0% found this document useful (0 votes)
30 views101 pages

Measurement of Temperature Using Sound-1

Uploaded by

nu834355
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as DOCX, PDF, TXT or read online on Scribd
0% found this document useful (0 votes)
30 views101 pages

Measurement of Temperature Using Sound-1

Uploaded by

nu834355
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as DOCX, PDF, TXT or read online on Scribd
You are on page 1/ 101

PROJECT REPORT

ON

Measurement of Temperature
Using Sound
[FOR THE PARTIAL FULFILMENT OF 6TH SEMESTER UG (PHYSICS) 2019-20]

Submitted by

Piyush

Roll No.- 011702PH032

UG-III (6TH semester), PHYSICS

Under the guidance of

DR. Smt. Pranati Kumari Rath

P.G. DEPARTMENT OF PHYSICSKHALLIKOTE (AUTO.) COLLEGE, BERHAMPUR


Certificate
“TO WHOM IT MAY CONCERN”

Certified that project report on “Measurement of Temperature using Sound”


submitted by
Piyush
Roll No.-011702PH032
UG 3rd year, 6th semester PHYSICS
KHALLIKOTE AUTONOMOUS COLLEGE, BERHAMPUR
has duly completed his/her assignment during the session 2019-20 under my
guidance and completed the assignment successfully.

For the partial fulfilment of UG 3rd year PHYSICS, encompasses the results of
studies carried out by him/her.

I wish him/her all success in life.

Date: Signature of the Guide


ACKNOWLEDGEMENT

I owe my deepest gratitude to Dr. Smt. Pranati Kumari Rath for being very
supportive and guiding me throughout the assignment and helping me
completing the project report on time, for partial fulfilment of 3rd year, 6th
semester PHYSICS.

I’d also like to thank Dr.Charupama Mishra, our esteemed HOD, PG


department of PHYSICS, for letting me do my project of interest and for being
supportive.

Signature of student
PREFACE

With the fast-growing world of knowledge, the challenge is to keep up


with the pace with which world is advancing. Mere acquiring the
knowledge in one field and specialising it would make you one
directional closing all the doors of exploration and letting you deserted
of ideas, hence interdisciplinary knowledge is one goal to achieve in this
age of global age of intelligence.
One of the best fields to realise these phenomenon, is science, to be
more General Physics, which has been long known for producing ideas
of interdisciplinary kind.

Signature of the student


Contents
1) Goals and Accomplishments

2) Physical basis of the method

3) Anisotropic diffusion filter for robust timing of ultrasound echoes

4) Pilot Scale Testing

5) Conclusion

6) Bibliography
Acronyms

The National Energy Technology Laboratory (NETL)

One-dimensional (1D)

Two-Dimensional (2D)

Three-Dimensional (3D)

Thermocouple (TC)

Resistance temperature device (RTD)

Coefficient of thermal expansion (CTE)

Fiber bragg grating (FBG)

Ultrasound (US) Nondestructive testing (NDT)

Speed of sound (SOS)

Time of flight (TOF)

Ultrasound measurements of segmental temperature distribution (US-MSTD)


Internal diameter (I.D.)

Signal-to-noise ratio (SNR)

Oxy Fuel combustor (OFC)


Partial stabilized zirconia (PSZ)

Tetragonal zirconia polycrystal (TZP)

Yttria stabilized zirconia polycrystal (Y-TZP)

Surface acoustic wave (SAW)

Abstract –
A gasifier’s temperature is the primary characteristic that must be monitored to
ensure its performance and the longevity of its refractory. One of the key
technological challenges impacting the reliability and economics of coal and
biomass gasification is the lack of temperature sensors that are capable of
providing accurate, reliable, and long-life performance in an extreme
gasification environment. This research has proposed, demonstrated, and
validated a novel approach that uses a noninvasive ultrasound method that
provides real-time temperature distribution monitoring across the refractory,
especially the hot face temperature of the refractory. The essential idea of the
ultrasound measurements of segmental temperature distribution is to use an
ultrasound propagation waveguide across a refractory that has been engineered
to contain multiple internal partial reflectors at known locations. When an
ultrasound excitation pulse is introduced on the cold side of the refractory, it
will be partially reflected from each scatterer in the US propagation path in the
refractory wall and returned to the receiver as a train of partial echoes. The
temperature in the corresponding segment can be determined based on
recorded ultrasonic waveform and experimentally defined relationship between
the speed of sound and temperature. The ultra sound measurement method
offers a powerful solution to provide continuous real time temperature
monitoring for the occasions that conventional thermal , optical and other
sensors are infeasible, such as the impossibility of insertion of temperature
sensor, harsh environment , unavailable optical path , and more. Our developed
ultrasound system consists of an ultrasound engineered wave guide, ultrasound
transducer/ receiver, and data acquisition, logging, interpretation, and online
display system, which is simple to install on the existing units with minimal
modification on the gasifier or use with new units. This system has been
successfully tested with a 100 kW pilot scale down flow oxyfuel combustor,
capturing in real time temperature changes during all relevant combustion
process changes. The ultrasound measurements have excellent agreement with
thermocouple measurements, and appear to be more sensitive to temperature
changes before the thermocouples response, which is believed to be the first
demonstration of ultrasound measurements segmental temperature
distribution across refractories.
Chapter 1
Introduction
Goals and Accomplishments
The main objective of this project is to develop and validate ultrasound(US)
techniques , insitu sensors, and measurement systems for non-invasive or
minimally invasive measurements of temperature distribution in solids, with
particular emphasis on measurements in extreme environments. We are
specifically interested in temperature measurement across refractories of
combustion and gasification processes and containments of other energy
conversion processes, which are often characterized by extreme operation
conditions, including high temperatures , pressures , mechanical abrasion,
chemical aggressiveness , and radiation exposure. This report describes the
successful development of such techniques and their experimental testing in the
laboratory and during pilot-scale operation of the coal-fired oxy-fuel
combustor .1.1 Motivation
Non-invasive or minimally invasive alternatives to the traditional methods of
temperature measurements are particularly useful when: a) insertion of
temperature probes is undesirable, difficult, or impossible; b) extreme
environments affect longevity of conventional sensors, as is the case for many
energy conversion processes; and c) when optical line-of-sight measurements
are not practical because the medium is opaque or optically dissipative. A
further advantage of ultrasound measurement is their sensitivity to the
temperature distribution, as opposed to point-wise temperature measurements
provided by conventional techniques. The US temperature measurements can
be implemented in all traditional transducer-receiver arrangements. The pulse-
echo mode, emphasized in this project, is particularly appealing because in this
arrangement a single device – an US transducer/receiver – is ultrasonically
coupled to the proximal end of the waveguide, which may be located outside an
aggressive environment, at a stand-off location. To illustrate the need for the
developed technology, we take a closer look at coal and biomass gasification as
an example of an energy conversion process that severely challenges the
utilization of traditional sensors.

1.2 Temperature
Temperature is a measure of the average kinetic energy of
the atoms or molecules in the system.
Temperature is the manifestation of thermal energy, present in all matter,
which is the source of the occurrence of heat , a flow of energy , when a
body is in contact with another that is colder. Temperature
is measured with a thermometer.
Thermometers are calibrated in various temperature scales that
historically have used various reference points and thermometric
substances for definition. The most common scales are the Celsius
scale (formerly called centigrade), denoted °C, the Fahrenheit
scale (denoted °F), and the Kelvin scale (denoted K), the latter of which is
predominantly used for scientific purposes by conventions of
the International System of Units (SI)..

When a body has no macroscopic chemical reactions or flows of matter


or energy, it is said to be in its own internal state of thermodynamic
equilibrium. Its temperature is uniform in space and unchanging in time.
The lowest theoretical temperature is absolute zero, at which no more
thermal energy can be extracted from a body. Experimentally, it can only
be approached very closely, but not reached, which is recognized in
the third law of thermodynamics.
Temperature is important in all fields of natural science,
including physics, chemistry, Earth science, medicine, and biology, as
well as most aspects of daily life.
Effects-
Many physical processes are affected by temperature, such as:
● the physical properties of materials including
the phase (solid, liquid, gaseous or plasma), density, solubility, vapor
pressure, electrical conductivity,
● the rate and extent to which chemical reactions occur,
● the amount and properties of thermal radiation emitted from the
surface of an object, and
● the speed of sound which is a function of the square root of the
absolute temperature.
Scale –
The Celsius scale (°C) is used for common temperature measurements in
most of the world. It is an empirical scale that was developed by a
historical progress, which led to its zero point 0 °C being defined by the
freezing point of water, and additional degrees defined so
that 100 °C was the boiling point of water, both at sea-level atmospheric
pressure. Because of the 100-degree interval, it was called a centigrade
scale. Since the standardization of the kelvin in the International System
of Units, it has subsequently been redefined in terms of the equivalent
fixing points on the Kelvin scale, and so that a temperature increment of
one degree Celsius is the same as an increment of one kelvin, though
they differ by an additive offset of approximately 273.15.
The United States commonly uses the Fahrenheit scale, on which water
freezes at 32 °F and boils at 212 °F at sea-level atmospheric pressure.

