A practical way to measure intonation quality of woodwind instruments using standard equipment without custom made adapters

Similar documents
Class Notes November 7. Reed instruments; The woodwinds

Correlating differences in the playing properties of five student model clarinets with physical differences between them

about half the spacing of its modern counterpart when played in their normal ranges? 6)

Physics HomeWork 4 Spring 2015

about half the spacing of its modern counterpart when played in their normal ranges? 6)

ANALYSING DIFFERENCES BETWEEN THE INPUT IMPEDANCES OF FIVE CLARINETS OF DIFFERENT MAKES

How players use their vocal tracts in advanced clarinet and saxophone performance

USING PULSE REFLECTOMETRY TO COMPARE THE EVOLUTION OF THE CORNET AND THE TRUMPET IN THE 19TH AND 20TH CENTURIES

Music 170: Wind Instruments

3b- Practical acoustics for woodwinds: sound research and pitch measurements

WIND INSTRUMENTS. Math Concepts. Key Terms. Objectives. Math in the Middle... of Music. Video Fieldtrips

Regularity and irregularity in wind instruments with toneholes or bells

Physics Homework 4 Fall 2015

Vocal-tract Influence in Trombone Performance

Saxophonists tune vocal tract resonances in advanced performance techniques

Physical Modelling of Musical Instruments Using Digital Waveguides: History, Theory, Practice

Does Saxophone Mouthpiece Material Matter? Introduction

Standing Waves and Wind Instruments *

Acoustical comparison of bassoon crooks

THE VIRTUAL BOEHM FLUTE - A WEB SERVICE THAT PREDICTS MULTIPHONICS, MICROTONES AND ALTERNATIVE FINGERINGS

Harmonic Analysis of the Soprano Clarinet

Prelude. Name Class School

Open Research Online The Open University s repository of research publications and other research outputs

NOVEL DESIGNER PLASTIC TRUMPET BELLS FOR BRASS INSTRUMENTS: EXPERIMENTAL COMPARISONS

A PSYCHOACOUSTICAL INVESTIGATION INTO THE EFFECT OF WALL MATERIAL ON THE SOUND PRODUCED BY LIP-REED INSTRUMENTS

Create It Lab Dave Harmon

The Acoustics of Woodwind Musical Instruments

The Acoustics of Woodwind Musical Instruments

Simple Harmonic Motion: What is a Sound Spectrum?

Physical Modelling of Musical Instruments Using Digital Waveguides: History, Theory, Practice

Transient behaviour in the motion of the brass player s lips

The Story of the Woodwind Family. STUDY GUIDE Provided by jewel winds

Measurement of overtone frequencies of a toy piano and perception of its pitch

Zooming into saxophone performance: Tongue and finger coordination

2. AN INTROSPECTION OF THE MORPHING PROCESS

How do clarinet players adjust the resonances of their vocal tracts for different playing effects?

Practice makes less imperfect: the effects of experience and practice on the kinetics and coordination of flutists' fingers

Sounds of Music. Definitions 1 Hz = 1 hertz = 1 cycle/second wave speed c (or v) = f f = (k/m) 1/2 / 2

Experimental Study of Attack Transients in Flute-like Instruments

ECE438 - Laboratory 4: Sampling and Reconstruction of Continuous-Time Signals

How do clarinet players adjust the resonances of their vocal tracts for different playing effects?

MODELING PRINCIPLES FOR TIME-DOMAIN SIMULATIONS OF WIND INSTRUMENTS IN PLAYING SITUATIONS

STEVE TADD WOODWIND REPAIRS (.co.uk)

TEXAS BANDMASTERS ASSOCIATION

Clarinet Assembling the Instrument

We applaud your commitment to arts education and look forward to working with you. If you have any questions, please don t hesitate to call.

The Interactions Between Wind Instruments and their Players

Forestwood Middle School Band Instrument Selection Guide

Music Source Separation

Shock waves in trombones A. Hirschberg Eindhoven University of Technology, W&S, P.O. Box 513, 5600 MB Eindhoven, The Netherlands

Why Engineers Ignore Cable Loss

Edinburgh Research Explorer

Sound energy and waves

How do clarinet players adjust the resonances of their vocal tracts for different playing effects?

