You are on page 1of 16

1 D.B.

Sharp

Leak detection in pipes using acoustic


pulse reflectometry
D.B.Sharp and D.M.Campbell
Department of Physics and Astronomy, JCMB, King’s Buildings,
University of Edinburgh, Edinburgh, EH9 3JZ, UK

Abstract
This paper discusses the measurement of the acoustical properties of tubular systems
using pulse reflectometry. The technique involves the injection of a sound pulse into
the object under investigation and the recording of the resultant reflections. Analysis
of the reflections gives information about the bore profile and input impedance of the
object. The effect of small leaks on both of these properties is investigated, leading
to a method for detecting the presence of a leak in a general tubular system. For the
special case of a cylindrical pipe, the position and size of the leak are also determined.

1 Introduction
The detection of leaks is of crucial importance in many industrial situations. One of
the standard methods of testing that a pipe is airtight involves sealing the end of the
pipe, increasing the air pressure within to a certain level and confirming that this level
remains constant. However, in industry, many systems are inaccessible and impossible
to seal, so in these cases a non-invasive technique is necessary. Here, an alternative
method of leak detection, based on the technique of acoustic pulse reflectometry, is
presented.
Originally developed for use in seismic studies [Ware & Aki, 1969], pulse reflec-
tometry has more recently been applied to the investigation of ducts of varying cross-
section. In the medical field, extensive research has been carried out into the use of the
technique for the measurement of airway dimensions [Jackson et al, 1977; Fredberg et
al, 1980; Brooks et al, 1984; Marshall, 1992]. At present, the main focus of the work
being carried out on ducts is in the field of musical instrument research. Pulse reflec-
tometry enables the measurement of the input impulse response of a wind instrument
[Smith, 1988; Watson & Bowsher, 1988]. Using a suitable reconstruction algorithm, a
reconstruction of the internal bore profile of the instrument can be evaluated from the
input impulse response. The input impedance of the instrument can also be calculated
[Watson, 1989].
The present paper is concerned with the detection of leaks in tubular systems. The
technique of pulse reflectometry is described, discussing in detail the measurement of
2 D.B.Sharp

the input impulse response of a general tubular object with varying cross-section. From
the input impulse response, an internal bore reconstruction and an input impedance
curve of the object can be evaluated. The increased accuracy of bore measurement
obtained by using the loss-compensating reconstruction algorithm recently proposed
by Amir, Rosenhouse and Shimony [1995] has made it possible to develop a technique
for the detection of small leaks in tubular systems. For the special case of a cylindrical
pipe, it is demonstrated that the axial position of a side hole can be identified and its
size evaluated.

2 Basic Technique
2.1 Determination of the input impulse response
The technique of pulse reflectometry involves injecting a pulse of sound into a tubular
object and digitally recording the resultant reflections. Analysis of these reflections
provides information about the internal bore of the object.

COMPUTER

D/A A/D

AMPLIFIER
AMPLIFIER
& FILTER

MICROPHONE
l1 l2

LOUDSPEAKER SOURCE TUBE OBJECT

Figure 1: Schematic diagram of pulse reflectometer

Figure 1 shows a schematic diagram of the pulse reflectometer used in the present
study. The reflectometer and test object are mounted in an anechoic chamber, with the
electronics in an adjoining control room. An electrical pulse (of width 80µs and voltage
5V) is produced by a 12 bit D/A converter (located on an Iotech DaqBoard 100A
data acquisition board inside a Viglen 486DX 66MHz PC). The pulse is amplified
3 D.B.Sharp