Absolute zero -
At the absolute zero of temperature, no more energy can be removed
from matter as heat, a fact expressed in the third law of thermodynamics.
At this temperature, matter contains no macroscopic thermal energy, but
still has quantum-mechanical zero-point energy as predicted by
the uncertainty principle. This does not enter into the definition of absolute
temperature. Experimentally, absolute zero can only be approached very
closely, but can never be actually reached. If it were possible to cool a
system to absolute zero, all classical motion of its particles would cease
and they would be at complete rest in this classical sense. The absolute
zero, defined as0K, is approximately equal to273.15°C, or459.67°F.
Absolute scales -
Referring to the Boltzmann constant, to the Maxwell–Boltzmann
distribution, and to the Boltzmann statistical mechanical
definition of entropy, as distinct from the Gibbs definition, for
independently moving microscopic particles, disregarding interparticle
potential energy, by international agreement, a temperature scale is
defined and said to be absolute because it is independent of the
characteristics of particular thermometric substances and thermometer
mechanisms. Apart from the absolute zero, it does not have a reference
temperature. It is known as the Kelvin scale, widely used in science and
technology. The kelvin (the word is spelled with a lower-case k) is the unit
of temperature in the International System of Units (SI). The temperature
of a body in its own state of thermodynamic equilibrium is always positive,
relative to the absolute zero.
Besides the internationally agreed Kelvin scale, there is also
a thermodynamic temperature scale, invented by Kelvin, also with its
numerical zero at the absolute zero of temperature, but directly relating to
purely macroscopic thermodynamic concepts, including the
macroscopic entropy, though microscopically referable to the Gibbs
statistical mechanical definition of entropy for the canonical ensemble,
that takes interparticle potential energy into account, as well as
independent particle motion, so that it can account for measurements of
temperatures near absolute zero. This scale has a reference temperature
at the triple point of water, the numerical value of which is defined by
measurements using the aforementioned internationally agreed Kelvin
scale.
International Kelvin scale -
Many scientific measurements use the Kelvin temperature scale (unit
symbol: K), named in honor of the physicist who first defined it. It is
an absolute scale. Its numerical zero point, 0 K, is at the absolute zero of
temperature. Since May, 2019, its degrees have been defined through
particle kinetic theory, and statistical mechanics. In the International
System of Units (SI), the magnitude of the kelvin is defined through
various empirical measurements of the average kinetic energies of
microscopic particles. It is numerically evaluated in terms of
the Boltzmann constant, the value of which is defined as fixed by
international convention.
Statistical mechanical versus thermodynamic temperature
scales -
Since May 2019, the magnitude of the kelvin is defined in relation to
microscopic phenomena, characterized in terms of statistical mechanics.
Previously, since 1954, the International System of Units defined a scale
and unit for the kelvin as a thermodynamic temperature, by using the
reliably reproducible temperature of the triple point of water as a second
reference point, the first reference point being 0 K at absolute zero.
Historically, the triple point temperature of water was defined as exactly
273.16 units of the measurement increment. Today it is an empirically
measured quantity. The freezing point of water at sea-level atmospheric
pressure occurs at approximately 273.15 K = 0 °C.
Classification of scales –
There is a variety of kinds of temperature scale. It may be convenient to
classify them as empirically and theoretically based. Empirical
temperature scales are historically older, while theoretically based scales
arose in the middle of the nineteenth century.
Empirical scales -
Empirically based temperature scales rely directly on measurements of
simple macroscopic physical properties of materials. For example, the
length of a column of mercury, confined in a glass-walled capillary tube, is
dependent largely on temperature, and is the basis of the very useful
mercury-in-glass thermometer. Such scales are valid only within
convenient ranges of temperature. For example, above the boiling point
of mercury, a mercury-in-glass thermometer is impracticable. Most
materials expand with temperature increase, but some materials, such as
water, contract with temperature increase over some specific range, and
then they are hardly useful as thermometric materials. A material is of no
use as a thermometer near one of its phase-change temperatures, for
example its boiling-point.
In spite of these limitations, most generally used practical thermometers
are of the empirically based kind. Especially, it was used for calorimetry,
which contributed greatly to the discovery of thermodynamics.
Nevertheless, empirical thermometry has serious drawbacks when judged
as a basis for theoretical physics. Empirically based thermometers,
beyond their base as simple direct measurements of ordinary physical
properties of thermometric materials, can be re-calibrated, by use of
theoretical physical reasoning, and this can extend their range of
adequacy.
Theoretical scales -
Theoretically-based temperature scales are based directly on theoretical
arguments, especially those of kinetic theory and thermodynamics. They
are more or less ideally realised in practically feasible physical devices
and materials. Theoretically based temperature scales are used to
provide calibrating standards for practical empirically based
thermometers.
Microscopic statistical mechanical scale -
In physics, the internationally agreed conventional temperature scale is
called the Kelvin scale. It is calibrated through the internationally agreed
and prescribed value of the Boltzmann constant, referring to motions of
microscopic particles, such as atoms, molecules, and electrons,
constituent in the body whose temperature is to be measured. In contrast
with the thermodynamic temperature scale invented by Kelvin, the
presently conventional Kelvin temperature is not defined through
comparison with the temperature of a reference state of a standard body,
nor in terms of macroscopic thermodynamics.
Apart from the absolute zero of temperature , Kelvin temperature of a
body in a state of internal thermodynamic equilibrium is defined by
measurements of suitably chosen of its physical properties, such as have
precisely known theoretical explanations in terms of the Boltzmann
constant. That constant refers to chosen kinds of motion of microscopic
particles in the constitution of the body. In those kinds of motion, the
particles move individually, without mutual interaction. Such motions are
typically interrupted by inter-particle collisions, but for temperature
measurement, the motions are chosen so that, between collisions, the
non-interactive segments of their trajectories are known to be accessible
to accurate measurement. For this purpose, interparticle potential energy
is disregarded.
In an ideal gas, and in other theoretically understood bodies, the Kelvin
temperature is defined to be proportional to the average kinetic energy of
non-interactively moving microscopic particles, which can be measured
by suitable techniques. The proportionality constant is a simple multiple of
the Boltzmann constant. If molecules, atoms, or electrons, are emitted
from a material and their velocities are measured, the spectrum of their
velocities often nearly obeys a theoretical law called the Maxwell–
Boltzmann distribution, which gives a well-founded measurement of
temperatures for which the law holds. There have not yet been successful
experiments of this same kind that directly use the Fermi–Dirac
distribution for thermometry, but perhaps that will be achieved in future.
The speed of sound in a gas can be calculated theoretically from the
molecular character of the gas, from its temperature and pressure, and
from the value of Boltzmann's constant. For a gas of known molecular
character and pressure, this provides a relation between temperature and
Boltzmann's constant. Those quantities can be known or measured more
precisely than can the thermodynamic variables that define the state of a
sample of water at its triple point. Consequently, taking the value of
Boltzmann's constant as a primarily defined reference of exactly defined
value, a measurement of the speed of sound can provide a more precise
measurement of the temperature of the gas.
Measurement of the spectrum of electromagnetic radiation from an ideal
three-dimensional black body can provide an accurate temperature
measurement because the frequency of maximum spectral radiance of
black-body radiation is directly proportional to the temperature of the
black body; this is known as Wien's displacement law and has a
theoretical explanation in Planck's law and the Bose–Einstein law.
Measurement of the spectrum of noise-power produced by an electrical
resistor can also provide an accurate temperature measurement. The
resistor has two terminals and is in effect a one-dimensional body. The
Bose-Einstein law for this case indicates that the noise-power is directly
proportional to the temperature of the resistor and to the value of its
resistance and to the noise band-width. In a given frequency band, the
noise-power has equal contributions from every frequency and is
called Johnson noise. If the value of the resistance is known then the
temperature can be found.
Macroscopic thermodynamic scale -
Historically, till May 2019, the definition of the Kelvin scale was that
invented by Kelvin, based on a ratio of quantities of energy in processes
in an ideal Carnot engine, entirely in terms of macroscopic
thermodynamics. That Carnot engine was to work between two
temperatures, that of the body whose temperature was to be measured,
and a reference, that of a body at the temperature of the triple point of
water. Then the reference temperature , that of the triple point, was
defined to be exactly 273.16 K. Since May 2019, that value has not been
fixed by definition, but is to be measured through microscopic
phenomena, involving the Boltzmann constant, as described above. The
microscopic statistical mechanical definition does not have a reference
temperature

Temperature conversions

from Celsius to Celsius

Fahrenheit [°F] = [°C] × 9⁄5 + 32 [°C]=([°F]32)× 5


⁄9

Kelvin [K] = [°C] + 273.15 [°C]=[K]273.15

Rankine [°R] = ([°C] + 273.15) × 9⁄5 [°C]=([°R]491.67)× 5


⁄9

Delisle [°De]=(100[°C])× 3
⁄2 [°C]=100[°De]× 2
⁄3

Newton [°N] = [°C] × 33⁄100 [°C] = [°N] × 100⁄33

Réaumur [°Ré] = [°C] × 4⁄5 [°C] = [°Ré] × 5⁄4


Rømer [°Rø] = [°C] × 21⁄40 + 7.5 [°C]=([°Rø]7.5)× 40
⁄21

1.3 Gasifier -

⮚ Here we have considered the case of a gasifier as it has a harsh condition


inside making difficult to measure temperature .
⮚ Q . What is a Gasifier ?