Guide to Band Instruments

Section V: Technique Building V - 1

THE KARLSON REPRODUCER

Note on Posted Slides. Noise and Music. Noise and Music. Pitch. PHY205H1S Physics of Everyday Life Class 15: Musical Sounds

Consonance in Music and Mathematics: Application to Temperaments and Orchestration

Sample Teaching Inventory Database:

Toward a Single Reed Mouthpiece for the Oboe ( )

Uses of Fractions. Fractions

PHYSICS OF MUSIC. 1.) Charles Taylor, Exploring Music (Music Library ML3805 T )

This question will most likely be the favorite one asked by your prospective switcher to tuba. The answers are fairly simple indeed:

Clarinet. History Assembly Cleaning

2018 Fall CTP431: Music and Audio Computing Fundamentals of Musical Acoustics

The characterisation of Musical Instruments by means of Intensity of Acoustic Radiation (IAR)

Clarinet Basics, by Edward Palanker

Marion BANDS STUDENT RESOURCE BOOK

Clarinet Quarter-Tone Fingering Chart

Linear Time Invariant (LTI) Systems

HOW CAN MY STUDENTS SOUND LIKE PROFESSIONALS?

The role of vocal tract resonances in singing and in playing wind instruments

STEVE TADD WOODWIND REPAIRS (.co.uk)

Section VI: Chorales VI - 1

The Research of Controlling Loudness in the Timbre Subjective Perception Experiment of Sheng

Hidden melody in music playing motion: Music recording using optical motion tracking system

HOW TO SELECT A NEW CLARINET by Tom Ridenour

Flute. Selection 10% of students will be selected for FLUTE. Flute Player Personality Conscientious

DISPLAY WEEK 2015 REVIEW AND METROLOGY ISSUE

Instrument Recognition in Polyphonic Mixtures Using Spectral Envelopes

Clarinet Care. Parts of a Clarinet - Ten Clarinet Care Imperatives:

A Quick Anatomy of the Flute

CHAPTER 14 INSTRUMENTS

Making Medieval & Renaissance. Woodwind Instruments. Modern instruments. Medieval and Renaissance instruments

Open Research Online The Open University s repository of research publications and other research outputs

Jaw Harp: An Acoustic Study. Acoustical Physics of Music Spring 2015 Simon Li

Bass Clarinet Quarter-Tone Fingering Chart

Spectral Sounds Summary

Page 5 Lesson Plan Exercises Score Pages 64 74

The influence of Room Acoustic Aspects on the Noise Exposure of Symphonic Orchestra Musicians

Low-Noise, High-Efficiency and High-Quality Magnetron for Microwave Oven

Shock waves in trombones

Jewel M. Sumner High Marching Band 2015 Spring Auditions

Mini- Junior-Harmonie

Investigation of Digital Signal Processing of High-speed DACs Signals for Settling Time Testing

CTP 431 Music and Audio Computing. Basic Acoustics. Graduate School of Culture Technology (GSCT) Juhan Nam


When you open your case, this is what you should see: LOWER JOINT UPPER JOINT. Instrument Assembly

West Des Moines Elementary Band West Des Moines Community Schools Elementary Band

Transcription:

A practical way to measure intonation quality of woodwind instruments using standard equipment without custom made adapters W. Kausel and H. Kuehnelt Inst. f. Wiener Klangstil, Univ. f. Music, Anton von Webernplatz 1, A-1030 Vienna, Austria kausel@mdw.ac.at 507

