.
Since the start of the project we noticed the relative lack of experimental knowledge
on double-reeds, especially when compared to single-reeds. For this reason it was
important to collect some experimental data that could help determining some particular
aspects of the reed and the flow.

9.1. Quasi-static model of the reed

9.1.1. The double reed as a nonlinear exciter

One of the fundamental questions was to know whether the non-linearity that char-
acterises the exciter was changed with respect to other reeds. While other reed in-
struments usually have a regime where the reed oscillates between non-closed states,
double-reed instruments tend to play exclusively in regimes where the reed abruptly
shuts during part of its cycle. A model had been proposed by Wijnands and coll. relating
this tendency to an additional singular loss in the flow inside the reed [Wijnands and Hirschberg, 1995].
In terms of the non-linear characteristic curve of the reed, the maximum-low pressure
(PT ) should be displaced towards higher pressures so that multiple flow values could be
observed for a same pressure difference across the reed.
Our measurements showed that this model does not apply to oboe or bassoon reeds
(chap. 4). There is no qualitative change in the non-linear characteristic of the reed
(there are no multiple flow values for a same pressure, other than hysteretic effects due
to visco-elasticity of the cane reed).
In quasi-static measurements for a complete oboe reed with its staple, the displace-
ment to PT is opposite of what was predicted by Wijnands’ model. In terms of the reed
150 Chap. 9: Conclusion

model, this means that no singular losses are observed in the flow inside the reed. More-
over we have shown that this displacement can be justified by the pressure recovered
along the staple due to the expansion of the flow (diffuser effect, see section 5.10).
In oscillating regimes, it is not clear if the diffuser effect should be taken into account
in the exciter, because the staple can be seen as a continuation of the resonator. Due
to the complexity of the measurements in dynamic regimes, this ambiguity could not be
completely solved.
Because of this difficulty, a complete model of the instrument was built where the
diffuser effect can be controlled by means of the CP (pressure recovery coefficient) pa-
rameter. This parameter does not seem very relevant when observing the pressure and
reed opening time evolutions, nor their spectrum. In listening tests, the effect is audi-
ble, increasing the brightness of the sound and making the model more responsive in
transitions (chap. 8).

9.1.2. Reed opening geometry

With respect to the reed opening (which controls the volume flow into the reed for a given
pressure difference accross the reed), different thoughts could be found in the literature
concerning the variation of its shape throughout the opening range. In particular, it
was proposed for tenora reeds [Barjau and Agulló, 1989] that the reed closes from the
corners as the two blades approach, so that both the width and the height of the opening
is reduced as the blades approach. In such case, the reed area is proportional to the
square of the distance between blades. This view is shared by some musicians, who
seek this kind of behavior when making their reeds.
Our observations show that in oboe and bassoon the reed blades touch at the same
time throughout the length of the reed (sect. 3.3). Quantitative analysis of the reed
opening using an algorithm developed for this purpose (appendix H) confirm that for
these reeds the opening area is proportional to the distance between reeds. It is possible
that the behavior is different for some reed designs, which should be checked with reeds
designed specifically for this kind of behavior.

9.1.3. Elastic model

The simplest approach for the mechanical behavior of the reed is to consider it as a
spring, driven by the pressure difference across it. In a linear spring, the pressure
difference induces a proportional deformation of the string. In the reed, the relation
between pressure and deformation depends on the conditions of the reed: when it is
dry, there is some non-linearity and the reed is stiffer when nearly closed. A more linear
relation is achieved when the reed is humidified, a condition met when the reed is blown
by the musician (sect. 3.4).
Measurements of the deformation against pressure revealed a viscoelastic behavior
of the reed material (confirmed through relaxation time measurements in section 3.5).
Some of the long-term deformation is kept after the force is released. This should cause
9.2 Open questions 151

an offset in the reed equilibrium position through time, but no significant changes in
the oscillating regime of the reed.
Other measurements and observations on the reed properties were made throughout
a number of experiments that allowed to gather important knowledge and data about
the properties of the reed. They were listed in chapter 7 and are not recalled here.

9.1.4. Sound synthesis


A synthesis model for double-reed instruments was derived from a clarinet model in
chapter 8. Conical waveguides were implemented, keeping the traveling-wave formal-
ism, and the parameters, many of them measured throughout the experimental part of
this work, were adapted to double reeds. The exciter model was changed to include the
quasi-static effect of the conical diffuser.
The model for the conical resonator together with the change in the reed parameters
already approach the timbre of the model to that of an oboe, for instance. In a subjective
analysis, the pressure recovery does not greatly affect the timbre, but the behavior in
the transitories seems to be improved.
The application of the simulation to sound-synthesis requires an extensive work over
the control and parameter mapping of the model. For a similar model which was im-
plemented for other wind instruments in the framework of the project WINDSET1 (a
partnership between IRCAM and Arturia) this work was effectuated in order to be used
for musical synthesis.

9.2. Open questions


Although the quasi-static behavior of the double-reed is at this point well characterised,
there are many questions that remain unanswered which may affect the behavior of the
double-reed when used in the instrument.
One of the main issues is the step towards the dynamic model. The model used
in chapter 8 for the reed is mostly a static model. Although dynamic effects of the
reed mechanics are taken into account through the model of the harmonic oscillator,
unsteady effects of the flow should be considered also, since they would contribute to
and potentially modify the reed dynamics. An analysis of magnitudes shown in section
6.2 suggests that both the dynamics of the flow and the flow induced by the reed are
important and should have consequences on the behavior of the instrument (although
no consequences of the induced flow were noticed at the reed output, as mentioned in
section 6.4.3). These two aspects should thus be studied in more detail.
Another unsolved issue is the separation between the exciter and the resonator.
There is clearly a region where the acoustic wave equations are valid. Inside the reed
however, the variable boundary conditions and the velocity magnitudes call for the
use of the complete equations of fluid mechanics. In our model, for the exciter we
used an approximation of these equations which is valid in incompressible, stationary
1
see http://www.arturia.com/en/brass/samples.php for sound examples
152 Chap. 9: Conclusion

conditions. Even in the case where the exciter model is extended to include dynamic
effects of the flow, it will still be necessary to define a boundary for the transition to the
acoustics region.

9.3. Perspectives

The broad approach to the double-reed presented in this document can be seen as an
advantage, because it allows to relate several distinct features of the reed properties,
and to the empirical knowledge involved in its fabrication and utilisation. But it can
also be seen as a weakness: in fact in many of the measurements, the variability of the
results do not allow us to be more categorical.
Naturally, one of the open paths for future research is to investigate the reasons of
the variability observed for instance in the characteristic measurements. This can be
achieved with a finer control of the experimental conditions (in particular the use of
artificial lips and teeth) but also by a better measurement of the parameters of the reed
during the experiment.
Another kind of variability is related to the astonishingly wide range of timbres that
can be associated with double reeds. With similar working principles and geometries
(although different dimensions), the timbre of a double-reed can be mellow, like in the
oboe, flute-like like in the guanzi, strident, like in the bombarde... Is this variability
related to the reed geometry? Its mechanical properties induced by different fabrication
traditions? Is it due to the way they are played (between the lips against free from lips)?
These questions have to be studied through a collaboration between the musicians,
who know what to change in a reed to achieve a particular result in the sound, and the
scientific community which can provide the means of testing the physical consequences
of these changes and provide models to explain them. A collaborative project is being
planned with musicians to list the techniques used in reed making, evaluating the
agreement on a particular technique between several musicians, and trying to propose
experiments and models for the effects of these techniques.
In order to better compare the simulations to experiments it is necessary to be able to
control and measure the evolution of the parameters applied in experiments. This is the
aim of the artificial mouth (appendix G), where in principle it is possible to control each
action of the mouth independently in a stable way (i. e.,without being limited by muscle
fatigue, breath supply, etc.), and measurement access is simpler than in the mouth of a
musician. However, the relation between the parameters in the mouth and those in the
simulation can be complicated, and despite the accuracy of the control for stationary
conditions it is still not easy to impose precise time evolutions of the controls. These
issues will be addressed by the ANR project CONSONNES2
One other aspect that undeniably calls for more attention is the dynamical descrip-
tion of the reed. Measurements of the reed opening and flow profiles during oscillations
unveil a great number of interesting and puzzling phenomena, for which only the first

2
CONSONNES : CONtrôle de SONs instrumentaux Naturels Et. Synthétiques
9.3 Perspectives 153

questions were formulated. . . An interesting approach would be to continue the exper-


imental measurements of the dynamic flow, in particular for lower frequencies and in
bounded (because coupled to the bore) conditions, as well as a computational simulation
of the fluid-structure coupling in the reed.
A simple unidimensional model of the fluid structure coupling, and extended to in-
clude unstationnary flows can be tested through similar simulations as those shown in
chapter 8.
154 Chap. 9: Conclusion
Appendices
A. Summary of double reed instruments
Oboes and bassoons are only two examples of double-reed instruments. Several others
exist, and the differences in the way they are played, their resonators, or the range of
notes they are aimed at. Reed dimensions are also widely variable.
These differences account to the large varieties of sounds produced by double-reed
instruments. For instance, the bore shape influences the balance between odd and
even harmonics in the spectrum. By reducing and damping the motion of the reed, the
contact between the lips and the reed blades has a softening effect on the timbre.
The difficult question to answer is thus if double-reeds have a characteristic and
recognizable sound.
One of the interests of the physical model is that if it is sufficiently detailed and
precise, it should be able to describe and reproduce these variations in timbre, by
inputing it with the correct parameters. Although this was not fully investigated in this
work, it is clear that some of these effects arise by adjusting such parameters as reed
damping and mass (for the lip contact) or bore tapering angle (for the bore shape).
In the following sections we list some of the double instruments that can be found both
in modern, ancient and traditional music, which are eventually a source of information
about the behavior of the double-reed.

A.1. Classification of double reed instruments


One important characteristics of double-reed instruments that influence their timbre
are the bore shape which can be cylindrical, conical or in some cases smoothly pro-
gressing from cylindrical to conical (renaissance shawms) or still stepped conical. The
bell shape also slightly influences the timbre and mostly the sound volume produced
by the instrument.
Another important characteristic is the interaction between the player’s lips and the
reed, either in contact (in some cases a pirouette is added so that the lips have a fixed
position on the reed) or free, either inside a bag (in bagpipes) inside a protective wind-cap
or completely inside the mouth.
Table A.1 lists some examples of modern, ancient and traditional double-reed instru-
ments, and the main characteristics referred above.
The note range for which some of these instruments are designed to is also listed,
because it can be an interesting information when comparing two different instruments.
Some of these however constitute a family (for example shawms, crumhorns) so that
they can be found in several sizes and musical ranges.
Traditional double-reed instruments are spread over almost all countries in Europe
(Tenora, Bombarde and many more), most of Asia (Guan, Zurna and many more) and
158 Chap. A: Organology

Name Bore Blow Bell Range Obs

Oboe Conical Lips Slightly flaring B3b − G6

Bassoon Conical Lips Slightly flaring B1b − E5

Contrabassoon Conical Lips Flaring B0b − C4

English Horn Conical Lips Bulb E 3 − C6

Heckelphone Conical Lips Bulb A 2 − G5


# #
Oboe d’amore Conical Lips Bulb G 3 − C5

Crumhorn Straight Free (windcap) No bell (var)

Shawm Conical Pirouette Flaring (var)

Dulcian Conical Lips C 1 − G4

Rauschpfife Conical Free (windcap) No bell

Racket Straight Lips No bell

Bagpipe (chanter) Conical Free (bag)

Bombarde Conical Lips Flaring

Tenora Conical Pirouette Flaring

Guan Straight Lips No bell

Zurna Stepped cone Free (mouth) Flaring

Table A.1.: Examples of double-reed instruments and their main characteristics


A.2 Variability of reed shapes 159

Northern Africa. Bagpipes exist mostly in Europe, and their chanters can have both
conical or cylindrical shapes. Usually, to a conical chanter is associated a double-reed
and to a cylindrical one a single-reed (which is different from a clarinet or saxophone
reed because it is not striking).

More examples of double-reed instruments can be found in literature such as [Midgley, 1976]
or in the Wikipedia1 article for double-reeds.

A.2. Variability of reed shapes

Although a double-reed instrument is, as the name indicates, constituted by two blades,
the actual shape of the reed can vary substantially from one reed to another. Looking
into traditional instruments, it is possible to find reeds whose structure is considerably
different (for example the duduk, which has a duck-beak like shape). Double-reeds can
also be made out of different plants, which may give them different elastic properties
for instance (for example, the reed used in the shehnai in northern India, made out of
pala-grass and in the Pi nai or in sralai, a palm leaf is used instead of the reed) .