by a Pioneer A-119 stereo amplifier, and used to drive a Fane Professional MD2050
compression driver loudspeaker. The resultant sound pressure pulse travels along a 6m
long copper source tube (of internal radius 4.8mm and wall thickness 1.2mm) which
is clamped to a wooden board in a spiral of approximately 200mm radius. A Knowles
microphone embedded halfway along the tube records the reflections returning from
the tubular object under test, which is coupled to the far end of the source tube. The
microphone output is amplified by a second amplifier and low-pass filtered (using a
Barr and Stroud EF4-03 filter set to 20kHz) to prevent aliasing. The resultant signal is
then sampled by a 12 bit A/D converter (using a sampling frequency of 50kHz and a
sample length of 1024 points, giving a sample time of approximately 20ms) and stored
on the PC. The A/D converter is located on the DaqBoard 100A data acquisition board.
This procedure is repeated 1000 times and the samples are averaged to improve
the signal-to-noise ratio. Precise time alignment of successive samples is achieved by
triggering the sampler from the electrical pulse to the loudspeaker. A delay of 180ms
is included before each repetition to ensure all the signal from the previous step has
died away.
The source tube half-length l2=3m is necessary to ensure that the input pulse has
fully passed the microphone before the first of the returning object reflections reaches
it. The minimum duration of the input pulse is in practice limited by the requirement
that the pulse carries sufficient energy to ensure a good signal-to-noise ratio in the
measured reflections.
After the object reflections pass the microphone they are further reflected by the
loudspeaker. The source tube half-length l1 =3m is necessary to separate the object
reflections from these source reflections. The object reflections must be sampled over
a time period no longer than the time taken to travel the distance 2l1, to ensure that no
source reflections are recorded. This restricts the maximum length of objects that can
be measured to 3m in the present case.
For an ideal delta function sound pressure pulse, the reflections obtained from the
tubular object under test would be its input impulse response. However, the sound
pressure pulse is not ideal; to obtain the input impulse response, the reflections are
deconvolved with the input pulse shape. The input pulse shape is measured by termi-
nating the source tube with a flat perspex plate of thickness 5mm and recording the
reflected pulse (so taking losses along the source tube into consideration) [Sondhi &
Resnick, 1983].
The deconvolution is carried out by Fourier transforming both the sample contain-
ing the object reflections and the sample containing the input pulse (each of length
1024 points). A complex division of the object reflections by the input pulse is then
carried out in the frequency domain. A constraining factor q is added to the denomi-
nator to prevent division by zero [Marshall, 1990]:

R ω I  ω
IIR ω  (1)
I ω  I  ω  q
where R ω  is the transformed object reflections, I ω  is the transformed input pulse,
4 D.B.Sharp

I   ω is the complex conjugate of I  ω and IIR  ω is the transformed input impulse


response.
The resulting array is then inverse Fourier transformed to give the input impulse
response.

2.2 Bore reconstruction


The reflections returning from the object occur at changes in impedance, such as ex-
pansions or contractions along the object’s bore. A suitable algorithm (such as the
layer-peeling algorithm developed by Amir, Rosenhouse & Shimony [1995], which
compensates for attenuation due to losses) allows the reflection coefficients arising
from these impedance changes to be evaluated from the input impulse response. It
is then a small step to calculate the changes in area along the bore. If the object is
assumed to have cylindrical symmetry, the changes in radius can also be calculated.
However, DC offsets in the input pulse and the object reflections generally cause a
small DC offset in the input impulse response. This offset manifests itself by causing
the reconstruction to expand or contract too rapidly. Although pulse polarity alterna-
tion can reduce the offset, for accurate reconstruction the offset must be completely
removed.
Originally, the DC offset value was obtained by an iterative process of estimation
and adjustment, until the bore reconstruction coincided with a directly measured diam-
eter (at a known position towards the end of the tubular object). This output calibra-
tion method proved very time consuming, requiring a reconstruction to be calculated
at each step of the iteration.
The present method of determining the DC offset value involves the insertion of a
403mm long steel cylindrical connector (of internal radius 4.7mm and wall thickness
1.7mm) between the source tube and the object under investigation. Since there should
be no signal reflected from this cylindrical connector (as it contains no expansions or
contractions) the input impulse response should be zero. The average value of the
measured input impulse response over this range thus gives the DC offset value. This
input calibration method is far less time consuming, requiring only one reconstruction
to be made.