Woodwinds are not much different from brass instruments in how intonation is determined by air column resonances. Nevertheless, it is easier to measure input impedance of brasses because the input cross-section at the mouthpiece rim is circular, plane and big enough to be easily coupled to standard measuring heads. In reed instruments the input cross-section is neither flat nor circular nor well defined. Flute instruments not even have any definable input cross-section as this is controlled by the player s lower lip. On top of that, flutes are played at an open end where some coupled impedance considerably influenced by the nearby environment like lips, mouth and face of the player has to be taken into account. Existing approaches to measure input impedance of woodwind instruments usually require custom made adaptors optimized to yield nearly accurate intonation results for a specific instrument and playing range only. The proposed new approach is to separate instrument head or mouthpiece from the body and measure both parts from their cylindrical ends. During the measurement a natural playing condition can be simulated at the now available end. Both measurements can be assembled computationally yielding useful intonation results as will be demonstrated for flutes and saxophones. 1 Introduction Acoustical input impedance measurements have become quite common for the analysis of brass wind instruments and it is agreed that such results do have considerable significance for main quality aspects of such instruments, like intonation and response. It is not a big technical challenge to couple a wide-band sound source with known source impedance to the circular cross-section of the mouthpiece and to measure the sound pressure close to that plane. Rugged commercial input impedance measurement systems for production tests in professional instrument manufacturing lines as well as for quality analysis of prototypes or for other diagnostic purposes in the workshops of traditional instrument makers exist for several years now [1][2]. Similar systems for woodwind instruments would be just as important but are still lacking. The main reason is, that there is no easy way to couple an impedance measurement head to the mouthpiece of any single or double reed instrument nor to the embouchure hole of a flute. Attempts have been made to replace the mouthpiece of e.g. a clarinet by a reed-less surrogate part with circular coupling cross-section which has been customized and may be even optimized in order to resemble the main acoustic properties of the original mouthpiece as perfectly as possible. This approach was working but unpractical. It turned out, that custom coupling parts had to be manufactured not only for instrument groups but rather for specific instruments. For flutes this approach is even less viable although it has been done experimentally [3]. First, the flute is played at an open end where acoustical characteristics of the near environment including the player s lips, nose and face cannot be neglected. Second, the player s embouchure varies significantly between playing registers and especially the area of the embouchure hole which is covered by the lips has to be taken into account. Therefore it was investigated whether woodwind instruments can be measured in parts using the natural separation plane between flute head and flute body (or between two saxophone or clarinet parts) as the reference plane (as shown in Figure 1). It will be shown that it is indeed feasible to mathematically derive the acoustical impedance at the reference plane starting from input impedance measurements of the separated instrument parts. The head part measurements can even be made while the player is pretending to play a certain note thus taking real embouchure parameters into account (see Figure 2). This is especially important for flutes in order to get realistic intonation results. Figure 3 shows flute body impedances measured with different fingerings. (a) Head (b) Body Figure 1: BIAS Head with flute parts. The player can simulate the embouchure during the measurement of the head. 2 Deriving acoustical properties of whole wind instruments by measuring their parts All resonances of a flute originate from standing waves inside the instrument. Pressure wavelets are propagated in the tube being partially reflected at all discontinuities 508