In most of the cases however, the basic profile of the reed remains more or less similar:
the entrance has an oblong cross-section, which progressively evolves into a circular
cross-section towards the reed output. In some cases, the downstream reed consists of
a metal conical tube (the staple), in others this tube is not part of the reed but part of
the instrument, where the reed is inserted.

For the most common double-reeds in the occident, even if the shape remains ba-
sically constant, the dimensions are significantly variable. In general, the lower the
range of the instrument, the larger is the reed, but as can be seen in table A.2, each
double-reed is not a scaled version of another reed.

Table A.2 lists some typical dimensions (see fig. A.1) in some of the most common
double-reeds which can be found in scientific articles (bassoon, oboe, english horn,
contrabassoon, chanter and tenora). A description of the dimensions used in this table
follows:

1
http://en.wikipedia.org
160 Chap. A: Organology

WI Reed opening width, measured inside the reed opening

HI Reed opening height at rest, measured inside the reed opening

WO Reed tip width, measured externally

HO Reed tip height at rest, measured externally

LR Reed length — length of a blade from the staple (or the narrowest part
of the reed duct, measured from the inside) until the tip

LSC Scraped length — length of the part of the reed tip where the bark is
removed

LST Staple length (or the length from the narrowest part of the reed duct,
measured from the inside until the butt)

NDS Narrowest diameter of the staple or reed duct (this can correspond to
two dimensions because the staple entrance is often oval)

XDS Diameter at the staple exit, or the butt

LT Total reed length

Bn Hb Ca Cbn Chnt Tn

WI 13.3 7.0 7.6 16.4 16

HI 0.9 – 1.2 0.8 – 1.0 1.2 1.7 1.6

WO 14.4 7.2 7.8 17.6 10

HO 1.6 – 2.1 1.0 – 1.2 1.6 2.4

LR 33 26 49 20

LSC 26 – 27.8 10 10 35 18

LST 23 47 24 20

NDS 3.7 2 × 2.5 2.0 × 3.5 4.5 1.6

XDS 5.3 4.8 5.0 5.7 3

LT 56 73 73 36

Table A.2.: Dimensions of the most common double-reeds in the occident (see figure A.1
for the meaning of the dimension abbreviations)

As a final remark, table A.2 should not be taken as a geometrical characterisation of


these double reeds. Variations between reed geometries go beyond those hereby listed.
A side view can vary from reed to reed, and perhaps more importantly, the profiles of
A.2 Variability of reed shapes 161

WO WO

WI WI

HI HO HI

LSC

LSC

LR
LR

LT

LST

LST

NDS

XDS

Figure A.1.: Reed dimensions


162 Chap. A: Organology

the thicknesses of the blades can be greatly variable.


B. Mechanical modes of vibration of the
double-reed

B.1. Mechanical response of the double-reed


In the section 2.4, an elementary model was proposed for the behavior of the reed when
used as an exciter of an instrument. This model is intrinsically quasi-static, that is, it
is valid in regimes where its variables change slowly with time.
In oscillatory regimes, other effects influence the motion of the reed and the evolu-
tion of the fluid flow. Among these we identified the inertia and damping of the reed
(sect. 2.5), which can be taken into account by modelling the reed tip as a harmonic
oscillator.
However, this model is still a simplification of the mechanical behavior of the reed
(even without considering its interaction with the flow), because the reed cannot be
reduced to its tip. The blades vibrate along their surface until the region where they are
attached to the staple, and for this reason several modes of vibration are expected to be
observed in the reed.
The analysis of these modes of vibration is important, on one hand because each
mode of vibration can have a different frequency of resonance, making a single harmonic
oscillator insufficient to describe the vibration of the reed tip, and on the other hand
because the motion of the reed inner surface is in contact with the fluid, inducing some
flow.

B.2. Modes of vibration of the reed


Equation (2.31) correctly describes a linear oscillator with one degree of freedom. In
general, a system with more than one degree of freedom shows several resonance peaks
in a plot of its mechanical response. As long as the linear hypothesis remains valid,
these systems can be described as a set of linear oscillators with different resonance
frequencies (ωr in equation (2.31)) and damping coefficients (α in the same equation).
Modal analysis studies the decomposition of an oscillating solid into this set of har-
monic oscillators, which are usually referred to as modes . This decomposition can also
be used to study the mechanical behavior of the reed as a 3-dimensional system with
boundary conditions, such as the fixed ends on the butt side, the free ends on the tip
and the contact conditions between the two blades on the reed’s rails. Modal analysis
was previously applied by Pinard ([Pinard et al., 2003]) to clarinet single-reeds.
In practice, modal analysis consists in decomposing the reed motion into its eigen-
modes of vibration, each of which has its own resonance frequency ωr and damping α.
164 Chap. B: Preliminary modal analysis

The whole system is a linear combination of these modes of vibration, and they can be
seen as independent harmonic oscillators. The excitation of these modes depends not
only on the frequency of the driving force, but also on the distribution of the force over
the reed. In reed instruments the excitation is particularly different from the excitation
in percussion instruments because the force is usually distributed along the surface of
the reed whereas in percussions the excitation is usually very localized.

B.3. An oscillator driven by an external force


When driven by an external sinusoidal force of frequency ω/(2π ), the harmonic oscillator
seen in section 2.4.3 will respond with an oscillation with the same angular frequency ω,
and with an amplitude that depends on the relationship between the driving frequency
ω and the resonance frequency ωr and quality factor Q = ωr /(2α ) according to the
formula of a Lorentz peak centered at the resonance frequency:
X (ω) 1/m
= (B.1)
F (ω) ω2
r − ω2 − 2ıαω
If the velocity (V ) of the oscillator is measured instead of the position (X ), a similar
formula can be obtained by differentiation of eq. (B.1), V (ω) = ıωX (ω):
V (ω) ıω/m
= (B.2)
F (ω) ω2
r − ω2 − 2ıαω
As a linear combination of simple harmonic oscillators, the response of a more com-
plex system with several modes of vibration as described in section B.2 is the sum of
the responses of each individual mode. Linear superposition of the modes is assumed
because the flows induced by the acoustic flow and reed motion are small enough to
induce a non-linear coupling between the modes.
An analysis of the spectrum will thus give an indication of the parameters ω r and α
of each mode of vibration of the reed. In general, because several modes of vibration
exist in the structure, it may be difficult to measure these parameters in each individual
peak.

B.4. Frequency analysis of reed vibration


B.4.1. Experimental approaches
In vibromechanics, a usual method for imposing a vibration of the object to be studied
is to rigidly connect a shaker to the object, so that it vibrates synchronously to the
exciting device.
In the case of the double reed, in normal playing conditions, it is driven by a force that
is distributed over the reed blades, rather than applied punctually. For this reason, we
preferred to use air vibrations as the excitation to the reed, provided by a loudspeaker
placed in the vicinity of the reed.
Two different methods were tried out to excite the reed vibration:
B.4 Frequency analysis of reed vibration 165

External excitation The reed is supported by the butt or the staple and an acoustics
source (a loudspeaker) is placed next to the reed, but not too close, so that the
wavefronts are approximately plane in the vicinity of the reed.

Internal excitation A small loudspeaker is inserted in a small chamber with an opening


where the base of the reed can be inserted, the same way it is inserted in a
resonator.

The method of external excitations has the disadvantage of requiring more power,
because the wave field spreads throughout the room rather than concentrating over
the reed. A compromise is obtained by approaching the reed to the loudspeaker, but
not too much, otherwise the pressure distribution is no longer homogeneous at high
frequencies.
Internal excitations require a weaker sound source but it is much more difficult to
maintain a constant amplitude excitation throughout the spectrum because of the influ-
ence of the acoustic modes of the loudspeaker chamber coupled to the reed. Moreover,
since the propagation of the waves is longitudinal with respect to the reed, the frequency
at which the pressure excitation is no longer homogeneous is lower than with external
excitations (provided that the reed is placed parallel to the loudspeaker in the external
case).
The excitation signal was a sweep (a sinusoidal wave of varying frequency) running
from 0 to 48000 Hz, 10 seconds long.
In both cases, the actual amplitude of the pressure oscillations was measured close
to the reed before the actual measurement of the amplitude of vibration of the reed. For
external excitations it would be possible to measure pressure simultaneously with the
vibration measurements, but for internal excitation it was not possible to maintain the
microphone inside the reed without disturbing the reed oscillations.
We then used the measured pressure variations to correct the amplitude of the exci-
tation signal so that it would be approximately flat at mid-range frequencies (1000-4000
Hz). A second measurement of the pressure variation with the new excitation signal was
used during the analysis to compute the reed response as a ratio of the measured reed
vibration to the pressure forcing it.
The vibration of the reed is measured using a laser vibrometer (Ometron VH300+).
This device measures the velocity of the structure at the point of incidence of the laser.

B.4.2. Data analysis

Analysis of the results consists simply in calculating the FFT of the reed vibration signal
and dividing it by the FFT of the reference (excitation) signal sent to the loudspeaker.

B.4.3. Comparison of excitation methods

Figure B.1 shows the response of one reed measured using the two different methods
described in section B.4.1.
166 Chap. B: Preliminary modal analysis

250
Internal
External
200

Amplitude (dB) 150

100

50
0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5
Frequency (Hz) 4
x 10

Figure B.1.: Response of reed 744 using two different excitation methods (external and
internal) as described in section B.4.1.

The results are quite different from one method to the other. The fact that the two
measurements were taken at two different dates doesn’t seem sufficient to explain these
differences. We suppose that the resonances of the loudspeaker cavity and reed can be
responsible for the differences observed, because it is difficult to guarantee the reeds
are placed in the same position for every measurement so that the cavity geometry is
the same.
In fact, measures of the actual pressure intensity inside the reed should be taken
simultaneously to the measurements of reed vibrations. This has not been accomplished
because this technique would require small microphones and perforation of the reed.
Due to the problems identified in the internal excitation method, we chose to work
with external excitation in the following sections, unless stated.

B.5. Admittance spectra of bassoon reeds


In the following paragraphs we present some admittance curves of double-reeds. Most
of the reeds used for these measurements are bassoon reeds made from natural cane.
In one particular case we present a measurement for a synthetic bassoon reed, and
a final measurement concerns a same oboe reed measured with two different excitation
methods.
These results require further investigation in order to investigate the origin of some
discrepancies observed for different methods of excitation and orientations of the reed.

B.5.1. Comparison of two different cane reeds

Figure B.2 plots the frequency response of two different cane reeds. Reed labeled 744
is tailored to be used in concert whereas reed 766 is not finalized, so that its blades are
thicker at the tip.
Both reeds have strong resonances between 2400 and 2700 Hertz, which we will call
B.5 Admittance spectra of bassoon reeds 167

50
744
766
0
Amplitude (dB)

−50

−100

−150
0 1000 2000 3000 4000 5000 6000 7000 8000 9000 10000
Frequency (Hz)

Figure B.2.: Comparison of two different cane reeds.

the mechanical frequency of resonance of the reed. These frequencies are similar to
those of clarinet reeds, which are situated in the 2100-2400 Hz region (for example,
[Pinard et al., 2003]). These values can also be compared to the ones found for chanter
reeds by S. Carral [Carral and Campbell, 2005]: for a reed with a similar geometry but
smaller dimensions (10 mm wide tip, 18 mm long for the chanter against 14 mm wide
and 28 mm long for the bassoon), the author found a resonance frequency of about
4800 Hertz for dry reeds. The larger dimensions of bassoon reeds probably account in
a great part for the lower resonance frequencies when compared to the chanter reeds.
The two measured bassoon reeds show similar response spectra below 5000 Hz,
although the reed that was not fully scraped sees its peaks displaced towards higher
frequencies, and a higher frequency of resonance. This is probably due to the increased
stiffness of the unscraped reed. However, reed makers usually scrape the tip of the
reeds to increase their frequency [Kopp, 2003].

B.5.2. Effect of humidity

For natural cane reeds, instrumentists notice a great dependence of the reed quality
with the damping of the reed. These usually have to be soaked for a few minutes before
being played, and blown a few times with moist hot air before they sound in perfect
conditions.
Figure B.3 plots the dependence of a reed spectrum on the time the reed is soaked
before the measurement. The first resonance frequency is substantially lowered after 5
minutes of soaking ( from 1900 Hz to 1640 Hz) which is probably due to the mass of
water captured by the reed. The quality of the resonances is slightly increased (peaks
are narrower). Simultaneously, results of section 3.4 and appendix F suggest that the
reed stiffness k increases when the reed is soaked (an increase in stiffness alone implies
a higher resonance frequency).
168 Chap. B: Preliminary modal analysis

100
Dry
5 min soaking
30 min soaking

Amplitude (dB)
50

−50
0 500 1000 1500 2000 2500 3000 3500 4000 4500 5000
Frequency (Hz)

Figure B.3.: Comparison between different levels of humidification.