Ans. A gasifier is device used to convert converts organic- or fossil fuel-


based carbonaceous materials into carbon
monoxide, hydrogen and carbon dioxide. This is achieved by reacting the
material at high temperatures (>700 °C), without combustion, with a
controlled amount of oxygen and/or steam. The resulting gas mixture is
called syngas (from synthesis gas) or producer gas and is itself a fuel.
The power derived from gasification and combustion of the resultant gas
is considered to be a source of renewable energy if the gasified
compounds were obtained from biomass. The pocess is called Gasfication.

1.4 Current Temperature Measurement Techniques -


Direct Measurement

Thermocouples (TC) remain the most common temperature sensors used to


characterize gasification processes. Platinum-rhodium TCs are common as they
can provide measurements of extreme high temperatures. However, when
exposed to a harsh operating environment inside the slagging gasifier, such
thermocouples often fail within hours of insertion, leaving operators with no
real-time means for temperature measurements. The conventional approach to
harden the thermocouples is to utilize ceramic sheathing and improved filler
materials. However , the heavy sheathing makes such devices less sensitive to
dynamic changes in temperatures and the currently available heathing materials
are still susceptible to slag penetration and erosion damage. As a result, the
service life of the most hardened TCs is typically less than 120 days inside the
gasifier. Failure rates are often up to 50% within 15 days, and 75% within 30
days. This is especially true for entrained-flow slagging gasifiers since even the
most hardened sensors are unlikely to survive for more than 1 or 2 months as
the inner surface of the refractory wall degrades and recesses, exposing sensors
directly to the corrosive slagging environment. Once exposed to the
environment that exceeds material limits , TCs rapidly disintegrate. Even within
performance limits, thermal variations may lead to failure due to mismatch in
the coefficient of thermal expansion of thermocouple and sheathing materials.
Our interviews with U.S. operators of industrial gasification units indicate that
they are willing to trade-off the accuracy of temperature measurements for a
longer service life of insertion thermocouples. One consequence of these
priorities is the practice of recessed installation of thermocouples inside the
measurement ports. By placing the tip of the thermocouple withdrawn into the
refractory wall, the sensor is better protected from slag or erosion damage [5]
but measures the temperatures that may be 10’s or even 100’s degrees lower
than the temperature on the surface of the refractory.

Indirect Measurement

Secondary measurements that are relatively easy to obtain – such as


temperatures, pressures and compositions of streams into and out of a gasifier
– can be used in conjunction with empirical or theoretical models and
correlations to estimate inaccessible operating parameters inside the reaction
zone. The sensors used for such inferential measurement areusuallylocatedup-
anddownstreamfromthegasifierandthusawayfromitsextreme environment.
Texaco Inc. has developed a method to monitor the weight of the slag and infer
the process conditions (including the temperature) and use this information in
automatical control. The temperature measurements of the cooling water was
used in to estimate the heat flux away from the reaction zone. An attempt to
correlate a large number of routinely measured process variables to the
composition of the produced syngas was reported by Guo etal. . Higmanand van
derBurgt concluded that the temperature of a dry slurry feed gasifier can be
monitored by measuring the concentration ofCH4 orCO2 in the product gas. A
similar conclusion was reached in the computational study of Sarigul, which
showed close correlation between CH4 concentration and the adiabatic flame
temperature of the gasifier. The same approach was used to estimate
gasification temperature during the Tampa Electric Integrated Gasification
Combined-Cycle Demonstration Project and is believed to be in common use by
atleast some operators of gasification units in the United States who ,in
ourinterviews , reported temperatures in ppm of methane. Despite limited
current use, the inferential sensors remain a promising approach in gasification
and, more broadly, extreme environment applications. However, two
fundamental limitations of inferential measurements must be taken into
account. First, the quality of inferences critically depends on modeling errors
and uncertainties, and unmodeled changes to the process itself (e.g., due to
ware and aging), its feed , and unknown process disturbances. Second, the
measurement accuracy, sensitivity, and response time of inferential
measurement compare poorly with the corresponding characteristics of the
direct measurements. Therefore, the direct measurements in gasification will
continue to be desirable despite a long development time, high development
cost, and technical challenges that must be overcome.

Non Invasive Measurements

Noninvasive measurements do not require the direct or partial insertion of a


fragile sensing element into the harsh environment. The most widely used
techniques in this category are optical measurements, which can be used to
measure combustion temperature and gas composition. Optic sensors may be
based on the measurements of different properties, including optical reflection,
scattering, interference, absorption, fluorescence, and thermal radiation . The
methods for optical measurements of temperature in an extreme environment
has been reviewed in and include optical fiber Bragg grating(FBG) sensors,
optical resonat or cavity measurements , optical sapphire fiber-based
fluorescence sensors, optical single crystal sapphire-based sensor, among
others. Photosensitive FBG sensor and optical resonator cavity sensor, such as
Fabry-Pérot cavity-based sensor, use silica fibers which do not survive
temperatures over 1000◦C. When FBG sensor was used at temperatures close to
or above 1000◦C for hundreds of hours in air, silica single mode fibers lost their
mechanical strength and became extremely brittle , which made their
subsequent handling impossible. Single-crystal sapphire fibers have a melting
temperature of ∼2050◦C and can be used at higher temperatures. Most
sapphire fiber sensors are based on Fabry-Pérot structures generating broad-
band interference fringe pattern that can be monitored as a function of
temperature [. Sapphire based sensors have been fabricated and demonstrated
in high temperature measurements in laboratory experiments . Such sensors
will still have to survive the difficult environment of industrial gasifiers and are
still untested at these conditions. A well-controlled dopant density is important
for measurement accuracy , which puts strict requirements on the fabrication
process. It has been difficult to achieve high-quality measurement results, since
the interference signal from the Fabry-Pérot cavity are degraded by multi-mode
electromagnetic fields inside sapphire fibers. These systems during industrial
service are prone to mechanical failure due to mismatch in the coefficient of the
thermal expansion (CTE) between optical fibers and the sensing element.
Optical pyrometry is an established method for measuring temperatures of
flames. For several years, it was used by Texaco in its pilot gasification unit.
However, conventional total radiation or single-wavelength pyrometers cannot
provide accurate measurement of the flame temperature because an unknown
or non-uniform emissivity of the flame. The interference of the background
radiational so contributes to the difficulties . A two-color pyrometry removes
the limitation of an unknown and changing emissivity by using the irradiance
ratio at two carefully selected wavelengths. The application of such an approach
has been demonstrated in utility furnaces and different open flames, such as
premixed and diffusion flames . More recently, the multi-color method has also
been developed. Though minimally invasive (require a transparent access port),
optical line-of-sight techniques are not suitable for temperature and
composition measurements when an optically transparent line-of-sight is
difficult or impossible to maintain, as in the case of slagging gasification or when
high particle concentration in the reaction zone prevents light transmission.
Continuously blown (nitrogen) gas may be used to maintain the line of sight but
this will change the gas composition and the temperature of the reactor.
Consequently , commercial use of optical pyrometry in gasification is limited ,
intermittent , and is not seen as replacement of thermocouples.
Chapter 2
Physical Basis of the Method-
Sound
In physics, sound is a vibration that propagates as an acoustic wave,
through a transmission medium such as a gas, liquid or solid.
In human physiology and psychology, sound is the reception of such
waves and their perception by the brain. Only acoustic waves that
have frequencies lying between about 20 Hz and 20 kHz elicit an auditory
percept in humans. Sound waves above 20 kHz are known
as ultrasound and are not audible to humans. Sound waves below 20 Hz
are known as infrasound. Different animal species have varying hearing
ranges.
Definition-
Sound is defined as "(a) Oscillation in pressure, stress, particle
displacement, particle velocity, etc., propagated in a medium with internal
forces (e.g., elastic or viscous), or the superposition of such propagated
oscillation. (b) Auditory sensation evoked by the oscillation described in
(a)." Sound can be viewed as a wave motion in air or other elastic media.
In this case, sound is a stimulus. Sound can also be viewed as an
excitation of the hearing mechanism that results in the perception of
sound. In this case, sound is a sensation.
Physics of Sound-
Sound can propagate through a medium such as air, water and solids
as longitudinal waves and also as a transverse
wave in solids (see Longitudinal and transverse waves, below). The
sound waves are generated by a sound source, such as the
vibrating diaphragm of a stereo speaker. The sound source creates
vibrations in the surrounding medium. As the source continues to vibrate
the medium, the vibrations propagate away from the source at the speed
of sound, thus forming the sound wave. At a fixed distance from the
source, the pressure, velocity, and displacement of the medium vary in
time. At an instant in time, the pressure, velocity, and displacement vary
in space. Note that the particles of the medium do not travel with the
sound wave. This is intuitively obvious for a solid, and the same is true for
liquids and gases (that is, the vibrations of particles in the gas or liquid
transport the vibrations, while the average position of the particles over
time does not change). During propagation, waves can
be reflected, refracted, or attenuated by the medium.
The behavior of sound propagation is generally affected by three things:

● A complex relationship between the density and pressure of the


medium. This relationship, affected by temperature, determines the
speed of sound within the medium.
● Motion of the medium itself. If the medium is moving, this movement
may increase or decrease the absolute speed of the sound wave
depending on the direction of the movement. For example, sound
moving through wind will have its speed of propagation increased by
the speed of the wind if the sound and wind are moving in the same

direction. If the sound and wind are moving in opposite directions, the

speed of the sound wave will be decreased by the speed of the wind.