2.2.1 Results
Figure 2 shows the internal profile of a 356mm long stepped tube (whose radius ex-
pands from 4.7mm to 6.2mm to 9.25mm), reconstructed using the input calibration
method of DC offset evaluation. Note the presence of the 403mm long connector at
the start of the reconstruction.
The overshoots at each expansion in the reconstruction are due to the presence
of ripple in the input impulse response. This ripple, an example of the Gibbs phe-
nomenon, is introduced when the input pulse and the reflections are Fourier trans-
formed, because of the finite bandwidth of the transformed spectra.
5 D.B.Sharp

10.0
PULSE REFLECTOMETER
DIRECT MEASUREMENT
9.0

8.0
radius/mm


7.0

6.0

5.0

4.0
0.0 0.2 0.4 0.6 0.8

distance/m
Figure 2: Bore reconstruction of stepped tube

radius/mm

-5

5
0
-5
0 0.05 0.1 0.15 0.2 0.25 0.3 0.35
distance/m

Figure 3: 3D display of stepped tube


6 D.B.Sharp

Figure 3 shows a 3D display of the same stepped tube. This time, the connector is
not displayed.

2.3 Input impedance


The input impulse response is a measure of the amount of input signal reflected at
discrete distances along the tubular object. As the reflections are caused by changes
in impedance within the object, it is clear that the input impulse response and the
input impedance are closely related. Indeed, the complex input impedance of a tubular
object may be evaluated from its input impulse response [Watson, 1989]. The analysis
assumes plane wave propagation within the source tube.
In the time domain

p t  δ t  iir t (2)
zst  u t  δ t  iir t (3)
where p t is the pressure recorded by the microphone at time t, deconvolved with the
input pulse shape, u t is the volume velocity at the microphone at time t, zst  ρc  S is
the acoustic impedance of the source tube (ρ is the air density, c is the speed of sound
in air, S is the cross-sectional area of the source tube) and iir t is the input impulse
response of the object.
In the frequency domain this gives

P ω  1  IIR ω (4)
zst  U ω  1  IIR ω (5)
where P ω is the Fourier transform of p t , U ω is the Fourier transform of u t and
IIR ω is the Fourier transform of iir t .
Hence, the input impedance is given by:

P ω 1  IIR ω
z ω   zst  (6)
U ω 1  IIR ω
in

In the evaluation of the input impedance, it is not necessary to calculate the DC offset
in the input impulse response.

2.3.1 Results
Figure 4 compares impedance measurements using pulse reflectometry (diamonds)
with an impedance curve measured using a conventional frequency domain technique
(both curves are for the stepped tube displayed in figures 2 and 3). In the conventional
technique, a sine wave is passed down a high impedance capillary into the object being
measured. The high capillary impedance provides an approximately constant volume
7 D.B.Sharp

8
10
impedance/ohms

7
10

6
10
PULSE REFLECTOMETER
SWEPT SINE WAVE

0.0 0.5 1.0 1.5 2.0 2.5


frequency/kHz
Figure 4: Impedance curve for stepped tube

velocity which ensures that, to the same approximation, the pressure amplitude is pro-
portional to the input impedance. The pressure amplitude is measured by a microphone
and converted to the input impedance by dividing by the volume velocity. This is re-
peated for a whole range of sine wave frequencies [Backus, 1974; Campbell, 1994].
The frequency resolution of the impedance measurement using pulse reflectom-
etry is constrained by the sample length of the object reflections, which is in turn
constrained by the source tube length l1 . For l1 =3m, the maximum sample length is
approximately 20ms, leading to a frequency resolution of approximately 50Hz. To
improve the resolution by sampling over a longer period of time either the source tube
length l1 must be increased or the source reflections removed. Although increasing
the source tube length improves the resolution of the impedance measurement, the
signal experiences extra losses. This attenuation of the higher frequencies means that
the impedance can only be accurately calculated at low frequencies. At present, work
is being carried out on removing the source reflections (by driving the loudspeaker
in such a way as to cancel out the incoming object reflections). This should result
in improved resolution without reducing the range over which the impedance can be
measured.
8 D.B.Sharp