B.5.3. Comparison between plastic and cane reeds

In some of our experiments it is more practical to use synthetic reeds (made out of plastic
materials) whose properties are more stable because they depend less on environment
conditions and on the effort applied during the experiment.

50
Plastic
744
0 766
Amplitude (dB)

−50

−100

−150
0 1000 2000 3000 4000 5000 6000 7000 8000 9000 10000
Frequency (Hz)

Figure B.4.: Comparison between plastic and cane reeds.

However, as can be seen in figure B.4, the dynamical properties of the plastic reed
are quite different from the cane reeds: its resonant frequency is lower (around 1900
Hz) and less damped, since the peak is narrower.
Dynamical measurements show how plastic reeds are less damped when compared to
cane reeds, although the lowest resonance peak stands close to the resonance frequency
of a real cane reed.

B.5.4. Oboe reeds

The same method can be used for oboe reeds, although due to their dimensions, their
vibration is fainter than in the case of bassoon reeds. The setup remains basically
B.6 Discussion 169

the same, but two different reed positions were tried out: with the rails facing the
loudspeaker and with the reed opening facing the loudspeaker.
Oboe reed mechanical response
50
debout
allonge
0
Amplitude (dB)

−50

−100

−150
0 1000 2000 3000 4000 5000 6000 7000 8000 9000 10000
Frequency (Hz)

Figure B.5.: Mechanical response of an oboe reed.

The two measurements present different peak configuration below 5000 Hz, but there
is one peak common to both experimental setups at 3492 Hz. In the first setup, the
resonance peak is partially hidden under a higher peak around 4200 Hz, which however
does not exist in the other setup. This is probably due to a vibration mode of the
structure supporting the reed.
This resonant frequency is substantially larger than that of the bassoon, as could be
expected based on the smaller dimensions of the oboe reed. However, when compared
to the chanter reed, the oboe reed has a smaller tip but a bigger vibrating length which
can explain the lower resonant frequency of the oboe reed.

B.6. Discussion
The methodology presented in this chapter is similar to other experimental setups
used to measure spectral characteristics of clarinet reed vibrations using external
excitations [Pinard et al., 2003] and chanter double reeds using internal excitations
[Carral and Campbell, 2005].
With our setups we tried to compare results using both methods for similar reeds,
however practical difficulties prevented a full recording of the exciting signal along with
the response signal in the case of internal excitations, which may be an explication of
the remarkable differences between the two methods. These differences were observed
in several different reeds (although figure B.1 only depicts this situation for one of the
reeds).
The fact that the two kinds of setups were performed in different days, and that for
instance it is possible that the laser vibrometer was pointed at a slightly different point
on the reed, might also be part of the explanation for the differences in the results,
although it cannot explain differences in the measured resonance frequency. Further
measurements, performed in more controlled conditions will be required to analyse the
origin of these differences.
170 Chap. B: Preliminary modal analysis

Nevertheless, and considering that external excitation is simpler than its counterpart,
this method proves useful in extracting important information about the dynamical
properties of the solid parts of the reed. For instance, the frequency of resonance of the
reed can be used to calculate the effective mass of the reed.
It is important to notice that the resonance frequency measured in this section, are the
mechanical resonance frequencies, which are usually very different from the frequency
of oscillation of the reed when blown alone (tuned squeak, see sect. 6.4.3)
Other than the resonant frequency, the width of each resonance peak is gives us the
quality factor (Q = ∆frf , where ∆f is the width of the peak at -3 dB).
Given the resonant frequency, the quality factor and the further knowledge of the reed
k
stiffness (from section 3.4), it is possible to determine the effective mass (m = (2πf )2
)
and damping (r = m ωQr ) of the modes of vibration of the reed that are relevant to its
description.

B.6.1. Importance of the dynamic aspects of the reed

Given that the forcing of the reed is usually determined by the pressure variations inside
the instrument (which have direct consequences in pr (t )), it is the left-hand side that
will determine the importance of the time-derivatives in the right-hand side of equation
(2.31). This can be better understood by doing a frequency decomposition of equation
(2.31):

∆P (ω)
= ωr2 ∆X (ω) + αıω∆X (ω) − ω2 X (ω) (B.3)
ms

In equation (B.3) the capitalized variables represent the Fourier transforms of ∆p(t )
and ∆x (t ). The Fourier transform of a derivatives of x (t ) correspond to multiplying the
frequency-domain function X (ω) by ıω.
Written in this form, equation (B.3) allows to see that the forcing ∆p(t ) is distributed
by the terms in the right-hand side. This way, it is the balance between ω r , α and ω
that will determine the relevance of each of the stiffness, damping and inertia terms.
For low frequencies (ω  ωr ) the inertia term can be neglected, and if ω  ωαr , damping
can also be neglected. We thus see that for quasi-static regimes, defined with the former
conditions on ω, equation (2.22) provides a good mechanical description of the reed.

B.6.2. Importance of higher-order modes of vibration of the reed

If the pressure under the reed can in fact be considered homogeneous, the fundamental
mode of vibration of the reed can be excited preferentially over the higher modes. Higher
harmonics of the bore pressure can excite higher modes of the reed, but depending on
the played note, the bore harmonic might or might not fall over a reed resonance.
Pinard [Pinard et al., 2003] gives indications on the first two flexural (longitudinal)
and torsional (transverse) modes of vibration of the clarinet reed. Due to the contact
between the sides of the double reeds, the flexural modes are not expected to be observed
B.6 Discussion 171

in oboes and bassoons because flexural modes have maximum vibration regions (anti-
nodes) along the sides.
Thompson [Thompson, 1979] studied the influence of the reed resonance frequency
on the tone produced by clarinets. It is probable that higher resonances of the reed might
have similar reinforcement effects on some of the partials produced by the instrument.
Section 6.3.5 presents some observations of the reed movement, with particular at-
tention to high-frequency oscillations during the period that the reed remains open.
These oscillations can be associated to several causes, and among them, the mechan-
ical eigenmodes of the reed. To these oscillations on the reed position are associated
simultaneous oscillations of the pressure and volume flow.
With respect to higher flexural modes, these are not very important in terms the
fluid-structure coupling, because they have several anti-nodes that vibrate in opposite
directions. The total flow ‘‘pumped’’ by the reed (average flow) is less important for a
same vibration amplitude, although it is possible that local flows are induced under-
neath the reed.
172 Chap. B: Preliminary modal analysis
C. The momentum equation in fluids

C.1. Differential form


The differential form of the conservation of momentum equation is known as the Navier-
Stokes equation. It can be formulated as:

1
∂t ui + uj ∂j ui = − ∂i p + ν∂2 ui (C.1)
ρ
In this equation, ui stand for the components of the flow velocity field vectors, ρ is
µ
the volume mass of the air, and ν = ρ is the cinematic viscosity (µ being the dynamic
viscosity).

C.2. Integral form


Equation (C.1) can be integrated in a control volume Ω, bounded by the control surface
Σ:

1
Z Z Z Z
∂t ui dV + uj nj ui dS = − pni dS + σij nj dS (C.2)
Ω Σ Σ ρ Σ

C.3. Vena Contracta


We can now apply equation (C.2) to a Borda Tube immerse in a flow.
The control volume Ω will be bounded by the duct walls, its bounding surfaces will
be perpendicular to the flow, and will entirely include the borda tube.
The flow is considered to be inviscid, stationary and irrotational, so that the first term
in the left-hand side vanishes, as well as the last term in the right-hand side. Equation
(C.2) becomes:
Z  
1
uj nj ui + pni dS = 0 (C.3)
Σ ρ
Along the duct walls in the upstream region the pressure is homogeneous, and the
same is true for the downstream region, so that in this region the second term in
equation (C.3) cancels out.
Since the flow is parallel to the walls, uj nj is zero along the walls of the duct, cancelling
out the first term of the equation as well.
The two cross-section surfaces Σ1 and Σ2 remain in the equation.
174 Chap. C: Momentum

PSfrag replacements Σ

p1 p2
u2
S1 Sj St

S2

Sw

Figure C.1.: Flow into a Borda tube

In the upstream surface (Σ1 ), the flow is homogeneous along the cross-section of the
duct. The velocity is ux = u1 . The pressure is also homogeneous and equals p1 .
In the downstream surface (Σ2 ), the flow is zero everywhere except for the region
where the jet crosses Σ2 : over the surface Σj the flow velocity is u2 . As for the pressure,
it equals p1 everywhere outside the Borda tube and p2 inside the tube. Here it is
homogeneous, because there is the pressure does not change across a jet.
Collecting all these assumptions leads us to the formula:

S1 p1 + S1 ρu12 = (S1 − St )p1 + St p2 + Sj ρu22 (C.4)


Cancelling out equivalent surfaces on both sides of the equation:

S1 ρu12 = St (p2 − p1 ) + Sj ρu22 (C.5)


or
q2
ρ = St (p2 − p1 ) − Sj ρu22 (C.6)
S1
using the equation of conservation of mass:

u1 S1 = u 2 Sj = q (C.7)

∆p is related to the flow by the Bernoulli formula:


1
∆p = ρu22 (C.8)
2
Supposing that the duct is much wider than the Borda tube, S1  St , the term on the
left-hand side is much smaller than those on the right-hand side, and can consequently
be neglected. It is now possible to calculate the discharge coefficient C D

Sj ∆p 1
CD = = = (C.9)
St ρu22 2
C.4 Singular losses 175

PSfrag replacements
C.4. Singular losses
A similar method to that described in the previous section can be used to calculate the
pressure drop in a singular loss scenario.

Σ

p1 p2
Sj
u1 u2
S1 S2
Sw

Figure C.2.: Singular losses in a channel: some of the jet’s (in the downstream region)
kinetic energy is dissipated through turbulence in the mixing region (in
gray)

In this case, the control volume Ω encloses the mixing region, where the dissipation of
the kinetic region occurs. The cross-section surfaces S1 and S2 are situated respectively
in the upstream and downstream region from the mixing region, such that the flow can
be considered as laminar.
Considering the hypothesis of inviscid and stationary flow, as in section C.3, equation
(C.3) can be used in the control surface Σ. With similar considerations, it yields

Sj ρu12 + S1 p1 = S2 ρu22 + S2 p2 (C.10)

If the mixing region is not very long, such that the cross-section does not vary sub-
stantially between S1 and S2 , these areas can be considered similar.
The conservation of mass can be used to simplify equation (C.10), stated in the form:

Sj u1 = S 2 u2 = q (C.11)
This yields for the pressure drop across the mixing region:
 2  
q Sj Sj
p2 − p 1 = ρ 1− (C.12)
S2 S2 S2

C.5. Potential theory for the Vena contracta


In a flow that can be described by the Euler equations [Landau and Lifchitz, 1989], or
in other words, without viscosity, and if in addition it is incompressible and irrotational,
it can be described by a potential φ such that,

u = ∇φ (C.13)

For bidimensional flows, this relation can be expanded into:


∂φ ∂φ
ux = uy = (C.14)
∂x ∂y
176 Chap. C: Momentum

A’ A

B’ B

P0

C’ C
a∞

Figure C.3.: Flow in the complex spatial plane (z)

Another interesting definition is the stream function:

∂ψ ∂ψ
ux = uy = − (C.15)
∂y ∂x

For further analysis, we will use the following complex variables:

z = x + ıy u = ux − ıuy w = φ + ıψ (C.16)

Using these variables, we can see that equations (C.14) and (C.15) express the Cauchy-
Riemann conditions for the differentiability of function w.