● The viscosity of the medium. Medium viscosity determines the rate


at which sound is attenuated. For many media, such as air or water,
attenuation due to viscosity is negligible.
When sound is moving through a medium that does not have constant
physical properties, it may be refracted (either dispersed or focused).
The mechanical vibrations that can be interpreted as sound can travel
through all forms of matter: gases, liquids, solids, and plasmas. The
matter that supports the sound is called the medium. Sound cannot travel
through a vacuum.
Longitudinal & Transverse Wave
Sound is transmitted through gases, plasma, and liquids as longitudinal
waves, also called compression waves. It requires a medium to
propagate. Through solids, however, it can be transmitted as both
longitudinal waves and transverse waves. Longitudinal sound waves are
waves of alternating pressure deviations from the equilibrium pressure,
causing local regions of compression and rarefaction, while transverse
waves (in solids) are waves of alternating shear stress at right angle to
the direction of propagation.
Sound waves may be "viewed" using parabolic mirrors and objects that
produce sound.
The energy carried by an oscillating sound wave converts back and forth
between the potential energy of the extra compression (in case of
longitudinal waves) or lateral displacement strain (in case of transverse
waves) of the matter, and the kinetic energy of the displacement velocity
of particles of the medium.
Sound Wave Properties & Characteristics –
Although there are many complexities relating to the transmission of
sounds, at the point of reception (i.e. the ears), sound is readily dividable
into two simple elements: pressure and time. These fundamental
elements form the basis of all sound waves. They can be used to
describe, in absolute terms, every sound we hear.
In order to understand the sound more fully, a complex wave such as the
one shown in a blue background on the right of this text, is usually
separated into its component parts, which are a combination of various
sound wave frequencies (and noise). Sound waves are often simplified to
a description in terms of sinusoidal plane waves, which are characterized
by these generic properties:

● Frequency, or its inverse, wavelength


● Amplitude, sound pressure or Intensity
● Speed of sound
● Direction
Sound that is perceptible by humans has frequencies from about 20 Hz to
20,000 Hz. In air at standard temperature and pressure, the
corresponding wavelengths of sound waves range from 17 m (56 ft) to
17 mm (0.67 in). Sometimes speed and direction are combined as
a velocity vector; wave number and direction are combined as a wave
vector.
Transverse waves, also known as shear waves, have the additional
property, polarization, and are not a characteristic of sound waves.

Ultrasonic Sound -
Ultrasound is sound waves with frequencies higher than 20,000 Hz (or 20 kHz) .
Ultrasound is not different from "normal" (audible) sound in its physical
properties, except in that humans cannot hear it. Ultrasound devices operate with
frequencies from 20 kHz up to several gigahertz.
Infrasonic Sound-
Infrasound is sound waves with frequencies lower than 20 Hz. Although sounds
of such low frequency are too low for humans to hear, whales, elephants and
other animals can detect infrasound and use it to communicate. It can be used to
detect volcanic eruptions and is used in some types of music.
2.2 Background -
Ultrasound (US), because of its high sensitivity, penetrating power, fast response
time, great accuracy, and noninvasive operation, has become a widely used
probing modality primarily in nondestructive testing (NDT). It can be used to
detect and characterize flaw, perform dimensional measurements, characterize
material properties, assess microstructural composition and the associated
mechanical properties . Ultrasound can be used in real time to provide a feedback
to control fabrication and synthesis processes ; the materials stability during
transport , storage and processing ; and the rate of degradation during service life
of the materials . The application of ultrasound in medicine includes diagnostic
imaging (sonography) and therapies . Sonography is used for evaluating the
condition of internal organs and tissues, commonly for neonatal fetuses, heart
imaging , and blood flow measurement. High intensity ultrasound is gaining
prominence as a method for non-invasive tumor ablation , targeted drug delivery,
and other therapies . Ultrasonic waves are stress waves at frequencies above 20
kHz that can propagate though gases, liquids, and solids . The propagation of a
harmonic wave can be described as

y(x,t)= y0cos(t2x ) , (2.1)

where y is the displacement caused by the propagating sound wave with respect
to distance, x, and time, t; y0 is the amplitude of the displacement, ω is the
angular wave frequency, and λ is the corresponding wavelength. The values of ω
and λ depend on the propagation medium; they are related to the speed of
ultrasound propagation (the speed of sound, SOS) and the frequency f as
2.2.1 Basic Ultrasound Instrumentation

Figure2.1 shows a block diagram of a basic ultrasonic measurement system used


to generate and detect ultrasonic waves in a solid specimen. The
synchronization generator sends trigger signals to the pulser, often with a high
repetitionrate. The pulser provides electrical excitation to the transducer which
produces ultrasonic waves at the same repetition rate. In the pulse-echo mode,
after traversing the specimen and reflecting from the interface, the reflected
ultrasonic waves (echoes) return to the receiver (which is if often the same
device as the transducer) and are converted to an electrical signal that can be
displayed and analyzed.

2.2.2 Speed of Ultrasound Propagation

In the traditional pulse-echo mode, the transducer-generated US pulse


propagates through as ample and produces an echo caused by its reflection
from the sample’s distal end relative to transducer location. For the sample
maintained at isothermal conditions, the velocity of ultrasound propagation, c,
can be found by measuring the time of flight (TOF), tof, of the returned echo:

where L is the distance between the proximal and the distal ends of the sample.
The speed of sound and its propagation modes depend on the transmitting
medium. In solids , the SOS depends upon the type of awave (e.g. , p-ors-
waves), the elastic properties of the medium, its density , and, in some cases,
the frequency. The ratio of applied stress (force/area) to axial strain
(extension/length) is the elastic (or Young’s) modulus, E, of the material.
Piezoelectric transducers can generate longitudinal (compression, also known as
primary wave or p-wave) and shear waves (s-wave), but one type will be
dominant depending on the specific design. The relationship between
longitudinal SOS and the elastic modulus in isotropic solids is given by

where K and G are the bulk modulus and shear modulus of the elastic materials,
respectively; E is the Young’s modulus, ρ is the density, and ν is the Poisson’s
ratio. The relationship between the Young’s modulus and the Poisson’s ratio has
the following form:

The SOS of s-waves is equal

The SOS of longitudinal waves depends both on the compression and shear
resistance properties of the material , while the speed of shear waves depends
on the shear properties only. Foralong , thin rod with a diameter shorter than a
wave length, the SOS of longitudinal waves may be simplified and is given by:

Young’s modulus of elasticity is sensitive to most structural changes and


material damage, including micro-cracking, densification, and phase transitions
between different crystal structures. The quantitative assessment of
microstructural changes can be carried out through the measurements of
ultrasound properties. Liquids and gases cannot resist shear stresses. The SOS in
fluids is expressed as
where β = 1/K is the compressibility of liquid. For an “ideal gas,” the SOS can be
shown to be

where γ is the adiabatic index, equal to the ratio of specific heats Cp of a gas at
a constant pressure to its heat capacity Cv at a constant-volume; P is the
ambient pressure and R is the molar gas constant.

2.2.3 Impedance and Attenuation

Acoustic impedance is defined as

and is an important characterization of the propagation medium. Another


important characteristic is the loss of the acoustic energy as the ultrasound
signal propagates through the medium. US attenuation [56,44] characterizes the
reduction in the amplitude of a ultrasonic waveform , and is defined as the
logarithm of the ratio of the magnitudes of the original to the attenuated
amplitudes, a and a0, expressed in decibels (dB):

Two types of processes contribute to attenuation of the ultrasound waveforms


propagating through the medium. The first process depends on the interaction
with the medium, which can cause beam spreading or focusing, energy
absorption, and dispersion. The presence of microscopic structural defects ,
such as point defects and dislocations, and macroscopic defects, affects the
degree of hardness and the elastic properties of the material and give rise to
absorption that occurs in both metals and nonmetals. The second process
contributing to the attenuation is due to wave interactions observed during
transmission across interfaces ; scattering by material variation and in
homogeneities , such as grains, grain boundaries, and pores; and the Doppler
effect .