3 Detecting the presence of a leak in a tubular object


A small leak in the wall of the object being measured presents a reduction in the
impedance seen by the incoming pulse. This change is similar to the change in impedance
caused by a widening of the bore. Hence, the leak appears as an expansion in the bore
reconstruction.
This fact reveals another advantage of using the input calibration method of DC
offset calculation. A leaking object measured using this method will give a recon-
struction which is correct as far as the position of the leak (after which it will expand
to an extent dependent on the size of the leak). However, if the output calibration
method is used, the reconstruction will be forced to coincide with a measured diame-
ter towards the end of the object, despite containing spurious expansions. The whole
bore reconstruction will thus be corrupted.
Using the input calibration method provides a method of leak detection. If a bore
reconstruction made using this method passes through a measured diameter at a known
point towards the end of the object, the object can be considered airtight. If, however,
the reconstruction has a larger diameter than the measured diameter, there must be a
spurious expansion present and the object must contain a leak.

3.1 Results
Figure 5 shows bore reconstructions of a 501mm long cylindrical pipe (of internal
radius 5mm and wall thickness 1mm) both with and without a 0.5mm radius sidehole.
Again, note the presence of the 403mm long connector at the start of the reconstruction.
The expansion in the reconstruction of the cylindrical pipe with the hole is clear.
For the case of a cylinder, it is quite obvious that this expansion is caused by the leak
because the radius at the end of the pipe reconstruction should be equal to the radius
at the beginning. The position of the leak, 301mm from the output end of the pipe, is
also clear.
For an object whose bore profile is not known in advance, an expansion in the re-
construction due to a leak may be indistinguishable from an actual widening of the
bore. Hence, the position of a leak may be difficult to determine. However, the pres-
ence of the leak can still be confirmed by observing a discrepancy between the radius
of the reconstructed bore and the directly measured radius at the output end of the tube.

4 Evaluating the size of a leak in a cylindrical pipe


The size of a hole in a leaking cylindrical pipe can be calculated from the pipe’s com-
plex input impedance (measured as described in section 2.3), if the position of the leak
and the radius, length and wall thickness of the cylinder are known. In order to find
the expression for the hole size, the leaking cylinder is thought of as consisting of two
non-leaking cylinders, one before the hole and one after.
9 D.B.Sharp

10.0

8.0
radius/mm


6.0

4.0
CYLINDER WITHOUT HOLE
CYLINDER WITH HOLE
2.0
0.0 0.2 0.4 0.6 0.8 1.0

distance/m
Figure 5: Bore reconstruction of cylindrical pipe with and without sidehole

4.1 Impedance of a non-leaking cylinder

r
z input z load

Figure 6: Schematic diagram of non-leaking cylinder

Figure 6 shows a schematic diagram of a non-leaking cylinder. Assuming plane


wave propagation but including losses, the input impedance of an airtight cylinder of
length l and radius r is given by [Kinsler et al, 1982]:
10 D.B.Sharp

 
 zload kπr2 
j tan kl
ρω ρω 
zinput 
kπr2   (7)
zload kπr2
1 j tankl
 ρω
where zinput is the input impedance, zload is the load impedance (at the end of the cylin-
der), ρ is the air density, ω is the angular frequency, and k is the complex propagation
constant (k  k  jα).