C.5.1. Ecoulements par des orifices


This section describes the flow from a reservoir entering a straight tube through a duct
whose walls are at an angle with the tube walls. The Borda tube is a particular case
of this flow, where = pi.
The coordinate system is placed at the center of the entrance cross-section of the
tube. The upstream flow is bounded by the upstream duct, whereas downstream the
flow boundaries may not coincide with the tube walls because of the Vena Contracta
effect.
A hodograph is the representation on the complex plane of all the flow velocity val-
ues. In Landau [Landau and Lifchitz, 1989] it is supposed that the flow velocity tends
to 0 upstream. In addition, after the detachment from the duct walls, the velocity is
supposed to be constant and equal to u∞ . This can only be true for the external stream-
lines, establishing the limit between the jet and the stagnant fluid. This assumption is
C.5 Potential theory for the Vena contracta 177

|v| = v∞ C ≡ C0

B B’

A ≡ A0

Figure C.4.: Hodograph — Set of velocities (u) found in the flow

justified by the fact the stagnant fluid has a pressure equal to P0 all the way from the
tube entrance because otherwise there would be some fluid moving between two points
in the stagnant region. In the jet, however, this is not true because the flow velocity
varies along the x direction.
With these assumptions, the hodograph is a circular section (see fig. C.4).
This hodograph can be represented in a simpler way by taking its logarithm:

 
i ∗
ζ = − log − v (C.17)
u∞

which maps the circular section into the half-strip represented in figure C.5. This will
be useful for the problem solution.
For the potential (w) the analysis must be done independently for the two quantities.
The stream-function (ψ) is constant on each streamline (psi and φ iso-contours are
orthogonal). Walking upstream from the jet, ψ must vary at a constant rate (− ∂ψ
∂x
= u1 ,
yielding ψ = −u1 x + K). K is chosen so that the rightmost streamline has ψ = 0
(fig. C.6).
∂φ
φ can be integrated on each streamline. On the jet side we have ∂y = u1 , that is
φ = u1 y + K. K is now chosen such that φ = 0 at the point where the jet detaches from
the tube walls, and φ tends to −∞ . On the reservoir side, flow velocity varies roughly
as 1r , r being the distance to the tube entrance. φ thus varies as ln(r ) and tends to +∞
as r → +∞.
The w plan is thus crossed by an infinite horizontal strip with u1 a1 as width, where
a1 is the jet width at infinity.
In order to calculate the shape of the outermost streamline, we must plot a new
simplified graphic which transforms the flow in the imaginary half-plane. The new
178 Chap. C: Momentum

B’ A’

C ≡ C0


B A

Figure C.5.: Transformed hodograph (ζ )

C B A

C’ −v∞ a∞ B’ A’

Figure C.6.: Flow potential set (w


C.5 Potential theory for the Vena contracta 179

-1 1
A’ B’ C0 ≡ C B A

Figure C.7.: Auxiliary variable v

variable v can be related to ζ and w:


u∞ a∞
w = − log v (C.18)
 π 
2
ζ = i − 1 arcsin v (C.19)
π

With these relations, we can compute the streamline shape by noticing that on the
line (BC), ψ = 0 and thus the complex potential w is reduced to its real part.
 dφ 
= u = u∞ e −iθ (C.20)
dz BC

=(u )
where θ = − arg u = − arctan <(u ) corresponds to the flow velocity direction. In conse-
quence,
1
dz = dx + idy = e iθ dφ (C.21)
u∞
ζ is also simplified on this streamline:
π 
(ζ )BC = − log −ie iθ = i

−θ (C.22)
2
From (C.18) and (C.19) φ can be deduced as a function of θ:

     
u∞ a∞ πζ u∞ a∞ π  π
φ=− log sin i =− log sin θ− (C.23)
π 2 π π − 2 2
so that  
u∞ a∞ π  π
dφ = − cot θ− dθ (C.24)
2 π − 2 2
180 Chap. C: Momentum

Equation (C.22) is used again for the streamlines:


 
a∞ iθ π  π
dz = − e cot θ− dθ (C.25)
2 π − 2 2
which is integrated over θ between and π2 separately for the two parts of the complex
plan z, with boundary conditions x = a2 et y = 0.

C.5.2. Borda tube ( = −π)


For the particular case of the Borda tube, the two following differentials need to be
integrated:

 
a∞ θ π a∞
dx = − cos θ cot − dθ = − (1 − sin θ) (C.26)
2π 2 4 2π
 
a∞ θ π a∞
dy = − sin θ cot − dθ = − tan θ(1 − sin θ) (C.27)
2π 2 4 2π

The simplest method for this integration is to suppose that x and y are known for
θ = π. This corresponds to the point of detachment of the jet, avoiding that infinities
arise as integration constants.
The integrals in C.27 are:
 
a∞ θ cos θ
x = + + Kx (C.28)
2 π π
   
a∞ θ θ
y = 2 log cos + sin + sin θ + Ky (C.29)
2 2 2

The difference in x for −π and π is the distance between the tube walls and the jet at
infinity: 
a∞ 1 1 π 1 1  π  a

∆x = + cos + − + cos − = (C.30)
2 2 π 2 2 π 2 2
The tube diameter is filled by the jet diameter (a∞ ) surrounded by two stagnant
regions which add up to a∞ , so that we conclude that the jet occupies half the tube’s
diameter.
C.5 Potential theory for the Vena contracta 181

Figure C.8.: Jet boundary profile, according to equations (C.28) et (C.29). The scale unit
length is the jet length at infinity (a∞ ).

a∞ a∞
2 2

a∞

Figure C.9.: Flow shape for the Borda tube


182 Chap. C: Momentum
D. Magnitudes of viscous effects in the
double-reed
The geometry of the double reed channel is in certain aspects very different from the
geometry of the mouthpiece chamber as seen by the jet entering a clarinet or a sax-
ophone. The main difference resides in the fact that the narrow reed channel in the
clarinet mouthpiece is relatively short (about 2 mm long), while in the double reed the
channel remains narrow throughout the reed length (about 4 cm).
Moreover, the double reed terminates in a constriction, a region where the total cross-
section has a minimum, which is where the reed staple begins (the metallic piece the
reed is tied to). .
Both factors can be a source of additional viscous losses that do not exist in single reed
instruments. The relevance of these distributed losses can be estimated by considering
a model of a long rectangular channel, with one narrow dimension in its cross-section
and the other sufficiently wide.

D.1. How important are viscous losses?


Let us suppose that the flow is stationnary (quasi-stationnary model, as in the char-
acteristics measurements of chapter 4). Let us suppose also that the head loss is due
only to viscous losses inside the reed. Then the head variation for a given travelled
length (∆L) inside the reed where the smallest dimension in the cross-section (d) does
not vary considerably (from the N.-S. equation, keeping only the pressure gradient and
the viscosity terms, because the velocity cannot vary if the cross-section is constant due
to mass conservation [Kundu and Cohen, 2002]):

U
∆P ' µ ∆L (D.1)
d2
As a worst-case scenario, a lower limit for d is the reed opening width, which is
typically 1 mm . In this case, taking ∆L = 10 cm, µ = 1.73 × 10 5 (S.I.) and U = 100 m/s,
the pressure drop inside the reed is around 150 Pa, a value 100 to 1000 times lower
than typical mouth pressures used when playing the instrument.
As the reed closes, this value greatly increases. The inter-blade distance d tends to 0,
but over a small length. It can be supposed that even if the two blades only touch at the
tip, there is a considerable reduction of the inter-blade distance over a fraction of the
bending length of the reed. In particular, since the reed is very flexible at the feather
edges (of very reduced thickness), it is possible that the two reeds come into contact
over a length of approximately 5 mm .
184 Chap. D: Viscous effects

In this case the pressure drop is greatly increased at the tip. If we suppose that
the thin entrance channel is ∆L = 10−3 m in length, and only this part is considered
important in the pressure drop, the above value, at the standard opening of 1 mm is
100 times smaller, that is 1.5 Pa.
As the mouth pressure increases, the reed gradually closes, but the flow velocity also
increases. From the standard Backus model (see sect. 2.4), this velocity is a function of
d as:
r
2kd
u= (D.2)
ρ
This velocity can be used to calculate the function of the viscous pressure drop at the
reed feather edges against the inter-blade distance d:
r
2k
(∆p)vt = µ d −3/2 ∆L (D.3)
ρ
If the inter-blade distance is reduced 10 times, the pressure drop increases approxi-
mately 30 times, bringing it to 50 Pa. In order to the viscous pressure drop to become
comparable to mouth pressure values ((∆p)vt 1kPa), the inter-blade distance d has to
be reduced of a factor of 100, that is near the beating pressure (around 30 kPa).
Given these values, viscous effects inside the reed can be considered to be unimpor-
tant, ven if the inter-blade channel is much longer. An upper limit for the length of
this channel is the distance over which the reed blade is scrapped (about 5 mm in the
oboe), so that viscous effects would stil only be noticeable very close to the reed beating
pressure.

D.2. Turbulent flows


The above discussion is only valid in laminar flow conditions. In most of the pressure
range used in the reed however, the flow was found to be turbulent.
The exact value of the distributed losses in turbulent flows is difficult to calculate
in the internal geometry of the reed. A discussion similar to the above can be made
by using well-known results for cylindrical ducts, which are expressed in the ‘‘Moody
chart’’ [White, 2001]. It relates the friction factor f to the geometrical parameters of the
duct and the Reynolds number of the flow.
If no gravity effects are considered, the friction factor can be used to calculate the
pressure drop in the flow due to the turbulent dissipation:

L ρu 2
∆p = f (D.4)
h 2
Although this formula is given only for cylindrical pipes, we will use it to estimate
the pressure drop in the reed tip. Again, the distance between reeds determines the
magnitude of this formula, because the perpendicular dimension is about 10 times
larger. For its maximum opening, we consider this distance to be 1 mm over a length
D.3 Discussion 185

of 1 cm ( dL = 10). In the reed tip, the maximum velocity has a magnitude of about 100
m/s, implying Re ' 7 × 104
The friction factor is also dependent on the roughness (ϸ) of the tube walls. The fibers
are taken as determinant in the roughness of the reed walls. Their diameter is about
10−4 m [Casadonte, 1995]. The ratio between the roughness and the channel height is
thus hϸ = 0.05 if we suppose that only half a fiber can emerge above the other fibers.
A remark has to be made about the concept of roughness: in the Moody diagram, the
roughness is supposed to be isotropic over the duct walls surface. In the reed, this is
not the case, since the fibers are directed along the reed. This fact can influence the
results of this analysis, but using the fiber width should allow to have a reasonable
upper bound for the pressure drop due to turbulent diffusion.
In these conditions the friction factor can be found to be f ' 0.07, so that the pressure
drop due to turbulent diffusion is upper bounded by ∆p ∼ 4200Pa.
As the reed closes, this value should increase faster than the inverse proportion of
the distance between reeds h, because the ratio hϸ increases as well.

D.3. Discussion
Due to the viscous character of the flow and the roughness of the reed walls, the dis-
tributed losses may not b negligible. In fact the upper bound of 4.2kPa for these losses
is of the same magnitude of the pressure recovery that was found in section 5.10.3.
Due to the particular texture of the reed walls, which should be rough when observed
in a transverse direction in relation to the main flow, the effective roughness should be
considerably smaller than the value used above, which could reduce the friction factor
(and the pressure drop) by up to 100 times. When the reed is almost closed, however, it
is very likely that the turbulent diffusion will affect substantially the total pressure drop
in the reed.
The distributed turbulent losses can cancel some of the effects of pressure recovery
found in section 5.10 within the reed. This may be part of the justification as to why
the conical diffuser should be supposed to start at the staple input.
186 Chap. D: Viscous effects
E. Effect of gravity in the bending of a jet
This appendix presents a short estimation of the bending due to gravity that should
affect the jet in figures 5.1.
Based on the pressure applied in the artificial mouth (supposed to be the pressure
drop in the reed tip (∆p)1 , the velocity at the reed entrance is
s
2∆p
utip = = 5m/s (E.1)
ρ

Using the conservation of mass, the velocity at the reed output, is determined from
the velocity at the reed tip

S 2
uout = utip ' ' 0.5m/s (E.2)
Sout 18
In figure 5.1 the bending of the flow can be determined to be about 2 mm at a distance
L
of 10 mm from the reed output. A fluid particle takes uout = 19 ms to cover these 10
mm.
The same fluid particle of volume dV is affected by a weight force

dFg = dV (ρCO2 − ρair )g (E.3)

where g the acceleration of gravity, which induces an acceleration a in the fluid particle
given by
dFg = dVρCO2 a (E.4)

The total acceleration in the vertical direction of the fluid particle is then
 
ρair
a= 1− g (E.5)
ρCO2

The vertical displacement of the jet during the above calculate interval of time ∆t is

a (∆t )2 3.3 × (18 × 10−3 )2


∆y = = ' 5 × 10−4 m (E.6)
2 2
This displacement is four times smaller than what is observed in figure 5.1.
188 Chap. E: Effects of gravity
F. Alternative measurements of reed elasticity
This appendix is a complement to the data presented in section 3.4. The elasticity of
other reeds than the synthetic plastic reed was measured using a similar procedure.
The pressure control is slightly different, and this is the reason why the applied pressure
usually does not decrease underneath a few kPa.