2.2.4 Scattering

Scattering is the primary mechanism by which ultrasonic energy is lost during


propagation through many solids. Materials with significant scatter effect the
feasibility and effectiveness of engineered waveguides used in the proposed
method. Scattering is the process of redirecting an ultrasonicwave as a result of
the interaction between a primary ultrasonic propagating wave and the
anisotropic grains (inhomogeneities) inside of the medium [59]. Scattering
occurs when the dimension of grains is small relative to the wavelength
λ =c/f, where f is the central frequency of the propagating US . If physical
properties of grains and other in homogeneities , such as density or elasticity ,
are different from those of the bulk medium , the discontinuity in the velocity of
ultrasound propagation at the grain boundary will lead to the localized
reflections and loss of acoustic energy. The extent of scattering depends on the
number of grains per volume , their size relative to the wavelength (therefore ,
the scattering process is frequency dependent), and mismatch in acoustic
impedance at the boundaries

2.2.5 Reflections and Transmissions

When an ultrasonic wave encounters an interface , several phenomena may


occur , including reflection, transmission, refraction, and mode conversion.
These interactions are the phenomena our proposed approach relies upon to
create multiple ultrasound echoes from known locations along the length of the
US propagation. Ultrasonic waves are reflected at an interface of two media if
there is a mismatch in acoustic impedances (Z) across the interface . Assuming
the incident angle is normal to the interface, the fraction of the incident wave
intensity that is reflected can be characterized by the reflection coefficient
which depends on the acoustic impedances of the materials on either sides of
the interface:

where z1 and z2 are acoustic impedances of media 1 and 2. The transmission


coefficient, which characterizes the ultrasound energy transmitted through the
interface , is calculated as

Note that

indicating that the reflection and transmission coefficients represent the


percentage of acoustic energy which is either reflected or transmitted at a
boundary. The greater the impedance mismatch, the greater the percentage of
energy that will be reflected at the interface.

2.3 Physical Basis of the US-MSTD Method

Velocity of ultrasound propagation is temperature dependent in gases , liquids ,


and solids. This dependence is the physical basis of several temperature
measurement techniques, such as those reported in . Here, we are primarily
concerned with measuring temperature in solids. Though the velocity of US
propagation (colloquially referred to as the speed of sound, SOS) of all elastic
waves is temperature dependent, and thus can be used in thermometry, the
longitudinal (primary) waves were emphasized in this project. The US
techniques provide non-invasive or minimally invasive alternatives to the
traditional methods of temperature measurements. They are particularly useful
when:

• Insertion of temperature probes is undesirable, difficult, or impossible;

• Extreme environments affect longevity of conventional sensors, as is the case


for many energy conversion processes; and

• When optical line-of-sight measurements are not practical because the


medium is opaque or optically dissipative.

The US temperature measurements can be implemented in all traditional


transducer receiver arrangements. The pulse-echo mode, emphasized in this
project, is particularly appealing because in this arrangement, a single device –
an US transducer/receiver – is ultrasonically coupled to the proximal end of the
waveguide, which may be located outside an aggressive environment, at a
stand-off location. If a sample is maintained at a uniform temperature , by
measuring the time of flight , tof , over the known propagation path of length L,
the velocity of ultrasound propagation, c, can be calculated using equation(2.4).
If the relationship between the speed of sound and the temperature, T, is
known from theoretical considerations or empirical correlations,

then the measured tof can be used to obtain the sample’s uniform temperature
as :

where we assumed that the inversion of (2.16) is unique. Note that, in addition
to the time of flight measurements emphasized in this project, other ultrasound
characteristics, such as a phase change of echoes produced by at one burst
excitation , may be used to characterize temperature-dependent variations in
the speed of sound. The application of this basic idea to measure the
temperature in gases is known as acoustic pyrometry and is well established
and commercially used in many high temperature applications . The advantage
of the approach is the ability to obtain real-time temperature measurements
over an extremely large range of temperatures (from 0 to 3500◦F), which makes
it applicable to process monitoring from a cold start up to normal high-
temperature operation. Disadvantages include significant measurement
uncertainties when temperature along the propagation path between the
transducer and the receiver varies significantly and unknown changes in the
adiabatic constant due to variability in the gas composition. The utilized acoustic
frequency range is low (typically, ≤3kHz) because higher frequency ultrasound
does not propagate through gases. The consequence of low excitation
frequencies is interference from combustion instabilities (sounds produced by a
turbulent flow) and other disturbances, collectively known as a passive acoustic
signature. Such low frequencies also limit the achievable spatial resolution of
measurements when multiple transducers-receivers are used in order to
measure the temperature distribution inside of a containment. In addition to
high-temperature gas-phase measurement, acoustic thermometry was used to
detect temperature changes in the ocean by receiving low-frequency ultrasound
(< 100 kHz) transmitted across an ocean basin . Examples of ultrasound
measurements of the uniform temperature in solids are relatively few. Lee et al.
reported the development of an acoustical temperature measurement system
which uses the TOF measurements of an acoustic wave introduced into the
silicon wafer through an excitation quartz rod. The wave, partially reflected
from the quartz-silicon interface, travels through the wafer until reaching a
second quartz rod through which the wave reaches the receiver. The difference
between an arrival time of the reflected wave and the wave reaching the
receiver through the second rod gives the time of flight through the wafer,
which is used to estimate the wafer temperature. Lee et al. Reported that ±5◦C
accuracy was achieved in the range from room temperature to 1000◦C. Another
example of high temperature application of US thermome try is given by the
form of Balasubramaniam et al. who used an ultrasonic sensor to measure
viscosity and temperature of molten material up to 1000◦C with the
temperature resolution of 5◦C. US thermometry was also used in biomedical
applications. For example, Arthur et al. investigated the use of backscattered
ultrasound energy in temperature measurements to monitor and control
noninvasive thermal therapies of tumors. Using 7MHz linear ultrasound phased
array transducer, they demonstrated temperature measurements in ex vivo
phantom tissue from 37 to 50◦C in 0.5◦C steps. The project did not progress
towards in vivo testing be cause the quality of temperature measurements was
severely affected by subject motion, unavoidable in subjects due to breathing
and other disturbance.

Simonet al. Developed a 2D temperature estimation method based on the


detection of shifts in echo location of the backscattered ultrasound from a tissue
undergoing thermal therapy. The application of US thermometry becomes
problematic when the temperature of the sample is non-uniform. At the same
time, a non-uniform temperature distribution (often with large thermal
gradients) in containment sand other solid structures is the most relevant case
in most energy, defence, and other applications, as is the primary focus of this
project. When the temperature along the ultrasound propagation path is not
uniform, the measured time of flight of ultrasound echoes encodes the spatial
distribution of the temperature along the entire propagation path. The
relationship between the measured tof and an unknown temperature
distribution, T(z), is given by the following measurement equation: tof =2∫L 0 1
f(T(z)) dz. (2.18) Unlike the isothermal case described by (2.17), there are
arbitrary many temperature distributions T(z) such that, when used in (2.18) to
predict tof, the result will be identical to the measured time of flight.
Consequently, the deconvolution of the measurement equation (2.18) is an ill-
posed problem. The lack of unique dependence of the measured tof on the
temperature distribution presents a difficulty that may be resolved in one of
several ways. In the first approach, the problem is regularized by imposing
additional constraints on the temperature distribution. This has an effect of
parameterizing “admissible” temperature distributions by prescribing a
functional form that depends on one or more unknown parameters , which are
then found from ultrasound and, perhaps, other measurements.
Parameterizations may include: a)an assumption that the temperature along
the US propagation is constant and given by (2.17); b) the temperature
distribution is linear along the sample; and c) the requirement that T(z) satisfies
a heat transfer model with an appropriately selected parameters (such as
thermal conductivity) and boundary conditions.