MOhm 1000 100 10 1 0,1 Headjoint 1, low register embouchure Headjoint 1, high register embouchure Headjoint 1, open embouchure hole 0,01 0 1000 2000 3000 4000 Hz Figure 2: Flute head impedances (different embouchure) Acoustics 08 Paris input impedance Z in in the time domain applying the inverse Fourier transform to both spectra, it will immediately become obvious that the latter exhibits strong periodicity with little decay caused by two strong reflections at the begin and at the end of the tube, while the former decays quickly as it does not have any reflection at the stimulated end absorbing perfectly all the waves traveling back into the source. The former could be thought of as an open-loop pressure transmission function while in the latter case the loop is closed by the characteristic impedance of the input cross-section. Fortunately there is a mathematical way to transform the input impedance Z in into the input impulse response IIR. It is the transform IIR(e jθ )= 1 Z 0 Z in e jθ 1+ Z 0 Z in e jθ (1) with Θ being the discretized frequency [4, p.41]. With the complex input impedance vector defined by Z in = Z mag e jzarg, IIR becomes MOhm 100 Body, b3 Body, g4 Body, c#5 10 1 0,1 0 1000 2000 3000 4000 Figure 3: Flute body impedances (different fingerings) Hz IIR = Z mag (cos(z arg )+i sin(z arg )) Z 0 Z mag (cos(z arg )+i sin(z arg )) + Z 0 (2) The characteristic impedance Z 0 of the open cross sectional area with the radius r (area S, air density ρ, speed of sound c) at the entry of the part is mainly a function of temperature temp in [ C] and can be approximated by Z 0 = ρc 428 (1 0.0017 temp) S r 2 (3) π The open-loop input impulse responses IIR head and IIR body of both parts can be multiplied in the frequency domain to obtain the open-loop input impulse response function IIR tot of the whole instrument. of the bore. The biggest discontinuity of the flute body is, of course, the first open tone hole or the open end of the instrument, if all holes are closed. The main reflection of the flute head when measured from the separation plane will be the open embouchure hole with its specific radiation impedance defined by its open, uncovered area and the acoustic properties of the near proximity given by the player s lips, nose, face and - to an again smaller extent - the room. A Dirac-impulse generated by a non-reflecting (perfectly impedance-matched) sound pressure source would reveal the complete acoustical characteristics of an instrument part, provided that the pulse response is recorded using a microphone at the separation plane until all higher order reflections have decayed to a negligible level. The Fourier-transform of such a pressure-to-pressure reflection function is here named input impulse response (IIR) function of flute head resp. flute body, and must not be mistaken with the corresponding input impedance functions. The input impedance function (Z in )is the frequency domain spectrum of a pressure response to a sound flow stimulation and is usually measured at a more or less perfectly reflecting source end. Comparing the input impulse response IIR and the IIR tot (e jθ )=IIR head (e jθ ) IIR body (e jθ ) (4) IIR tot can be interpreted as the primary pressure response spectrum of the complete instrument in the separation plane when it has been stimulated by a δ- pulse at that same position. Secondary reflections could therefore be calculated according to IIR 2 tot, tertiary reflections by IIR 3 tot, and so on. The total response would consist of the stimulus and an infinite number of reflections. This transmission function H ClosedLoop (Eq (6)) will exhibit all resonances of the assembled instrument which do not accidentally have pressure nodes in the separation plane (see Figure 4). H ClosedLoop =1+IIR tot + IIR 2 tot + IIR 3 tot +... (5) H ClosedLoop =1+ IIRtot n 1 = (6) 1 IIR tot n=1 The impedance Z cut at the separation plane could be calculated according to Z cut Z 0,cut = 1+IIR tot 1 IIR tot (7) 509