2.5
Reed opening area (mm )
2

1.5

0.5

0
0 10 20 30 40
Pressure difference (kPa)

Figure F.1.: Measurement of the reed opening area against the applied pressure for a
Fibercane (Glottin) reed

Figure F.1 plots elasticity data for a synthetic Fibercane (Glottin) reed.
In figure F.2 the reed is also a natural cane reed, different from the previous figure,
and that had not been used before in experiments.
Figure F.3 shows data for the same reed as in figure F.2, but the reed was soaked in
water for 15 minutes before the experiment.
These measurements show that the linearity observed in section 3.4 is not perfectly
observed in other reeds than the synthetic plastic reed. The stiffness is seen to increase
slightly for small reed openings in some of the natural cane reeds.
Soaking the reed has the effect of linearising the elasticity data, and at the same time
increasing the visco-elastic effects.
190 Chap. F: Elasticity data

3.5

Reed opening area (mm²)


2.5

1.5

0.5

0
0 10 20 30 40
Pressure difference (kPa)

Figure F.2.: Measurement of the reed opening area against the applied pressure for a
dry natural cane reed (2)

anche4mouillée
4

3.5
Reed opening area (mm )

3
2

2.5

1.5

0.5

0
0 10 20 30 40
Pressure difference (kPa)

Figure F.3.: Measurement of the reed opening area against the applied pressure for a
wet reed (2)
G. Artificial mouth

G.1. Artificial mouth

In normal playing conditions, the reed is introduced between the instrumentist’s lips,
so that the observation of its motion is difficult (the part of the reed that remains
accessible from the outside of the mouth has a small amplitude of motion which is also
very different from the motion of the tip). Another problem about making measurements
with a real instrumentist is to ensure a constant control over the instrument (blowing
pressure, embouchure...). It is thus necessary to use a device that can apply the
same ranges of controls to the instrument while keeping them constant throughout
the experiments.

Such an experimental device, called artificial mouth [Almeida et al., 2004a] (fig. G.1),
was conceived and built at IRCAM by Alain Terrier, and is used in several experiments
requiring or not visualizations, as a complement to measurements performed on real
instrumentists. It was inspired by a device used by D. Ricot, R. Causs é and N. Misdariis
[Ricot et al., 2005] to study the single-reed of the accordion.

The artificial mouth is fed by a compressed air source (connected to the compressed
air circuit existing in the laboratory) which arrives directly into a cylindrical chamber.
On the top of this chamber, the air can flow through a transparent cylindrical tube
which has a 90 degree angle before its exit. The double reed is placed at the exit of this
tube by inserting it by the bottom side (the side that is usually inserted in the resonator)
into a rubber plug which fits securely to the exit of the tube.

When the reed is secured in this position, its tip remains close to the 90 degree
angle, and to the tube walls. To maximize the visual access to the reed tip and avoid
deformations by the curved wall of the tube, the angle is in fact replaced by a cubic
piece made of plexiglass. Two faces of the cube are connected to the tube while the
other four have polished windows through which the reed can be observed (figure G.2).

Lips are replaced by a soft matter (synthetic foam) which is pressed against the reed
by a pair of screws attached to the tube walls. The position of the screws acts as a
control which in fact adapts several parameters of the reed. Some of these parameters
can be measured directly (such as the reed opening at rest) but others must be deter-
mined inderectly through experiments (for example damping or effective mass). The lips
replacement might be used or not in the experiments.
192 Chap. G: Artificial mouth

Diaphragm
Double reed

Pressure sensor

Window for filming

Figure G.1.: Artificial mouth

Figure G.2.: Reed introduced into the artificial mouth for observation. Synthetic lips
are removed from the reed.
G.2 Stroboscope 193

G.2. Stroboscope
The reed can be observed inside the artificial mouth during its oscillation. Because this
motion is too fast to be directly observed or captured by a 25 Hz camera, a stroboscope is
used to examine the periodic motion of the reed at a frozen phase or in slow motion. The
frequency of the reed motion can slightly vary along the experiment, so that the sound
produced by the instrument is used as a synchronization signal for the stroboscope.
For this purpose, the sound signal is low-pass filtered.
When recording a motion picture, a homogeneous lighting can only be achieved if the
strobe flashes are contained within the exposure time of the camera. It is thus important
that the strobe frequency be a multiple of the camera frame rate. Most ordinary cameras
have a fixed rate (25 images per second), so that slight frequency adjustments have to
be performed on the instrument, which will automatically adjust the frequency of the
stroboscope. This is done either by selecting an appropriate note by closing or opening
tone holes or by doing slight adjustments of frequency with the embouchure. However,
an exact proportionality ratio between the strobe frequency and the camera image rate
cannot be maintained during the whole recording because of the above mentioned slight
deviations in the sound frequency. This issue is addressed in section H.2.

G.2.1. Aim of the device


The apparatus described in this section is used to observe and record on video the
reed behavior for different playing conditions. These videos, once digitized and post-
processed (see appendix H) are used to automatically estimate some geometrical features
of the reed opening, such as the width or the area. This is done in chapter 3.
194 Chap. G: Artificial mouth
H. Image Analysis

H.1. Introduction

H.1.1. Image analysis in musical instrument acoustics

Analysis of visual data is a very useful way of gathering information and quantitative
data about a physical phenomenon. On one side this approach can provide an overview
of the system behavior through a direct observation, on the other it is usually less
expensive than other methods based on sensors and far less intrusive. A survey on
various visualization methods and applications to musical acoustics is presented by
Molin [Molin, 2003].
For wind instruments, visual methods have been successfully used to study either the
motion of solid parts (mostly lips for brass instruments) [Martin, 1942],[Gokhshtein, 1981],
[Yoshikawa and Muto, 2003], [Copley and Strong, 1996], flows [Verge et al., 1994],[Skulina et al., 2003],
[de la Cuadra et al., 2005], [Vergez et al., 2005] or waves in fluids [Pandya et al., 2003].
Scientists have used visual observations for a long time in a qualitative way, or in
handmade measurements [Backus, 1963]. Improvements such as stroboscopic light-
ing, photography and video recording have extended the possibilities of visual analysis
(for instance, J. Gilbert in his thesis[Gilbert, 1991] used a video recording of strobe
motion of the reed to validate reed opening measurements through a capacitive de-
vice). D. Jorno [Jorno, 1995] proposed some quantitative analysis of lips opening of the
trombone player, which however required heavy handwork preparation (lips painted
in blue). However the same method has recently been applied to more simple setups
[Bromage et al., 2005]. Digital image has been introducing quantitative methods to
analyse systematically huge amounts of data.
This paper is devoted to the presentation of such an approach oriented toward the
investigation of double-reed functioning, for which there is a lack of experimental data
in the literature.
A key advantage of the methods hereby presented is their non-intrusiveness. In fact,
they allow the studied instrument to be adjusted at will in the experimental device (an
artificial mouth presented in section G.1) without the drawbacks of sensors applied on
the sensitive reed. The systematic analysis based on pictures of the reed, conjugated
with more traditional techniques to measure pressures or flows, provides a simple but
powerful method to study the behavior (sect. 6.3.1) and parameter values (sect. 3.4, 3.5)
for the double-reed model. Moreover, these quantitative analysis can be kept rather
simple so that even real-time measurements are possible.
196 Chap. H: Image Analysis

Image sequence
(gray levels or color)

lighting
balance

Binarization

B/W image
sequence

Identification of
reed opening

Area etc.
Width Height

Figure H.1.: Flowchart of the image analysis procedure (see figure H.6 for more details)

H.1.2. Challanges of automatic image analysis


Although measuring a distance by hand in a photograph is a simple task (because
the target region, corresponding to the reed opening is easily identified by eye), the
automation of this task requires several steps in order to identify and extract the target
region. The procedure flowchart of the image analysis is represented in figure H.1 and
detailed in figure H.6. Each step in the chart will be described in detail in the following
sections.
Original images, captured by the camera and then digitized, either color or gray level,
must be binarized, that is, transformed into black and white in a way that ideally the
reed corresponds to a region with one of these colors and is completely surrounded by
pixels with the other color (sect. H.3). The binarization criteria might vary from one
image to another if the lighting is not constant. This is why the lighting balance step
is introduced before binarizing the image (sect. H.2), so that a constant gray threshold
can be used in distinguishing pixels that will turn into black from the ones that will
turn into white. In the final step of the recognition (Identification of the reed opening)
the algorithm isolates the pixels corresponding to the reed opening (sect. H.4). Once
the reed opening region is identified, the algorithm can perform several measurements
upon it, as described in section H.5. Image processing operations are implemented with
the help of the Image Processing Toolbox of Matlab.

H.2. Lighting balance


The first operation done on the image is to adjust the pixel intensities so that the inten-
sity of a same material point on the image is kept approximately constant throughout
the whole movie. We propose two ways of balancing the image intensities:

Average intensity histogram: consists in retaining the histogram of a reference image,


and trying to transform each image of the film so that its histogram approximately
H.3 Binarization 197

matches the reference histogram.

Reference area intensities: a reference area of the images is given as an input parameter,
corresponding to a region where physical movements are small, and having the largest
possible range of gray-levels. The histogram of the reference area of a chosen image de-
fines the reference histogram. A one-to-one relation is determined that best transforms
the reference histogram into the histogram of the reference area in the current image.
This transformation may be extrapolated to take into account the whole range of gray
levels existing in the current image. The inverse of this transformation is applied to the
current image, to obtain an image whose lighting is closer to the original image.

H.3. Binarization

Considering that the former step was successful in returning a set of images with ho-
mogeneous lighting, it should be possible to determine an intensity level that stands
between the gray levels of the reed border and the ones of the reed opening. Binarization
sets pixels with gray-levels on one side of this threshold to ‘‘on’’ (1) and those below to
‘‘off’’ (0). The result of this operation is thus a binary (black and white) image, where
the reed opening should ideally show as a single-color region surrounded by pixels of
the other color as seen in the second frame of figure H.2 (the color of the reed opening
region can either be black or white depending on whether the lighting was made from
the front or from behind, through the reed channel).

Figure H.2.: First stages of the image processing: Original Image and Binarization

H.4. Identification of the reed opening

After binarization, several black non-connected regions usually persist on the white
background. A subset of all the black pixels corresponds to the reed opening. The aim
of this step is to separate the reed opening region from other regions that are accidentally
connected to it, to identify the several pieces which correspond to the reed opening and
then return the subset of black pixels corresponding to the reed opening.
Several operations may be involved in this step, and correctly selecting them is the
key to perform a correct analysis of the whole image sequence.
198 Chap. H: Image Analysis

H.4.1. Removing thin connections or cracks

Due to problems in the binarization phase, maybe because the choice of the binarization
threshold is not optimal for all images, the reed opening region might be connected to
another region in the background, or it might break into several regions. To cope with
this problem, morphological operations are used to remove small-scale components of
the image. Typically, the functions applied are morphological opening and closure.
They transform individual pixels depending on the values of surrounding pixels and of
a structuring element (see appendix H.7 for details), allowing that a small area of black
pixels turns into white ones if enough white pixels suround it.

Figure H.3.: Binarized image and result of the application of a morphological opening.
On the left we display the whole image of the reed opening, and on the right
a zoom of the area where the reed opening connects to the background

Figure H.3 shows how a morphological opening is used to avoid that the reed opening
region connects to the background (on the right edge of the reed).
These operations should however be reduced to a minimum, because they can alter
the shape of the detected reed opening. Indeed, an indepth comparison of both images
in figure H.3 would show that some pixels are missing from the left edge of the reed
opening region when it becomes too narrow.

H.4.2. Reducing the number of regions

A significant number of the smallest regions in the image is already removed by using the
procedure described in the preceding section. In some images, the background of the
image that surrounds the reed has the same color as the reed opening after binarization
(like in fig. H.3). This can interfere with the identification of the reed opening region.
These areas can be removed by contiguously filling connected regions from the corners
or borders of the image. This is also called flood-filling (see figure H.4, bottom frame).

H.4.3. Segmentation and region identification

The final step to identify the reed opening region consists in segmenting the image
obtained. The role of the segmentation is to identify different non-connected regions,
that is, to label all pixels belonging to each region with a same identifier.
H.4 Identification of the reed opening 199

Figure H.4.: Secondary stages of image processing after morphological recontitution:


Inversion and flood filling from edges

Then, the region of interest is selected among the other labeled regions by comparing
their centroids to a reference point. This point is either provided at start, as an input
parameter of the program or it is the centroid of the selected region in the previous
image. This last method also requires a reference point for the first image.