Typically, an assumption of a constant temperature Ta across the ultrasound


propagation path , calculated from (2.17) using the measurements of tof ,
results in the least accurate temperature estimation. The relationship between
the calculate “average” Ta needed to match the measured TOF and the
unknown temperature distribution T(z) is given by the following equation

Such approximation is particularly problematic when strong thermal gradients


are present. This is the case when the temperature distribution across
refractories of energy conversion processes or other containments of extreme
environments are of interest. The TOF measurements alone are insufficient to
reconstruct the temperature distribution parameterized with more than a single
unknown. For example, the linear parametrization is determined by an
unknown slope and an intercept, both of which cannot be determined
simultaneously from a single measurement of tof.
Several examples of the deconvolution of the TOF measurements based on
thermal models are found in the literature. Takahashi and Ihara [77] assumed
that a one dimensional heat transfer model adequately describes the
temperature distribution along the length of the ultrasound propagation. They
further more assumed that the relationship (2.16) is linear, and the boundary
temperature at the transducer location is known (e.g., independently
measured). Therefore, only the boundary condition at a distal end of the sample
is needed to calculate the model prediction of the temperature distribution,
T(z), and the corresponding time of the flight, predicted by the measurement
equation (2.18). In temperature of the distal end was selected as the boundary
condition needed to completely define the heat conduction model. An
alternative boundary condition that specifies the heat flux at the distal end of
the sample was used in . In both cases, the selected distal boundary condition
(either the distal temperature, or the heat flux) was estimated as the value that
minimizes the difference between the predicated TOF and its measured value.
The temperature distribution through the sample was then found as the
solution of the obtained heat conduction problem.

An alternative to reducing the number of unknowns by parameterizing the


temperature distribution is to use more than a single TOF measurement in
temperature reconstruction. This is possible when multiple transducers and
receivers are used to measure the transmit and echo delays along different
ultrasound propagation paths, followed by the reconstruction of the
temperature distribution using a procedure similar to computer tomography.
System theoretical discussion of similarities of ultrasound and other noninvasive
measurement modalities, in which the acquired data depends on the spatial
integral of the property of interest, is given in.

The third distinct approach is to devise a way to extract more information from
each ultrasound transducer-receiver than a single time of flight measurement.
This is possible if each ultrasound excitation pulse produces multiple ultrasound
reflections, caused by echogenic features encountered as it propagates through
a sample. In an approach reminiscent to Bragg grating of optical fibers
Hanscombe and Richards proposed a method in which an ultrasound
waveguide is engineered to have regions of periodic ultrasound gratings, each
producing US echoes with different dominant frequencies determined by
grating dimensions.

Temperature variations change these dimensions through thermal expansion


and thus shift the frequency content of the echoes. By using gratings of different
dimensions at different spatial locations along the waveguide, multiple
frequency-separated echoes are produced, each encoding temperature changes
in a different spatial location. In this approach, the cross-talk between
overlapping echoes is reduced. However, the frequency content of echoes
produced by grating zones is insensitive to the temperature distribution
between them. The described approach apparently has not been tested in
experiments, leaving many unanswered questions. For example, it is not clear
how long each grated zone should be to insure narrow frequency content of
each echo. The achievable accuracy as a function of grating design also remains
unknown. The accuracy of temperature measurements when thermal gradients
are present within each grated zone , leading to a wider frequency band of each
echo, has not been quantified, as well.The method for ultrasound
measurements of segmental temperature distribution (USMSTD) in solids
developed during this project may be viewed as an approach which extracts
more than a single TOF data from each US excitation. In this approach, the
ultrasound propagation path consists of n segments defined by echogenic
features bounding each segment. The difference in the time of flight between
consecutive echoes, produced by partial ultrasound reflections from the
echogenic features, is used to estimate the temperature distribution across the
segment. Different parameterizations of segmental temperature distribution
may be used. Therefore, the US-MSTD method may be viewed as the
development of a hybrid approach which extracts more information from each
ultrasound excitation than a single time of flight measurement and
deconvolutes the information encoded in the TOF of each echo by
parameterizing the segmental temperature distributions.

2.4 Method

The US-MSTD method is graphically summarized in Fig. 2.2 for the case when
the temperature across the containment of an aggressive process is measured.
Its essential components are:

•The structured ultrasound propagation paths with echogenic features creating


partial ultrasound reflections from known locations;

• An ultrasound instrumentation needed to create the excitation pulse and


receive the train of resulting echoes;

• The signal processing of the ultrasound wave forms needed to accurately


determine the speed of sound or its change in different segments of the
propagation path by measuring the echoes’ time of flight;

• The relationship between the SOS and the temperature; and

• The method to translate the segmental speed of sound into the temperature
distribution such that the predicted time of flight, according to the
measurement model (e.g., equation 2.18), matches the measured values.

In the described approach, the sensitive electronic components are kept away
from harsh gasification environments and it is only required that the US
transducer be acoustically coupled to the cold side of the refractory,
representing minimal modifications to the gasifier. In the following, the
components of the method are discussed in more details.

2.4.1 Structure US Propagation Path


The central idea of the US-MSTD method is to have an US propagation path that
incorporates echogenic features at known locations that redirect a portion of
the US energy of the excitation pulse back to the transducer where the train of
echoes is recorded. The TOF of these echoes encode the temperature
distribution with finer granularity by providing information specific to the
temperature distribution within individual segments.

For cementitious samples, several methods for producing partial internal US


reflections were investigated . It was found that inclusions, stratifications, and
variations in the waveguide geometry are the adequate means in creating the
structured US propagation path. For example, it was found that by casting
multiple cementitious layers of identical composition and allowing time for a
partial curing before consecutive castings, enough variation in acoustic
impedance is introduced to create partial US reflections at the interface of the
layers. With such implementation, thermal, chemical, and mechanical
properties remain essentially constant throughout the structured material. In
another example, echogenic features were obtained by drilling small holes
along the length of a ceramic (alumina) waveguide.

An alternative to purposefully structuring the waveguide, the echogenic


features occurring naturally in the material or introduced for reasons unrelated
to the needs of the US-MSTD method may also be used. These may include
natural stratifications, natural inclusions, or geometric features, such as a rifling
inside a gun barrel used in to estimate the temperature variations caused by
firing a gun. It is only required that the spatial
2.4.2 Acquisition of Echo Waveforms

A measurement of the temperature distribution in a pulse-echo mode begins


with an excitation of an ultrasound pulse by a transducer. As the US pulse
propagates through the structured material , multiple echoe sare produced by
each echogenic feature encountered along its path. The waveforms of returned
echoes (labeled in Figure 2.2 by their time of flight TOF1, TOF2, TOF3, ...),
acquired by the US transducer, are the primary data used to measure the
segmental time of flight, and to estimate the temperature distribution in each

segment. The time of flight of the echo produced by a feature located at zi is


related to the temperature distribution along its propagation path as
where it was assumed that the trasducer /receiver is located at z = 0. The
temperature distribution in the i-th segment of the propagation path is inferred
from the difference in the time of flight, , between consecutive echoes
produced by echogenic features which bound the segment at the locations zi
and zi−1:

Where (zi−zi−1) is the segment’s length. The TOF of the first echo depends on
the temperature distribution across the 1st segment of the refractory, between
the cold surface and the first echogenic feature. The next return echo will
originate from the second echogenic feature. Then the difference in the TOF
between the second and the first echoes gives the information on the
temperature distribution in the second segment, and so on until the estimate of
the temperature distribution across the entire refractory is obtained. With that
distribution known, the last echo, reflected from the refractory-reactor zone
surface, can be used to determine the temperature of the refractory’s interior
hot surface. Note that we often use the first segment between the transducer
and the first echogenic feature at z1 as a delay line and reference the time of
flight of all subsequent echoes to the arrival time of the first echo, .

2.4.3 Signal Processing

The speed of sound is calculated as the distance traveled by an ultrasound pulse


divided by the time of propagation (or time of flight, TOF). Therefore, a method
for precise measurements of the time of flight is essential to accurate
measurements of temperature distribution. The simplest approach to the
measurements of the TOF and its changes is to use temporal location of a
single-point wave form feature , such as the first zero crossing or the peak value
of the waveform. Though standard, these timing techniques are sensitive to
measurement noises. Furthermore, when broad-band excitations are used, the
timing accuracy of single-feature methods deteriorates further due to waveform
distortions and broadening caused by stronger attenuation of higher-frequency
content of ultrasound pulses. More robust and accurate measurements of
may be achieved when the entire shape of the waveform is utilized in
timing. In this case, both amplitude and phase information are taken into
account ,which makes timing results less sensitive to measurement noises and
shape distortions. Mathematically, the cross-correlation between two signal f(t)
and g(t) is represented as:
where f∗ is the complex conjugate of f and τ is the lag time between two
signals. The temporal shift τ needed to obtain the best match between the
waveforms may be found by maximizing their cross-correlation , minimizing
l1 and l2 norms of their difference or by maximum likelihood . Figure 2.3 shows
∆TOF between two echoes is obtained by finding the best match of the entire
normalized shape of the two waveforms as based on both phase and amplitude
information, which makes timing results less sensitive to noises and shape
distortions.