H CL 16 14 12 10 8 6 4 2 0 Body g1/2 + Head, low register embouchure Body g1/2 + Head, high register embouchure 0 1000 2000 3000 4000 Hz Figure 4: Closed loop transmission function of flute 3 Length Correction A certain complication is caused by the fact that the acoustically effective length of the whole instrument is not simply the sum of the acoustical lengths of its parts. Usually some centimeters of the length of either part is lost when the parts are joined together. This difficulty can be overcome mathematically as the measured input impedance is given by the required unknown impedance at the effective coupling cross section multiplied by the well-known transmission matrix of a nearly cylindrical wave guide. If this matrix is inverted, the unknown termination impedance required to assemble the flute can be calculated. Useful models for such wave guides are not reviewed here. A fairly general lossy transmission line model has been published by Mapes-Riordan in [5] based on the work of Keefe, Caussé, Benade et al. [6, 7, 8, 9, 10]. It has been reviewed by the author in [11] in a way ready to implement it in a computer program. The matrix ( ) a11 (jθ) a A = 12 (jθ) (8) a 21 (jθ) a 22 (jθ) describing the usually cylindrical part of the duct which will become ineffective after composition has to be inverted in order to calculate the desired unknown termination impedance Z out = p out u out. The input impedance Z in = pin u in can be measured directly. ( ) pin (jθ) = A(jΘ) u in (jθ) ( pout (jθ) u out (jθ) ) (9) It should be noted that it is not at all required for the eliminated part of the duct to be cylindrical. Any arbitrary bore shape can be eliminated by composing it from simple conical or cylindrical elements, multiplying their transmission matrices A i before solving for Z out : Z out = a 22Z in a 12 a 11 a 21 Z in (10) 4 Flute Intonation Two different flutes have been studied in greater detail to evaluate the proposed method of determining their intonation. The head-joints have been measured without player and with a player simulating an embouchure required to play the lower register and one required to overblow the notes. The body part was measured with all fingerings required to play a complete chromatic scale from B4 to D7. All measurements have been made with the BIAS measurement system. A two-second long chirp signal was used as a stimulus signal and it took a few seconds only to obtain calibrated impedance curves for each fingering, ready to be saved in the user s database. In automatic mode the player is prompted for all selected fingerings while all the other steps are repeated automatically for all notes. The BIAS software has also been used to do the length corrections of all body measurements and to combine all body measurements with all available head-joint measurements. This way one or more complete sets of flute transmission functions for all playable notes can be generated by a single command. The resulting intonation chart shows the intonation in Cent of all available notes in one diagram (see Figure 6). A comparison between the intonation of one of the instruments when played by an expert player and when composed from part impedances is shown in Figure 5. The head joint has been measured with the embouchure required to play the lowest resonance of the separated head joint with a closed end (a sharp A4) and with an embouchure to overblow this note into the twelfth (about an E6). From these plots it can be seen how important the coverage of the embouchure is to control the intonation and that it should be taken into account when the head-joint impedances are measured. Especially in the low register there is an almost perfect match between played and measured intonation values. In the higher register players have to adjust their embouchure quite significantly in order to correct the natural intonation offered by the instrument. This is mainly done by covering a bigger part of the embouchure hole with the lips thus reducing the effective cross-sectional area making the pitch more flat. It should be noted that there are additional parameters which do have influence on intonation when flutes are actually played. These parameters, like temperature distribution and chemical composition of air (CO 2 content, humidity) in the flute have been investigated by Coltman [12, 13] quite early. Even the jet itself seems to have considerable influence on air column resonance frequencies. It is being studied currently by means of Lattice-Boltzman simulations [14] and PIV (particle image velocimetry) experiments [15]. 5 Saxophone Intonation It has also been investigated whether the method can be adapted for saxophones. Especially alto and tenor saxophones seem to be well suited as they can be separated between neck and body. Preliminary results from 510

100 80 60 Flute 1, H CL with low register embouchure Flute 1, H CL with high register embouchure Flute 1, played intonation cent 40 20 0-20 -40 b3 c4 c#4 d4 d#4 e4 f4 f#4 g4 g#4 a4 a#4 b4 c5 c#5 d5 d#5 e5 f5 f#5 g5 g#5 a5 a#5 b5 c6 c#6 d6 d#6 e6 f6 f#6 g6 g#6 a6 a#6 b6 c7 c#7 d7 Figure 5: Intonation overview: Flute 1, A4 = 440Hz Figure 7: Altosax Schagerl, A4 = 443Hz, measured at the place of the reed Figure 6: Flute 2, A4 = 442Hz, embouchure for high register measurements done on two different alto saxophones indicate that application of the composition method does yield useful results for this type of wind instruments. The instruments have been measured using the BIAS system in two different ways. First, it was tried to determine their resonance frequencies in their assembled state by using a soft rubber pad to couple the mouthpiece without the reed tightly to the measurement head. Second, measurements have been made in the disassembled state where the reed was pressed to the mouthpiece using tape in a way to ensure air tight closure of the mouthpiece when measuring the neck part. The intonation charts for both instruments measured in the two different ways can be compared in the Figures 7..10. The results are quite consistent for both methods and expert players state that the measured values do agree with their subjective experiences. Especially when played with heavy reeds results seem to be realistic and useful. Anyhow, more thorough evaluations have still to be made. Figure 8: Altosax Schagerl, A4 = 443Hz, measured in separate parts 6 Conclusion It can be concluded that the proposed method of measuring the resonance frequencies of wood wind instruments by determining the input impedances of their parts and composing their acoustics numerically is feasible and practical. Results obtained this way are accurate and consistent with results obtained by other methods and they are in acceptable agreement with intonation figures obtained from actually playing these instruments. No customized adaptor parts are required to obtain useful intonation charts which makes this method suitable for quality control of wood wind instruments in a production environment. The matching between measured intonation values and intonation values obtained by human players de- 511