Small openings It is common that the image has some defects such as random noise,
parts of the reed which are slightly out-of-focus, or fibers pending from the reed. When
the reed opening becomes only a few pixels wide, these defects can split the region
corresponding to the reed opening in the binarized image into several non connected
regions (see figure H.5). To correct this, the size of the structuring element used in the
morphological operations (sect. H.4.1) would need to be exaggerated in a way that the
final shape of the reed opening would differ too much from reality.

Figure H.5.: Image of an almost closed reed (top) and demonstration of the splitting
issue of the reed opening region into several sub-regions.

In this case, several subregions need to be taken into account. Therefore, for these
particular cases a new detection method is used: the current binarized image is in-
tersected with the smallest reed opening region detected so far with the method seen
above. Thus, even if it consists of several non-connected regions, all the ‘‘on’’ pixels are
taken into account. We chose to switch to this method when the reed opening length
becomes smaller than a given ratio of the reference length (length of the reed opening
in the reference image). In figure 3.8, simultaneous measurement of pressure and reed
opening shows that the measured area is smooth, indicating that switching between
200 Chap. H: Image Analysis

the two methods does not induce any discontinuity, which is a way of validating our
approach.

H.5. Geometrical measurements

Once the correct region is identified, geometrical information about the reed opening is
extracted from the image: typically the area, length and width of the selected region (see
figure H.7 for an example).
In order to improve the precision in these measurements, the reed opening is approx-
imated by an equivalent ellipse that has the same area and second moments as the
detected region. The major and minor axis of this ellipse are considered to be respec-
tively the length and the width of the reed opening. In figure H.7, the equivalent ellipse
is traced over the reed opening region.
Another input parameter is needed to transform the measured lengths (in pixels) in
real lengths. The physical length of a pixel can be given directly, or by indicating the
dimension of a known object in the image. Of course, this object must be located in the
same plane as the reed opening, in order to avoid biases due to perspective.
An example of the measured reed opening area during an oscillation of a plastic reed
is show in figure 6.1.

H.6. Discussion

H.6.1. Difficulties encountered during image processing

Lighting issues

Most of the problems with the technique of image analysis described in this appendix
are solved by improving lighting conditions. However some limitations are due to the
fact that the reed is introduced in the artificial mouth, and only a few windows are
available for both observation and lighting.

Strobe/camera synchronization

In experiments which use stroboscopic lighting, the main difficulty arises from the syn-
chronization between the camera and the stroboscope, which has to be done manually.
Because the stroboscope is triggered by the sound generated by the instrument, which
is not precisely stable, some of the images in the video are illuminated for a fraction of
the stroboscope flash, or not at all. A partially illuminated image can be corrected using
the lighting balance described in section H.2, whereas a frame not illuminated at all by
the stroboscope is blurred by the motion of the reed. One may moreover suspect that
the light power provided by a single flash is not the same for all flashes.
H.6 Discussion 201

Colour image

Gray-scale image

III.A
histogram mapping

lighting balance

mapping on a test area

III.B
Binarization

III.C.1
closure

morphological filtering

opening

III.C.2
connectivity (flood filling)

III.C.3 III.D
segmentation within regions

identification of the closest region


match an ellipse on the retained region
to the reference centroid

Reed length / Yes determination of the region


ref. length < ratio? area (pixel count)

No

logical AND with determination of the


the reference region region dimensions

sum of the area update reference


of remaining regions region and centroid

physical dimensions

Figure H.6.: Detailed flowchart of the whole image analysis procedure (see figure H.1
fore a more synthetic overview)
202 Chap. H: Image Analysis

Figure H.7.: The equivalent ellipse traced over the reed opening region. Length and
width are considered to be this ellipse’s major and minor axis.

Histogram Enhancement

The selection of the balance method is also important. The reference area histogram
(see H.2) gives in principle better results than the average intensities (see H.2), because
the reference can be chosen as an area of the image which does not change during the
film other than for lighting reasons (typically, a piece of foam surrounding the reed is
chosen as the reference area). If the whole image is used to calculate lighting variations,
there are usually other causes to the variation of the histogram than the lighting, for
example the variation of the area of the reed opening.
However, attention should be paid to select a reference area with maximum contrast.
As a matter of fact, since the histogram mapping calculated on the reference frame is
applied to the whole image, it is better to have information on the whole intensity range.

Inhomogeneous contrast

Lighting balance corrections can solve lighting differences between frames in the film.
There are other problems, noticeable inside a single frame that prove more difficult
to solve. For instance, parts of the image supposed to have the same intensity level
(same material, same colors) appear with different levels due to non-uniform lighting or
shadows.
Although usually there is a strong contrast locally between the reed material and the
reed opening, lighting differences might make it difficult during the binarization stage to
find an intensity level that uniquely distinguishes the brightness of the reed from that
of the reed opening for the whole reed. When this occurs, only moving the light to a
different place can help in solving the problem...

Reed material

However, it can be very difficult to achieve ideal lighting conditions. For instance, the
reed material is somewhat translucent so that the light arriving from the outside of
the reed can also lighten its inside. Another problem can be that light arriving from
the front of the reed will directly reach its inside as well, smearing the reed’s contours
H.7 Notions on mathematical morphology 203

(similar problems are described by Story and Titze [Story et al., 1996] when analysing
MRI images of the vocal tract).

Consequences on the binary image

Variations of the gray-level intensity on the reed’s contours (or the reed opening) can
cause:

• Variations in the detected reed opening, if parts of the reed opening and parts of
its borders have overlapping gray levels;

• Connection between the reed opening region and the background, if parts of the
reed tip have gray levels situated above the selected threshold while others are
below (such issues were also reported for the analysis of MRI scans in vocal tract
measurements [Narayanan et al., 1995]);

• Splitting of the reed opening into several regions, if parts of the reed tip have gray
levels situated above the selected threshold while others are below;

It is important to keep in mind that the solution to one of these problems in one of
the images can induce new problems in other images.

H.7. Notions on mathematical morphology


The following morphological operations are on/off functions and can thus be seen as
a non-linear filtering. For all operations we previously define, X ⊂ R2 (the set of ‘‘on’’
pixels of an image) ans a structuring element Bx ⊂ R2 which is a ball centered on point
x.

• Erosion: The eroded of X by B is defined as Y , {x |Bx ⊂ X } and noted Y = E B (X ).

• Dilation: The dilated of X by B is defined as Y , {x |Bx ∩ X 6= 0} and noted


Y = D B (X )

• Opening: The opening of X by B is defined


 as the dilated by transposition of B of
BT B
the eroded of X by B : Y =, D E (X )

• Closure: The closure of X by B is defined as the eroded by transposition of B of


BT B

the dilated of X by B: Y , E D (X )
204 Chap. H: Image Analysis
I. Analysis and synchronisation of velocity
measurements

I.1. Signal synchronisation

In order to build the dynamic velocity profiles shown in section 6.4, the measured
signals need to be synchronised. This is achieved by using the simultaneous pressure
signals recorded using the B& K microphone. The microphone is maintained at a fixed
position throughout the measurement of the velocity profiles, so that the waveform does
not change considerably from one position of the velocity probe to another.
For each recording a time delay must does be applied so that they are synchronised.
The delay is determined by locating the maximum of the correlation between two pres-
sure recordings. Since the maximum delay possible corresponds to one period, this
maximum value is searched within the delays between 0 and the signal period. The
average period is determined using a simple estimation of the fundamental frequency
1
on the whole signal. The correlation is nevertheless calculated on a 10-period long
sample of the signal, for a sake of speed.
Since the playing frequency of the reed alone is situated around 1.2 kHz, the signal
period is not longer than about 40 samples (using a sampling rate of 44100 Hz). In order
to increase the precision of the determined time delay, the signal is oversampled 10
times. The precision could be increased further by implementing a fractional detection
of the delay.

I.1.1. Splitting the periods of a signal

In section I.1 we described the method for synchronising two signals recorded at different
times and for different probe positions. We now describe the method and the motivation
for synchronising different periods within a same signal.
For the determination of statistics shown below, it was necessary to compare several
periods of a same recording. For this reason, an algorithm was conceived to identify all
the points having a same phase on a signal.
The average fundamental frequency (f0 ) of the signal xi is first determined. This
frequency should not vary much along the signal, although it is not necessary that it is

1
The signal spectrum is multiplied by copies of itself compressed along the frequency axes by an integer
factor. This amplifies the fundamental frequency peak, provided that there is enough energy in the
harmonics. The fundamental frequency is then determined using a sufficiently powerful high-frequency
harmonic. This increases the precision on the estimated frequency.
206 Chap. I: Signal processing

strictly constant. As a parameter to the analysis, the algorithm accepts a base time (t 0 ).
Points with the same phase (φ0 ) as t0 are searched for all periods of the signal.
As a first estimation of the time t1 for which the phase is again phi0 , the algorithm
uses t10 = t0 + 1/f0 . This estimation is improved by calculating the correlation of signal
x ([t0 , t0 + N ]) with x ([t1 , t1 + N ]) and the maximum correlation (∆t )1 is found. This is
the correction to be brought to t10 (t1 = t10 + ∆t ). The procedure is repeated until the end
of the signal.

I.1.2. Period statistics


Period marks ti allow us to split the velocity signal into period. These can be placed in
an array xij of dimension T × N , where T = fs /f0 and N is the total number of periods in
the signal. For each phase of the period (and for each probe position) several statistic
PN
measurements can be determined, for instance its average (< x >i = N1 j=1 xij ) and
q
standard deviation (σx,i = N1 sumjN=1 (xij − < x >i )2 ).
The standard deviation of the signal gives an indication of the phases and regions
where the flow velocity is more unstable, that is, where turbulence is stronger.
In figure 6.5 some average profiles are represented for one period, along with the
standard deviation, calculated as described above. Figure 6.6 plots the profile evolution
(along the longitudinal axis) with the standard deviation represented as a color scale
(blue for small deviations, red for large ones).

I.1.3. Image analysis and reed opening area


In most cases, the reed opening oscillations was video recorded before and after the
velocity measurements. Simultaneously to the recording of each picture, an impul-
sive signal was recorded into a WAV file synchronised with the signal captured by the
microphone, in a similar manner to the procedure applied to the velocity measurements.
Since the period reference marks are available from section I.1.1, the picture syn-
chronisation signal can be used to determine the phase corresponding to each picture.
This way, it would be possible to reconstruct a slow-motion movie of the reed opening
movement, eventually by interpolating between images to obtain a periodic rendering
of the motion. However, it seemed more useful for the current analysis to synchronise
the measurements performed on each picture (see appendix H) with the velocity profile
evolution.
The result of this synchronisation is shown in figure 6.9. In figure 6.8 the same reed
opening evolution is shown together with the average profile evolution.

I.1.4. Flow reversal


The hot wire probe used in this work measures the heat loss in the wire (even if in this
system the wire temperature remains constant independently of the flow). The more
air runs in the vicinity of the probe the more heat is lost to it, because more fresh
air is available for the probe to transfer its heat to. This implies that the probe can
I.1 Signal synchronisation 207

only measure the absolute value of the velocity and not its direction. In the case of
mostly unidimesnional flow, it is still possible for the flow to reverse its sense while the
measured velocity remains constant.
In our measurements, flow reversal seems to be present at some points in the period
(figure I.1). Far from the reed output, signal variations are smooth, but as the probe
approaches the main flow, the amplitude of variation of the flow increases, and a second
peak becomes visible in each period. A stronger discontinuity is observed near the
minima that surround this second peak. This suggests that flow reversal takes place in
these regions.
Supposing that the flow is directed along the x axis, the recorded signal would be |u x |.
In this case, and supposing additionally that ux (t ) and its derivative are continuous,
flow reversal would be identified as a discontinuity in ∂t |ux (t )| near the time-axis (as
long as ∂u∂t
x
6= 0 when ux (t ) = 0).
In practice, flow reversal was identified as the two minimum points in each period of
the measured signal.
Nevertheless, an observation of a single signal for a specific probe position shows
that it never becomes 0. This can suggest the presence of a residual component of
the velocity orthogonalq to the x axis (u⊥ ). If this component is constant, the measured
velocity is thus u 0 = 2 . By knowing the flow reversal points (described above),
ux2 + u⊥
both the residual velocity and the main flow can be deduced.
This processing provides expected results (figure I.2) for most signals, however it is
problematic for some signals where more than two minima have identical values. For
example, when three minima have similar values, one of them is a true minima of u x
close to zero, and not a flow reversal point. One way to overcome this problem may
be to use the knowledge about flow reversal times for other probe positions next to the
problematic signal, however this was not tried.
After some tries, this processing was abandoned because the reversal points are in
some cases not identified by the automatic processing. In some cases it is also difficult
to identify them manually, and a global observation of the profile evolution (such as
fig. 6.6) is required. In practice, it is difficult to be sure that the apparent discontinu-
ities correspond to flow reversals, or that the orthogonal component is constant. As a
conclusion, it seems safer to analyse directly the measured signals with the knowledge
of these problems, and to extract a maximum of conclusions from the observation of
these signals
208 Chap. I: Signal processing

0 36.3974 72.9097 109.564


time (samples)

Figure I.1.: Plot of the flow velocity signals along a reed diameter. From bottom to top,
each line corresponds to one probe position. Horizontal axis represents the
time in seconds. Vertical dotted lines show period reference marks (with the
same phase). Thicker plots are the signals for the reed output boundaries.
I.1 Signal synchronisation 209

30

20

10
velocity (m/s)

−10

−20

−30
60 80 100 120 140 160 180 200 220 240 260
time (samples)

Figure I.2.: Unfolding the velocity signal


210 Chap. I: Signal processing
J. Topography methods

In several aspects it is important to know the internal geometry of the reed channel. For
instance, it can be useful both for considerations on the flow inside the reed or for an
acoustical characterisation of it, as a resonator of reduced dimensions (crucial in the
analysis of reed oscillations without the bore — chapter 6).
The inside of the reed is of difficult access, due to its reduced cross-section dimen-
sions. Therefore, a cast of the duct provides a better way of accessing the geometry of
the double-reed. Due to the neck formed at the junction between the staple and the
reed, it is usually necessary to destroy the reed in order to recover the cast.
Different casting materials were tried out. The preferred was a soft silicone which did
not adhere to the reed and could be recovered without much dammage when opening
the staple.