Though cross-correlation and other shape-matching methods perform better


than single-point timing, the results may still be unacceptable when significant
distortion of ultrasound waveforms occurs, as is often the case when the pulse
propagates through attenuating and dissipative materials. It was suggested by
Le [89] that for waveform distorting samples , a higher precision can be
achieved if the envelopes of the wave forms are used in timing. The analytic

signal, ,of the waveform, , is the following complex function:

where =−1 and is the Hilbert transform of s(t):

The envelope of the waveform s(t) is then calculated at the amplitude of its
analytic signal:

We have recently shown that further improvements in timing accuracy can be


achieved by iteratively applying a nonlinear anisotropic diffusion filter to the
envelopes of the echo waveforms.
2.4.4 Temperature Dependence of the Ultrasound TOF

Theunknowntemperaturedistributionisestimatedfromthemeasurementsoftheti
meof fight of ultrasound echoes. The length of the propagation path and the
SOS both change with the temperature of the sample and thus influence the
echoes’ TOF. These factors typically contribute to the lengthening of the time of
flight as temperature increases. For relatively short waveguides and small
temperature ranges, the contribution of thermal expansion is negligible.

Note that it is possible to separate individual contributions of thermal expansion


and the changing SOS to the measured time of flight. However, as long as the
calibration curve (2.16) is obtained without such differentiation , there is no
practical need to distinguishthe degree each one of the two factors contribute
to the measured change in the TOF with temperature. The subsequent
discussion assumes that the correlation between the SOS vs. temperature, Eqn.
(2.16), was not corrected for the thermal expansion. This simplifies the method,
as it is becomes unnecessarily to adjust the length of the propagation path in
Eqn. (2.18) and (2.20) for the thermal expansion.

2.4.5 Estimation of Temperature Distribution

The measurements of the segmental time of flight encode the information


on the temperature distribution within i-th segment. As before, additional
assumptions are needed to estimate the segmental temperature distribution
from the measurement model (2.20). All parametrization options discussed in
the context of deconvoluting model Equation ?? may be used for this purpose,
and are discussed below.

Piecewise Constant Distribution


This distribution is obtained by assuming constant speed of sound within each
segment. Using this assumption in Equation (2.20), the constant SOS in the i-th
segment of the waveguide is obtained as

The corresponding constant temperature is obtained by inverting the


correlation (2.16). After repeating the process for all segments, the entire
temperature distribution along the waveguide is approximated as a piecewise
constant function. Infeasible temperature
discontinuitiesoccurringatthelocationsoftheechogenicfeaturesisanundesirablefe
ature of such approximation. Nevertheless, its accuracy is significantly better
than can be obtained by assuming a constant Ta along the entire path of
ultrasound propagation. Furthermore, by using a larger number of echogenic
features and the correspondingly finer segmentation of the waveguide, this
approximation can be further improved.

Piecewise Linear Distribution

The temperature continuity may be enforced by assuming that the temperature


changes linearly within each segment. For the i-th segment, this gives that

where mi and ni are unknowns. To further illustrate this parametrization,


consider a sample segmented by several echogenic features. Further assume
that at the transducer location, z = 0, the temperature of the sample’s proximal
end is measured independently (e.g., by using a surface thermocouple), and is
equal to T(0) = n1. By using the measured TOF of the first echo and the linear
temperature distribution in Eqn. (2.21), the following equation is obtained
from which the unknown slope of the distribution, m1, can be found. Similarly
for the second segment, the unknown slope m2 and intercept n2 are obtained
from the solution of the following two equations:

Where is the difference in the TOF of the second and first echoes and the
second equation enforces the continuity of the temperature at z= z1. The
process continues for all remaining segments until the piecewise linear
approximation of the temperature distribution over the entire sample is
obtained. For the i-th segment, the unknown slope mi and intercept ni are
obtained from the solution of the following two equations:

&

Where is the difference in the TOF of the i and i−1 echoes, and Equation
2.32 enforces the continuity of the temperature at z=zi−1.

Parametrization with thermal conductivity model

The temperature parametrization by the following one-dimensional heat


conduction models was used in :

where ρ, C, and k are refractory density, heat capacity, and thermal


conductivity, respectively. In both cases, the temperature of the distal end, Th,
was unknown and elevated, the temperature at the location of the transducer,
Tc, was assumed to be independently measured, and the temperature
distribution, T(z), was estimated by adjusting a single boundary condition in
order to match the predicted and the measured TOF of an echo produced by a
reflection of the excitation pulse from the distal end of the ultrasound
propagation path.

When a two- or three-dimension model is needed to provide an adequately


accurate description of the temperature distribution in the sample, additional
measurements will be required to reconstruct the temperature distribution. For
example, consider the case of a cylindrical waveguide with the transducer, used
to launch an excitation pulse in the axial direction, coupled to one of its ends.
Assuming the radial symmetry of the temperature distribution, constant density
ρ, heat conductivity k and capacity Cp, the temperature distribution inside the
sample must satisfy the following 2D heat transport model in cylindrical
coordinates:

where r is the radial position relative to the centerline of the sample. To


completely define the problem , three boundary conditions–at the proximal ,
distal and the cylindrical surfaces of the waveguide – are required. If the
waveguide is unstructured, only a single US echo will be produced by are
flection from a distal end of the sample and measurement of its TOF will allow
us to estimate only one of the three needed boundary conditions. The other
two boundary conditions must be obtained from independent measurements.
For example, if the temperatures of the distal and proximal ends of the
waveguide are independently measured , then the measured ∆tof can be used
to estimate the overall heat transfer coefficient h and define the remaining
boundary condition given as the heat flux through the cylindrical boundary of
the waveguide:

where Te is the ambient temperature of the environment. The time of flight of


multiple echoes received when the excitation pulse propagates through a
structured waveguide provides sufficient data to estimate all required boundary
conditions without the need for additional independent measurements. When
such independent measurements are available, they can still be in corporated in
to the US-MSTD method and may help improve the accuracy and the robustness
of the estimated temperature distribution.

2.4.6 Estimation of Heat Fluxes

The measurement of conductive heat fluxes through a solid is currently


obtained by attaching a flux sensor to the surface of the sample. It is therefore
capable of estimating only a localized heat flux in the immediate proximity of
the sensor. The approach proposed in this paper can be used to profile the
temperature distribution across the entire sample. It can, therefore, be used to
non-invasively estimate conductive heat fluxes through the sample, at a
considerable distance from the surface where the US transducer is attached.
Specifically, by differentiating (exactly or approximately) the estimated
temperature distribution T(z), the conductive heat flux, q, across the sample is
found as:

For the case of a piecewise constant temperature profile, one form of an


approximate differentiation gives:
The piecewise linear temperature distribution will result in a piecewise-constant

estimation of the heat flux distribution in the direction normal to the plane of
the ultrasound traducer. An even more detailed estimation is possible when the
heat flux is calculated based on the temperature profile that satisfies the heat
conduction model.

2.4.7 Segmental characterization of elastic properties

The proposed method essentially depends on the measurement of the


segmental speed of sound ( oritschange) in a solid sample. Factors , other than
temperature , that influence the speed of sound include the density of the
material and its elastic properties. For example, the velocity of longitudinal
waves (p-waves) in i-th segment of a “long” waveguide maintained at
isothermal conditions is equal to

Therefore, by measuring the SOS we can characterize the spatial changes in


Young’s modulus E of the sample and/or its density, ρ. Note that, as with the
estimation of the temperature distribution, the change in these properties may
be reconstructed under the assumption of their piece wise-constant or piece
wise-linear change in a different segment, or other suitable parameterizations.
Chapter 3
Anisotropic diffusion filter for robust timing of ultrasound
echoes
3.1 Time of Flight Measurements

Accurate measurements of return delays and transmission times are important


in many ultrasonic applications, including flaw localization , structure thickness
and strength measurements , range determination , measurements of fluid flow
rates , ultrasound imaging , and several biomedical applications . For example,
the accuracy of ultrasonic temperature measurements depends on our ability to
precisely measure the speed of sound (SOS) in materials, which we usually
determine by measuring the time of fight (TOF).

With focus on ultrasound measurements of temperature in this dissertation,


and the relationship between the time of flight and temperature, expressed as

3.1

for solid with known ultrasound propagation distance, accurate TOF


measurements are clearly essential for the accuracy of temperature
measurements.

3.1.1 Cross-correlation of Waveform

It has been pointed out in previous chapters that the simplest and standard
approach to the measurements of the TOF and its changes is to use temporal
location of a single-point wave form feature. However , such approach is
sensitive to measurement noises and small changes in experimental conditions.
Our initial approach to timing of US echoes was to use threshold crossing or by
matching the peak absolute values of the two round-trip echoes. When this
method was used to calculate the speed of sound in the aluminum sample, we

Figure 3.1: The difference in the TOF between echoes produced by an internal
interface of a two-layer cementitious sample and its distal end obtained by
threshold crossing (A), peak matching (B), cross-correlation(C), and envelope
cross-correlation of the waveforms (D).