1 ηω κω
k  ω  c and α  γ  1$
rc ! 2ρ #" ! 2ρc p %
γ is the ratio of the principal specific heats of air
c p is the specific heat of air at constant pressure
η is the coefficient of shear viscosity of air
κ is the thermal conductivity of air
c is the speed of sound in air

4.2 Impedance of the hole in a leaking cylinder


4.2.1 Expression of zh in terms of z1 and z2

zh lh
r
z0 z1 z2 z2rad

l1 rh l2

CYLINDRICAL CYLINDRICAL
SECTION 1 SECTION 2

Figure 7: Schematic diagram of leaking cylinder

Figure 7 shows a schematic diagram of a leaking cylinder. The complex impedance


at the end of the first cylindrical section, z1 , is made up of contributions from the input
impedance of the second cylindrical section, z2 , and the impedance of the hole, zh . The
impedance of the hole can be expressed in terms of the impedances z1 and z2 :
1 1 1 z2  z1
   (8)
zh z1 z2 z1 z2
11 D.B.Sharp

Thus
z1 z2
zh & (9)
z2 ' z1

4.2.2 Evaluation of z1
Equation 7 gives the input impedance of a cylinder with a complex load impedance
at its end. The first cylindrical section, of length l1, is such a cylinder, with an input
impedance z0 and a complex load impedance z1 . By substituting z0 for zinput , z1 for
zload , l1 for l and rearranging, an expression for the load impedance (i.e. the impedance
at the end of the first cylindrical section, z1 ) in terms of the measured input impedance,
z0 , is obtained:

ρω
z0 ' j tankl1
kπr2
z1 & (10)
z0 kπr2
1' j tan kl1
ρω

4.2.3 Evaluation of z2
Similarly, the impedance of the second cylindrical section can be calculated using
equation 7. The impedance at the input to this cylinder is z2 and, as the cylinder is
open-ended, the load impedance is the radiation impedance, z2rad . Substituting the
relevant values into equation 7 gives:
() ,-
)* z2rad kπr2 -
j tankl2
ρω ρω +
z2 & . (11)
kπr2 z2rad kπr2
1 j tankl2
+ ρω
In this case, the end is unflanged and the radiation impedance is given by [Kinsler
et al, 1982]:
1 ρω 2 2 ρω
z2rad & k r j0 / 6 kr (12)
4 kπr 2 + kπr2
so,
ρω 4 k r + j 1 0 / 6kr + tan kl2 2
1 2 2
z2 &
kπr2 0 1 1 ' 0 / 6kr tankl2 2
(13)
j 14 k2 r2 tankl2 3
+
4.2.4 Evaluation of zh
Substituting equations 10 and 13 into equation 9 gives the complex impedance of the
hole, zh , in terms of the measured impedance at the input to the leaking cylinder, z0 .
12 D.B.Sharp

4.3 Radius of the hole in a leaking cylinder


The complex impedance of a sidehole can also be expressed theoretically in terms of
its depth lh and radius rh [Kinsler et al, 1982; Keefe, 1982]:
ρωkh ρω 6 lh Erh 7
zh 4 j 5 (14)
4π 5 πrh2
rh 2
where E 4 1 8 595 9 0 8 58 : is the sum of the inner and outer end corrections for
r;
a hole set flush with the cylinder wall. The complex propagation constant is denoted
by kh 4 k 9 jαh because the wave is propagating in the hole. Hence, the attenuation
1 ηω κω
constant αh 4 6 γ 9 17 is dependent on the hole radius rh , not
rh c <= 2ρ 5 = 2ρc p >
the cylinder radius r.
Thus, the hole radius can be deduced from zh , measured as described in section
4.2.
Expanding equation 14 gives:

ρωk ρω ηω κω ρω rh3
zh 4 9 j 6 γ 9 17 8 9 8
r2 A
j l 1 595r 0 58
4πrh c <?= 2ρ 5 = 2ρc p > 5 πrh2 @ 5
h h

(15)
Considering only the imaginary part of the impedance zh :
ρωlh 1 8 595ρω 0 8 58ρωrh ρω ηω κω
zhimag 4 9 9 6 γ 9 17 (16)
πrh 5
2 πrh πr 2 4πrh c <= 2ρ 5 = 2ρc p >
Rearranging gives a cubic equation for rh :