flutes. Journal of Sound and Vibration, 243(1):127 144, 2001. [4] David B. Sharp. Acoustic pulse reflectometry for the measurement of musical wind instruments. PhD thesis, University of Edinburgh, Physics Department, 1996. Figure 9: Altosax Selmer III, A4 = 443Hz, measured at the place of the reed [5] Dan Mapes-Riordan. Horn modeling with conical and cylindrical transmission-line elements. Journal of the Audio Engineering Society, 41(6):471 483, 1993. [6] Douglas H. Keefe. Woodwind Tone Hole Acoustics and the Spectrum Transformation Function. PhD thesis, Physics Dept., Case Western Reserve University, Cleveland, OH, 1981. [7] Douglas H. Keefe. Woodwind air column models. Journal of the Acoustical Society of America (JASA), 88(1):35 51, 1990. Figure 10: Altosax Selmer III, A4 = 443Hz, measured in separate parts pends on the type of instrument, the type of reeds used for playing them and, of course, the skill and experience of the players themselves. Some of these influences can be taken into account during the measurements because the player has a chance to simulate her or his embouchure while performing the head or neck measurement. Special fingerings (e.g. partly closed tone holes or loosely pressed keys) can also be simulated while measuring the body part. Acknowledgments The authors would like to acknowledge Werner Winkler as the original inventer of the described measurement method. Many years ago he suggested to separate wood wind instruments for impedance measurements and to recompose them mathematically. He even worked out most of the mathematical formalism required to do this. Unfortunately he never published this useful method which has now been implemented in the impedance measurement system BIAS 1. References [1] Gregor Widholm, Heinrich Pichler, and Thomas Ossmann. BIAS - A Computer Aided Test System for Brass Instruments. An Audio Engineering Society Preprint, (2834), 1989. [2] G. Widholm and W. Winkler. Evaluation of musical instrument quality by computer systems. examples of realisation. In Proceedings of the SMAC 93, pages 560 565, Stockholm, 1994. Royal Swedish Academy of Music. [8] Douglas H. Keefe. Acoustical wave propagation in cylindrical ducts: Transmission line parameter approximations for isothermal and nonisothermal boundary conditions. Journal of the Acoustical Society of America (JASA), 75(1):58 62, 1984. [9] René Caussé and et al. Input impedance of brass musical instruments comparison between experiment and numerical models. Journal of the Acoustical Society of America (JASA), 75:241 254, 1984. [10] Arthur H. Benade. Equivalent circuits for conical waveguides. Journal of the Acoustical Society of America (JASA), 83:1764 1769, 1988. [11] Wilfried Kausel. Bore reconstruction of tubular ducts from acoustic input impedance curve. IEEE Transactions on Instrumentation and Measurement, 53(4):1097 1105, 2004. [12] John W. Coltman. Resonance and sounding frequencies of the flute. The Journal of the Acoustical Society of America, 40(1):99 107, 1966. [13] John W. Coltman. Mode stretching and harmonic generation in the flute. Journal of the Acoustical Society of America (JASA), 88(5):2070 2073, 1990. [14] Helmut Kühnelt. Simulating the sound generation in flutes and flue pipes with the lattice-boltzmannmethod. In Proceedings of International Symposium on Musical Acoustics 2004 (ISMA 2004), pages 251 254, Nara, 2004. Acoustical Society of Japan. [15] A. Bamberger. Vortex sound of flutes observed with particle image velocimetry. In The Proceedings of the 18th International Congress on Acoustics (ICA 2004), Kyoto, 2004. ICA. [3] J. Wolfe, J. Smith, J. Tann, and N.H. Fletcher. Acoustic impedance spectra of classical and modern 1 www.bias.at 512