J.1. Hand measurements

The first apporoach to measuring the reed is to mark several lines along the reed,
perpendicular to the reed axis, and measuring height h and width w on these lines on
a digital photograph.
A first approximation of the cross-section can be obtained by approximating each reed
cross-section as an ellipse with axis lengths h and w. The cross-section area is thus
A = πhw.

J.2. Artisanal scanner

A different approach was to use an artisanal method of topography consisting in mea-


suring the deviation of a laser plane that incided over the reed.
The cast lied on an accurately moving table (A), whose displacement can be precisely
measured (B). The table and the cast are illuminated by a laser beam (C) expanded by
a transparent cylinder (E) and striking the table with a small angle (20 deg.). This laser
plane would trace an arc of a circumeference (due to the presence of E) over the table if
the cast wasn’t present (the baseline , A on figure J.2). Images are captured from above
the reed (D), the direction of observation being perpendicular to the table.
Several images (fig. J.2) of the laser plane are captured for different displacements
of the cast (about 1 mm). These correspond to sections of the reed parallel to each
other. The photographs are then inserted into the computer in order to extract the
beam deviation at each point.
212 Chap. J: Topography methods

Figure J.1.: Scanner setup

Figure J.2.: Laser line illuminating the reed cast and computer recognition of the line:
A – baseline; B – line deviated by the cast

From each photograph several points on an oblique section of the reed are extracted.
The complete set of photos provides a sampling of a part of the reed’s surface, however
in order to reconstruct cross-sections of the reed interpolation must be done.
J.2 Artisanal scanner 213

10

20

30

40

4 50
2

20 60
15

Figure J.3.: Three-dimensional reconstitution of a reed cast


214 Chap. J: Topography methods
Bibliography

[Agulló, 2001] Agulló, J. (2001). A time-domain description of the acoustical behaviour


of the tenora (catalan tenor shawm). In International Symposium in Musical Acoustics,
pages 45–50.

[Almeida et al., 2006] Almeida, A., Vergez, C., and Caussé, R. (2006). Experimental
investigation of reed instrument functionning through image analysis of reed opening.
Submitted to Acustica.

[Almeida et al., 2004a] Almeida, A., Vergez, C., and Caussé, R. (2004a). Experimental
investigations on double reed quasi-static behavior. In Proceedings of ICA 2004,
volume II, pages 1229–1232.

[Almeida et al., 2004b] Almeida, A., Vergez, C., Caussé, R., and Rodet, X. (2004b).
Physical model of an oboe: Comparison with experiments. In Proceedings of ISMA
2004, pages 112–115.

[Antunes and Piteau, 2001] Antunes, J. and Piteau, P. (2001). A nonlinear model for
squeeze-film dynamics under axial flow. In Symposium on Flow-Induced Vibration.
ASME.

[Atig, 2004] Atig, M. (2004). Non-linéarité Acoustique localisée à l’extrémité ouverte d’un
tube. PhD thesis, Université du Maine.

[Azad, 1996] Azad, R. S. (1996). Turbulent flow in a conical diffuser: a review. Exp.
Thermal and Fluid Science, 13:318–337.

[Backus, 1961] Backus, J. (1961). Vibrations of the reed and the air column in the
clarinet. Journal of the Acoustical Society of America, 33(6):806–810.

[Backus, 1963] Backus, J. (1963). Small-vibration theory of the clarinet. Journal of the
Acoustical Society of America, 35(3):305–313.

[Bailly and Comte-Bellot, 2003] Bailly, C. and Comte-Bellot, G. (2003). Turbulence.


CNRS Editions.

[Barjau and Agulló, 1989] Barjau, A. and Agulló, J. (1989). Calculation of the starting
transients of a double-reed conical woodwind. Acustica, 69:204–210.

[Bate, 1975] Bate, P. (1975). The oboe. Instruments of the Orchestra. W. W. Norton
and Co., 3rd edition.
216 Bibliography

[Benade, 1988] Benade, A. (1988). Equivalent circuits for conical waveguides. J. Acoust.
Soc. Am., 83(5):1764–1769.

[Benade, 1976] Benade, A. H. (1976). Fundamentals of musical acoustics. Oxford Univ.


Press.

[Bromage et al., 2005] Bromage, S., Campbell, M., and Gilbert, J. (2005). Experimental
investigation of the open area of the brass player’s vibrating lips. In Forum Acusticum
2005, pages 729–734.

[Brown, 1990] Brown, A. E. (1990). Some aspects of intonation in the bassoon. In


Proceedings of the institute of acoustics, volume 12, pages 777–782.

[Campbell et al., 2004] Campbell, M., Greated, C., and Myers, A. (2004). Musical in-
struments. Oxford Univ. Press, New York.

[Carral and Campbell, 2005] Carral, S. and Campbell, D. M. (2005). The influence of
relative humidity on the physics and psychoacoustics of a scottish bellows blown
border bagpipe chanter and reed. In Forum Acusticum 2005, pages 379–384.

[Casadonte, 1995] Casadonte, D. J. (1995). The clarinet reed: introduction to its biology,
chemistry, and physics. PhD thesis, Ohio State University.

[Christensen, 2003] Christensen, R. M. (2003). Theory of Viscoelasticity. Dover, 2nd


edition.

[Copley and Strong, 1996] Copley, D. and Strong, W. (1996). A stroboscopic study of
lip vibrations in a trombone. J. Acoust. Soc. Am., 99(2):1219–1226.

[Dalmont et al., 2002] Dalmont, J.-P., Ducasse, E., and Ollivier, S. (2002). Saturation
mechanism in reed instruments. In Proceedings of the 3rd EEA European Conress on
Acoustics, Seville.

[Dalmont et al., 1995] Dalmont, J.-P., Gazengel, B., Gilbert, J., and Kergomard, J.
(1995). Some aspects of tuning and clean intonation in reed instruments. Applied
acoustics, 46:19–60.

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

[Dalmont et al., 2003] Dalmont, J. P., Gilbert, J., and Ollivier, S. (2003). Nonlinear
characteristics of single-reed instruments: quasi-static volume flow and reed opening
measurements. J. Acoust. Soc. Am., 114(4):2253–2262.

[de la Cuadra et al., 2005] de la Cuadra, P., Fabre, B., Montgermont, N., and de Ryck,
L. (2005). Analysis of flute control parameters: A comparison between a novice and
an experienced flautist. In Forum Acusticum 2005, pages L27–L31.
Bibliography 217

[Debut, 2004] Debut, V. (2004). Deux études d’un instrument de musique de type clar-
inette :Analyse des fréquences propres du résonateur et calcul des auto-oscillations par
décomposition modale. PhD thesis, Université de la Mediterranée Aix Marseille II.

[Deverge et al., 2003] Deverge, M., Pelorson, X., C., V., Lagrée, P.-Y., Chentouf, F.,
Willems, J., and Hirschberg, A. (2003). Influence of the collision on the flow through
in-vitro rigid models of the vocal folds. J. Acoust. Soc. Am., pages 3354–3362.

[Ducasse, 2001] Ducasse, E. (2001). Modélisation et simulation dans le domaine tem-


porel d’instrument á vent á anche simple en situation de jeu: méthodes et modèles. PhD
thesis, Université du Maine, ENSAM, Esplanade des Arts et Métiers 33405, Talence
Cedex.

[Fabre, 1992] Fabre, B. (1992). La Production du Son dans les Instruments de Musique
à Embouchure de Flûte: Modèle Aéro-Acoustique pour la Simulation Temporelle. PhD
thesis, Univ. du Maine.

[Fabre et al., 1996] Fabre, B., Hirschberg, A., and Wijnands, A. P. J. (1996). Vortex
shedding in steady oscillations of a flue-organ pipe. Acta Acustica, 82:811–823.

[Fletcher and Rossing, 1991] Fletcher, N. H. and Rossing, T. D. (1991). The Physics of
Musical Instruments. Springer-Verlag, 1 edition.

[Fletcher and Rossing, 1998] Fletcher, N. H. and Rossing, T. D. (1998). The Physics of
Musical Instruments. Springer-Verlag, 2 edition.

[Fritz, 2004] Fritz, C. (2004). La clarinette et le clarinettiste: Influence du conduit vocal


sur la production du son (The clarinet and the clarinet-player: influence of the vocal tract
on sound production). PhD thesis, Université Paris 6, IRCAM; 1, Place Igor Stravinsky;
75004 Paris; France.

[Fuks, 1998] Fuks, L. (1998). From air to music: acoustical, physiological and perceptual
aspects of reed wind instruments playing and vocal-ventricular fold phonation. PhD
thesis, Kungl Tekniska Högskolan.

[Fuks and Sundberg, 1996] Fuks, L. and Sundberg, J. (1996). Blowing pressures in
reed woodwind instruments. TMH-QPSR, 3:41–56.

[Gazengel, 1994] Gazengel, B. (1994). Caractérisation objective de la qualité de justesse,


de timbre et d’émission des intstruments à vent à anche simple. PhD thesis, Univ. du
Maine, Le Mans, France.

[Gilbert, 1991] Gilbert, J. (1991). Etude des Instruments de Musique à Anche Sim-
ple: Extension de la Méthode d’Equilibrage Harmonique, Rôle de l’Inharmonicité des
Résonances, Mesure des grandeurs d’Entrée (Study of single reed musical instruments:
extension of the harmonic balance method, role of inharmonicity of resonances, mea-
surement of input parameters). PhD thesis, Univ. du Maine, Avenue Ollivier Messiaen;
72085 Le Mans Cedex 9; France.
218 Bibliography

[Gilbert et al., 2005] Gilbert, J., Dalmont, J.-P., and Guimezanes, T. (2005). Nonlinear
propagation in woodwinds. In Forum Acusticum 2005, pages 1369–1372.

[Gokhshtein, 1979] Gokhshtein, A. Y. (1979). Self-vibration of finite amplitude in a tube


with a reed. Sov. Phys. Dokl., 24(9):739–741.

[Gokhshtein, 1981] Gokhshtein, A. Y. (1981). Role of airflow modulator in the excitation


of sound in wind instruments. Sov. Phys. Dokl., 26(10):954–956.

[Goossens and Roxburgh, 1980] Goossens, L. and Roxburgh, E. (1980). Oboe. Yehudi
Menuhin music guides. MacDonald Futura Pub., 2nd edition.

[Heinrich, 1979] Heinrich, J.-M. (1979). The bassoon reed. IDRS Journal, 7:17–43.

[Helie, 2003] Helie, T. (2003). Unidimensional models of acoustic propagation in ax-


isymmetric waveguides. J. Acoust. Soc. Am., 114(5):2633–2647.

[Helie and Rodet, 2003] Helie, T. and Rodet, X. (2003). Radiation of a pulsating portion
of a sphere: Application to horn radiation. Acustica, 89:565–577.

[Hirschberg, 1995] Hirschberg, A. (1995). Mechanics of Musical Instruments, chapter 7:


Aero-acoustics of Wind Instruments, pages 229–290. Springer-Verlag.