1 8 724πr2 zhimag 2
rh3 r
5CB ρω D h
0 8 431r2 ηω κω
6 γ 9 17 9 2 8 75r2 rh 9FE 1 8 724lhr2 G 4 0 (17)
5 B c < = 2ρ 5 = 2ρc p > D
Substituting the value for zhimag (calculated in section 4.2 from the measured input
impedance) into equation 17 and solving [Press et al, 1988], yields a value for the hole
radius.
The success of equation 17 in predicting the radius of the hole will clearly depend
on a suitable choice of frequency. A small hole has little effect on the impedance of
the air column if it is in the vicinity of a pressure node; thus, at frequencies which
correspond to this condition the prediction of hole size from impedance measurement
can be expected to break down. Likewise, the hole has little effect on the acoustical
properties of the pipe at frequencies much above the cutoff frequency, which is deter-
mined by the radii of the main cylinder and the hole. This sets an upper limit on the
usable frequency range.
13 D.B.Sharp

4.3.1 Results

7.0
REFLECTOMETRY (1.0mm)
REFLECTOMETRY (1.5mm)
6.0 REFLECTOMETRY (2.5mm)
DIRECT MEASURE (1.0mm)
DIRECT MEASURE (1.5mm)
5.0
hole radius/mm

DIRECT MEASURE (2.5mm)

4.0
H
3.0

2.0

1.0

0.0
0.6 0.8 1.0 1.2 1.4 1.6 1.8 2.0 2.2

frequency/kHz
Figure 8: Comparison of predicted and measured hole radii

Figure 8 illustrates the predictions of equation 17 for three different sizes of hole
in a cylindrical pipe of length 501mm, internal radius 5mm and wall thickness 1mm.
In each case, the hole is 301mm from the output end of the pipe. Graphing the pre-
dicted radius as a function of frequency in this way is useful, since the divergences of
the prediction allow an immediate identification of the frequencies for which pressure
nodes coincide with the hole. In the present case, the divergences occur at integer mul-
tiples of 570Hz, as expected from the known geometry. The predictions also start to
break down, although less dramatically, above approximately 2000Hz, which is in the
vicinity of the cutoff frequency for the largest hole.
As a prescription for selecting valid data for predicting the effective radius of a leak
in an open ended pipe measured in this way, it is suggested that the frequency F of the
first divergence is estimated by inspection of the graph. Values of rh in the frequency
ranges 1 I 25F J 1 I 75F and 2 I 25F J 2 I 75F are then averaged to give a best prediction
of the effective radius of the hole. Predictions drawn in this way from Figure 8 are
compared in Table 1 with direct measurements using calipers. Six predicted values are
averaged in each frequency range; the uncertainties quoted are single standard devia-
tions. For each of the holes, the predicted radius agrees with that measured directly
within experimental error.
14 D.B.Sharp

Measured Predicted radius (averaged Predicted radius (averaged


radius over 712.5-997.5Hz) over 1282.5-1567.5Hz)
1.0 K 0.05mm 0.979 K 0.02mm 1.044 K 0.04mm
1.5 K 0.05mm 1.440 K 0.02mm 1.463 K 0.07mm
2.5 K 0.05mm 2.479 K 0.05mm 2.627 K 0.16mm
Table 1: Comparison of predicted and measured hole radii

5 Discussion and Conclusion


The technique of acoustic pulse reflectometry is a useful method of obtaining infor-
mation about the internal state of tubular systems without leaks. The insertion of a
cylindrical section of known radius between the source tube and the system under in-
vestigation allows the bore profile to be directly calculated from the measured input
impulse response. The input impedance of the system can also be evaluated, with-
out the necessity for either acoustic velocity measurement or calibration using known
acoustical systems.
For a system which has a small leak at an unknown point, it has been shown that
the pulse reflectometry technique can help to identify the presence of the leak, and
under favourable circumstances can give an estimate of both its axial position and its
effective area.
The technique has already proved its utility in the investigation of historical brass
musical instruments [Sharp et al, 1995]. The bores of these tubular systems are often
impossible to measure directly, and are prone to develop very small leaks. In the case
of a nineteenth century cornet, pulse reflectometry was used to identify a leak which
had not been evident in playing tests of the instrument; the axial information allowed
the leak to be located in a failed joint in a valve section.

Acknowledgments
This work has been supported by an EPSRC research studentship. The authors are
grateful to Prof.C.Greated and Dr.R.Parks for help in the early stages of the project;
Dr.N.Amir provided valuable assistance, including the provision of the reconstruction
algorithm prior to publication.

References
AMIR N., ROSENHOUSE G., SHIMONY, U. (1995) “A discrete model for tubular
acoustic systems with varying cross section - the direct and inverse problems. Part 1:
15 D.B.Sharp

Theory”, Acustica, 81, 450-462


AMIR N., ROSENHOUSE G., SHIMONY, U. (1995) “A discrete model for tubular
acoustic systems with varying cross section - the direct and inverse problems. Part 2:
Experiments”, Acustica, 81, 463-474
BACKUS J. (1974) “Input impedance curves for the reed woodwind instruments”,
J.Acoust.Soc.Am., 56, 1266-1279
BROOKS L.J., CASTILE R.G., GLASS G.M., GRISCOM N.T., WOHL M.E.B.,
FREDBERG J.J. (1984) “Reproducibility and accuracy of airway area by acoustic
reflection”, J.Appl.Physiol., 57, 777-787
CAMPBELL D.M. (1994) “The sackbut, the cornett and the serpent”, IOA Acoustics
Bulletin, 19(3), 10-14
FREDBERG J.J., WOHL M.E.B., GLASS G.M., DORKIN H.L. (1980) “Airway area
by acoustic reflections measured at the mouth”, J.Appl.Physiol., 48, 749-758
JACKSON A.C., BUTLER J.P., MILLET E.J., HOPPIN F.G., DAWSON S.V. (1977)
“Airway geometry by analysis of acoustic pulse response measurements”, J.Appl.Physiol.,
43, 523-536
KEEFE D.H. (1982) “Theory of the single woodwind tone hole”, J.Acoust.Soc.
Am., 72, 676-687
KEEFE D.H. (1982) “Experiments on the single woodwind tone hole”, J.Acoust.
Soc.Am., 72, 688-699
KINSLER L.E., FREY A.R., COPPENS A.B., SANDERS J.V. (1982) “Fundamentals
of acoustics”, 3rd ed., John Wiley and Sons, New York
MARSHALL I. (1990) “The production of acoustic impulses in air”, Meas.Sci.
Technol., 1, 413-418
MARSHALL I. (1992) “Acoustic reflectometry for airway measurement”, PhD The-
sis, University of Edinburgh, UK
PRESS W.H., TEUKOLSKY S.A., VETTERLING W.T., FLANNERY B.P. (1988)
“Numerical recipes in C”, Cambridge University Press
SHARP D.B., MYERS A., PARKS R., CAMPBELL D.M. (1995) “Bore reconstruc-
tion by pulse reflectometry and its potential for the taxonomy of brass instruments”,
Proceedings of the 15th International Congress on Acoustics, Trondheim, Norway, 26-
30 June, 1995, 481-484
SMITH R.A. (1988) “It’s all in the bore!”, J.Int.Trumpet Guild, 12(4)
SONDHI M.M., RESNICK J.R. (1983) “The inverse problem for the vocal tract: nu-
merical methods, acoustical experiments, and speech synthesis”, J.Acoust.Soc.
Am., 73, 985-1002
WARE J.A., AKI K. (1969) “Continuous and discrete inverse scattering problems in a
stratified elastic medium. I: Planes at normal incidence”, J.Acoust.Soc.Am., 45, 911-
921
WATSON A.P., BOWSHER J.M. (1988) “Impulse measurements on brass musical in-
struments”, Acustica, 66, 170-173
WATSON A.P. (1989) “Impulse measurements on tubular acoustic systems”, PhD The-
sis, University of Surrey, UK
16 D.B.Sharp

You might also like