[Hirschberg et al., 1990] Hirschberg, A., van de Laar, R. W. A., Marrou-Maurières, J. P.,
Wijnands, A. P. J., Dane, H. J., Kruijswijk, S. G., and Houtsma, A. J. M. (1990).
A quasi-stationary model of air-flow in the reed channel of single-reed woodwind
instruments. Acustica, 70:146–154.

[Hirschman, 1991] Hirschman, S. E. (1991). Digital waveguide modeling and simula-


tion of reed woodwind instruments. Master’s thesis, Stanford Univ.

[Hofmans, 1998] Hofmans, G. C. J. (1998). Vortex sound in confined flows. PhD thesis,
Technische Universiteit Eindhoven.

[Jorno, 1995] Jorno, D. (1995). Etude théorique et expérimentale de l’auto-oscillation


des lèvres en présence d’un couplage acoustique. Application aux instruments à anches
lippales (theoretical and experimental study of lip auto-oscillation in presence of an
acoustic coupling. application to lip-reed instruments). Master’s thesis, Universit é
Paris 6, DEA ATIAM, ICP, IRCAM; 1, Place Igor Stravinsky; 75004 Paris; France.

[Karal, 1953] Karal, F. C. (1953). The analogous acoustical impedance for discontinu-
ities and constrictions of circular cross section. J. Acoust. Soc. Am., 25(2):327–334.

[Keefe, 1982a] Keefe, D. H. (1982a). Experiments on the single woodwind toneole. J.


Acoust. Soc. Am, 72(3):688–699.

[Keefe, 1982b] Keefe, D. H. (1982b). Theory of the single woodwind toneole. J. Acoust.
Soc. Am, 72(3):676–687.
Bibliography 219

[Keefe, 1990] Keefe, D. H. (1990). Woodwind air column models. J. Acoust. Soc. Am,
88(1):35–51.

[Kergomard, 1995] Kergomard, J. (1995). Mechanics of Musical Instruments, chapter 6.


Elementary Considerations on Reed Driven Instruments, pages 229–290. Springer-
Verlag.

[Kilne and Abbott, 1962] Kilne, S. J. and Abbott, D. E. (1962). Flow regimes in curved
subsonic diffusers. J. Basic Eng., 84:303–312.

[Kopp, 2003] Kopp, J. (2003). Physical forces at work in bassoon reeds.


http://koppreeds.com/physicalforces.html.

[Kundu and Cohen, 2002] Kundu, P. and Cohen, I. (2002). Fluid Mechanics. Academic
Press, 2nd edition.

[Landau and Lifchitz, 1989] Landau, L. and Lifchitz, E. (1989). Mécanique des Fluides,
volume 6. Editions Mir, 2 edition.

[Ledet, 1999] Ledet, D. A. (1999). Oboe reed styles: Theory and practice. Indiana Univ.
Press.

[Levine and Schwinger, 1948] Levine, H. and Schwinger, J. (1948). On the radiation of
sound from an unflanged circular pipe. Phys. Rev., 73(4):383–406.

[Lizée, 2004] Lizée, A. (2004). Doublement de Période dans les Instruments à Anche
Simple du Type Clarinette; approches numérique et expérimentale (period doubling
in clarinet-like single reed instruments; numerical and experimental approaches).
Master’s thesis, Université Paris 6, IRCAM; 1, Place Igor Stravinsky; 75004 Paris;
France. Rapport de DEA.

[MacIntire et al., 1983] MacIntire, M. E., Schumacher, R. T., and Woodhouse, J. (1983).
On the oscillations of musical instruments. J. Acoust. Soc. Am., 74:1325–1345.

[Maganza, 1985] Maganza, C. (1985). Excitations non-linéaires d’un conduit acoustique


cylindrique. Observations de doublements de période précédant un comportement chao-
tique. Application à la clarinette. PhD thesis, Université du Maine.

[Maganza et al., 1986] Maganza, C., Caussé, R., and Laloë, F. (1986). Bifurcations,
period doublings and chaos in clarinetlike systems. Europhysics Letters, 1(6):295–
302.

[Martin, 1942] Martin, D. W. (1942). Lip vibrations in a cornet mouthpiece. Journal of


the Acoustical Society of America, 13(80):305–308.

[Merzkirch, 1974] Merzkirch, W. (1974). Flow visualization. Academic Press, New York.

[Midgley, 1976] Midgley, R., editor (1976). Musical instruments of the world. Paddington
Press.
220 Bibliography

[Molin, 2003] Molin, N. E. (2003). Visualization of sound fields with application to


musical instruments. In SMAC proceedings, pages 703–706.

[Morse and Ingard, 1968] Morse, P. M. and Ingard, K. U. (1968). Theoretical Acoustics.
Princeton Univ. Press.

[Narayanan et al., 1995] Narayanan, S. S., Alwan, A. A., and Haker, K. (1995). An ar-
ticulatory study of fricative consonants using magnetic resonance imaging. J. Acoust.
Soc. Am., 98(3):1325–1347.

[Nederveen, 1998] Nederveen, C. J. (1998). Acoustical Aspects of Musical Instruments.


Northern Illinois Univ. Press.

[Obataya and Norimoto, 1999] Obataya, E. and Norimoto, M. (1999). Acoustic proper-
ties of a reed (arundo donax l) used for the vibrating plate of a clarinet. J. Acoust. Soc.
Am., 106(2):1106–1110.

[Okwuobi and Azad, 1973] Okwuobi, P. A. C. and Azad, R. S. (1973). Turbulence in a


conical diffuser with fully developed flow at entry. J. Fluid Mech., 57(3):603–622.

[Ollivier, 2002] Ollivier, S. (2002). Contribution à l’étude des Oscillations des Instruments
à Vent à Anche Simple. PhD thesis, Université du Maine, Laboratoire d’Acoustique de
l’Université du Maine – UMR CNRS 6613.

[Oppenheim and Schaffer, 1989] Oppenheim, A. V. and Schaffer, R. W. (1989). Discrete-


time signal processing. Prentice Hall.

[Ouziaux, 2003] Ouziaux, R. (2003). Mécanique des fluides appliquée. CNRS Editions,
3rd edition.

[Pandya et al., 2003] Pandya, B. H., Settles, G. S., and Miller, J. D. (2003). Schlieren
imaging of shock wave fronts from a trumpet. J. Acoust. Soc. Am., 114(6):3363–3367.

[Parsons, 1997] Parsons, R., editor (1997). ASHRAE Handbook fundamentals. ASRAE.
Chapter 32.

[Pinard et al., 2003] Pinard, F., Laine, B., and Vach, H. (2003). Musical quality assess-
ment of clarinet reeds using optical holography. J. Acoust. Soc. Am., 113(3):1736–
1742.

[Ponthot, 1988] Ponthot, J.-P. (1988). Le roseau et la musique (Exposition, Hy ères,


Château San Salvador, 1988, chapter Le roseau du var et la facture d’anche contem-
poraine. Aix-en-Provence: ARCAM: Edisud.

[Rader, 2004] Rader, M. (2004). Oboe reed questions. Web page.


http://experts.about.com/q/Oboe-2282/Oboe-Reed-Questions.htm.

[Rayleigh, 1945] Rayleigh, J. (1945). The Theory of Sound. Dover, 2nd edition.
Bibliography 221

[Ricot et al., 2005] Ricot, D., Caussé, R., and Misdariis, N. (2005). Aerodynamic excita-
tion and sound production of blown-closed free reeds without acoustic coupling: The
example of the accordion reed. J. Acoust. Soc. Am., 117(4):2279–2290.

[Rocaboy, 1989] Rocaboy, F. (1989). Proposed model for reed action in the bassoon. J.
Catgut Acoust. Soc., 1(4):20–25.

[Roscoe, ] Roscoe, B. Some thoughts about oboe reeds. Web page.


http://www.doublereeds.com/.

[Scavone, 1997a] Scavone, G. (1997a). Digital waveguide modelling of woodwind tone-


holes. In Proceedings of the international computer music conference. Computer music
assoc.

[Scavone, 1997b] Scavone, G. P. (1997b). An acoustic analysis of single-reed woodwind


instruments with an emphasis on design and performance issues and digital waveguide
modeling techniques. PhD thesis, Stanford University.

[Scavone, 2002] Scavone, G. P. (2002). Time-domain synthesis of conical bore instru-


ment sounds. In 2002 International Computer Music Conference Proceedings.

[Schumacher, 1981] Schumacher, R. T. (1981). Ab initio calculations of the oscillations


of a clarinet. Acustica, 48:71–85.

[Shalita, 2004] Shalita, J. (2004). Making oboe reeds. (ebook).

[Shimizu et al., 1989] Shimizu, M., Naoi, T., and Idogawa, T. (1989). Vibrations of the
reed and the air column in the bassoon. J. Acoust. Soc. Japan, 10(5):269–278.

[Skulina et al., 2003] Skulina, D. J., Campbell, D. M., and Greated, C. A. (2003). Mea-
surement of the termination impedance of a tube using particle image velocimetry. In
SMAC 2003 Porceedings, volume 2, pages 747–750.

[Smith, 1992] Smith, D. H. (1992). Reed design for early woodwinds. Indiana Univ.
Press.

[Smith, 1986] Smith, J. O. (1986). Efficient simulation of reed-bore and bow-string


mechanisms. In Proc. Int. Computer Music Conf., pages 275–280.

[Sprenkle, 1961] Sprenkle, R. (1961). The Art of oboe playing. Summy-Birchard C.

[Story et al., 1996] Story, B. H., Titze, I. R., and Hoffman, E. A. (1996). Vocal tract area
functions from magnetic resonance imaging. J. Acoust. Soc. Am., 100(1):537–554.

[Thompson, 1979] Thompson, S. C. (1979). The effect of the reed resonance on wood-
wind tone production. J. Acoust. Soc. Am., 66(5):1299–1307.

[van Walstijn, 2002] van Walstijn, M. (2002). Discrete-time modelling of brass and wood-
wind instruments with application to musical sound synthesis. PhD thesis, University
of Edinburgh.
222 Bibliography

[van Walstijn and Campbell, 2003] van Walstijn, M. and Campbell, M. (2003). Discrete-
time modeling of woodwind instrument bores using wave variables. Journal of the
Acoustical Society of America, 113(1):575–585.

[van Zon et al., 1990] van Zon, J., Hirschberg, A., J, G., and Wijnands, A. P. J. (1990).
Flow throught the reed channel of a single-reed music instrument. In Colloque de
Physique, Colloque 2, volume 51-C2, pages 821–824.

[Verge et al., 1994] Verge, M. P., Fabre, B., Mahu, W. E. A., Hirschberg, A., van Hassel,
R. R., Wijnands, A. P. J., and Hogendoorn, C. J. (1994). Jet formation and jet velocity
in a flue organ pipe. Journal of the Acoustical Society of America, 95:1119.

[Vergez, 2000] Vergez, C. (2000). Trompette et Trompettiste: un Système Dynamique


Non-linéaire à Analyser, Modéliser et Simuler dans un Contexte Musical. PhD thesis,
Université Paris 6, IRCAM, 1 Place Stravinsky, 75004 Paris.

[Vergez et al., 2003] Vergez, C., Almeida, A., Causse, R., and Rodet, X. (2003). To-
ward a simple physical model of double-reed musical instruments: Influence of aero-
dynamical losses in the embouchure on the coupling between the reed and the bore
of the resonator. Acustica, 89:964–973.

[Vergez et al., 2005] Vergez, C., de la Cuadra, P., and Fabre, B. (2005). Analysis of flute
control parameters: A comparison between a novice and an experienced flautist. In
Forum Acusticum 2005, pages 545–549.

[Välimäki, 1995] Välimäki, V. (1995). Discrete-time modeling of acoustic tubes using


fractional delay lines. PhD thesis, Helsinky univ. of technology.

[White, 2001] White, F. M. (2001). Fluid Mechanics. McGraw-Hill, 4th edition.

[Wijnands and Hirschberg, 1995] Wijnands, A. P. J. and Hirschberg, A. (1995). Effect


of a pipe neck downstream of a double reed. In Proceedings of the International
Symposium on Musical Acoustics, pages 149–152. Societe Française d’Acoustique.

[Wilson and S., 1974] Wilson, T. A. and S., B. G. (1974). Operating modes of the clar-
inet. Jour. Acoust. Soc. Am., 56(2):653–658.

[Yefchak, 2005] Yefchak, G. (2005). Making oboe reeds. web page.


http://www.yefchak.com/reed/.

[Yoshikawa and Muto, 2003] Yoshikawa, S. and Muto, Y. (2003). Lip-wave generation
in horn players and the estimation of lip-tissue elasticity. Acustica, 89(1):145–162.

You might also like