Obtained =6481m/s,which is very close to the hand book value. Though


thresholding and peak matching approach are generally successful in low

acoustic attenuation samples, they become problematic when used to time


echoes propagating through dissipative samples. Figures 3.1(A) and 3.1(B)
illustrate their application in timing the difference between two echoes
produced by an internal interface of a cementitious waveguide and its distal
end. The results obtained by the two methods are different largely due to the
deterioration in timing accuracy of single-feature methods due to waveform
distortions and signal broadening caused by stronger attenuation of higher
frequency content of ultrasound signals.
When the entire shape of the wave for misutilized in timing by using cross-
correlation and shape matching methods, a more robust and accurate
measurement of may be achieved. In this case, both the amplitude and the
phase information are taken into account , which makes timing results less
sensitive to measurement noises. The temporal shift is obtained when the best
match is found between the waveforms by maximizing their cross-correlation ,
minimizing ℓ1 or ℓ2 norms of their difference, or by maximum likelihood .

Mathematically, the cross-correlation between two signal f(t) and g(t) is


represented as:

(3.2)

where f∗ is the complex conjugate of f and τ is the lag time between two
signals.

The shift time τ between two signals is determined by maximizing .


Thecross-correlationmethodformeasuringTOFwasinitiallytestedwithanaluminum
standard. The experiments were performed in pulse-echo mode using an
immersion transducerwith1MHz central frequency (PanametricsmodelV302).
Figure3.2 shows the collected ultrasound waveform for an aluminum sample
where multiple echoes represent round trips of the same excitation pulse. The
difference in the time of flight of two consecutive round- trip echoes may be
used to calculate the speed of sound in the sample. Figure 5.3 shows the
original waveform obtained with an aluminum sample and its shift by the delay
needed to maximize the cross-correlation between the first and the second
echoes. The shift needed for such an alignment is precisely the value of the time
of flight through the sample.
The cementitious samples and refractory materials are substantially more
dissipative compared to metals. As a result, multiple round trip echoes are
usually not observed. In this situation , the described method cannot be used to
measure of the TOF. Our approach in this case requires that we first determine
a zero time reference point, which we found to be different from the trigger
time acquired from the ultrasound pulser .

3.1.2 Zero Time Reference

Our tests show that the recorded trigger time, which corresponds to the timean
electrical pulse is sent to the piezoel ectric transducer,is not equal to the
ultrasound zerotime , which is the instant when the ultrasound pulse starts its
propagation through the sample. Our approach is to find the correction to the
trigger time in order to obtain a reference time relative to which the TOF will
then be determined.

We started by using the cross-correlation method to find the trigger time


correction for materials with low ultrasound attenuation. Figure 3.4 shows the
result obtained with the aluminum standard. The red dot on the blue waveform
is the starting point of the first echo. By shifting time by the determined TOF
value , the needed trigger correction for the particular transducer was found to
be equal to 91.2 nanoseconds. To establish if the obtained trigger correction
depends on the sample , the experiments were repeated with several other
materials, including bronze, stainless steel, steel and
Figure 3.2: The ultrasound excitation pulse applied to aluminum samples
produces multiple round-trip echoes.

Figure 3.3: The shift needed to maximize the cross-correlation between the first
and second echoes (blue trace) of the acquired response to the ultrasound
excitation equal to the ultrasound time of flight through the aluminum sample.
Green lineshows the original waveform shifted by the calculated TOF.
Figure 3.4: The time of the electrical trigger must be corrected to find the time
when the ultrasound waveform starts its propagation through the sample.

Table 3.1: Ultrasound propagation delay relative to trigger obtained with


different materials

plastic. Results for all tested materials are summarized in Table 3.1 and show
that the trigger correction time, equal to the delay between the electrical
excitation sent to the transducer and the start of the ultrasound pulse
propagation do not vary significantly with the material. This suggests that once
the propagation delay is determined for a given transducer, it can be used to
identify the zero reference when testing with different materials.

While we observed no to little dependence of the trigger correction on the


material of the sample, it is reasonable to expect that the size and design of the
transducer may have an effecton the ultrasound propagation delay which
characterizes the time of transduction of an electrical pulse into the ultrasound
pulse propagating through the sample. To test this hypothesis , we repeated the
described experiment with a different 1MHz transducer (Panametrics model
A114s with 1 MHz central frequency). For the same aluminum sample , the
signal delay observed with A114s transducer is only 25.6 nanoseconds , which is
substantially shorter than the transduction delay for V302 transducer.

The possible reason for two transducers with same central frequency to have
different transduction delays is the difference in their excitation bandwidth.
While V302 is a vedioscan type transducer with a broadband bandwidth needed
to optimize its near surface resolution, the A114S transducer is of accuscan type
designed to have a relatively narrow frequency bandwidth preferable when an
enhanced penetration is desired.

3.1.3 Cross-correlation of Envelopes of Ultrasound Waveforms

Though the cross-correlation method (illustrated in Figure 3.1(C)) performs


better with attenuating materials than single –point timing ,the results mays till
be unacceptable when significant distortion of ultrasound wave forms occurs ,
as is often the case with dissipative materials (e.g. , materials with large grain
sizes). For waveform distorting samples , a higher precision can be achieved if
the envelopes of the waveforms are used in timing, as illustrated in 3.1(D) for
the case when the cross-correlation between wave form envelopes is used.

The analytic signal , , of a given waveform , s(t) , is defined as the following


complex function:
(3.3)

Where is the Hilbert transform of s(t), which

represents the convolution of the original waveform with the function .

The envelope of s(t) is then calculated as an instantaneous amplitude of its


analytic signal:

(3.4)

where∗ is the complex conjugation operator.

Several methods have been proposed to time ultrasound signals( echoes and
transmission times) using the waveform envelopes. In , the transmission time of
ultrasound pulses was measured by using the peak value of the waveform
envelope as a reference point. Though this is still a single-point timing method ,
now based on the features of A(t), it improved the accuracy when compared to
timing based on single features of original waveforms (e.g., the first zero-
crossing by s(t)). In the case of dissipative (cementitious) samples, a noticeable
improvement was observed with this approach over the results obtained with
waveform cross-correlation. A simplified version of the envelope cross-
correlation method for TOF determination is summarized as Algorithm 3.1.

Algorithm 3.1
However, during long-term pilot-scale tests we found that performing envelope
cross correlation of echo waveforms is not sufficiently robust for accurate timing
when the signal-to-noise ratio of the acquired signals is low. This motivated the
development of the following method.

3.2 Anisotropic Diffusion Filter

A new method proposed here uses the results of Peronaand Malik. They
developed an anisotropic diffusion algorithm for image processing that smooths
an image without blurring the edges found within. To achieve this outcome, the
original image is iteratively subjected to a diffusion operator with spatially
dependent diffusion coefficient that becomes small on the edges. As a result,
the image features away from the edges are smoothed by diffusion, while the
diffusion is impeded across the edges to prevent their blurring.

The anisotropic diffusion operator used in the Perona-Malik filter has this
following form:

(3.5)
(3.6)

where the initial condition, I(r) , that initializes the iterations is the original
image , r=(x,y) defines spatial position within the image, and τ is the ordering
parameter that enumerates iteration steps of the discrete implementation of
Equations 5.5 and 5.6. The diffusion coefficient D(r,τ) changes with position and
is selected by the designer to decay at the edges. Two common choices that
satisfy this requirement include :

(3.7)

(3.8)

where κ is a selected constant. In both cases, the diffusion is small on the edges
where the gradient of the grayscale image intensity is high. A proper choice of
the diffusion function preserves and even enhances edges while ensuring
numerical stability .

The described filter has been used in several ultrasound imaging applications.
To the best of our knowledge , this approach has not been previously used in
timing the arrival of ultrasound waveforms by matching single-point features or
shapes of their filtered envelopes.

According to the proposed method , a 1D version of the diffusion filter from


Equations 3.5 and 3.6 is iteratively applied to the envelope of the acquired
ultrasound waveform:

(3.9)
(3.10)

The filtering process is initialized with the original envelope A(t), given by
Equation 3.4. The discrete approximation of Equation 3.9 is applied iteratively,
with each new iteration using the results of the previous step as the initial
condition. The specific approximation of Equation 3.9 used by us during the
experimental testing of the method is the same as the one found in :

(3.12)

where ∆t = 1 and u(t−∆t,τ), u(t,τ), and u(t+∆t,τ) are the temporally consecutive
values of the filtered envelope obtained at the preceding iteration. We elected
to use the diffusion coefficient, D, given by Equation 3.7 because it tends to
better preserve high-gradient data segments of the envelopes, which are
analogous to high contrast edges in images. After approximating the gradient
appearing in Equation 3.7, the following values of the diffusion coefficient are
used in Equation 3.11:

(3.13)

(3.14)
Where the parameter κ was set to20. The updated value of the filtered
envelope after each iteration is obtained by approximating the derivative on the
left-hand side of Equation 3.11. For example, by using a simple forward
difference approximation, the following update equation is obtained:

(3.15)

Where rhs is the right handside of Equation 3.11 and ∆ τ is the “ discretization ”
step selected to provide the desired rate of convergence. Algorithm 3.2
summarizes the process of applying the anisotropic diffusion to envelope of
ultrasound waveform.

The described filtering process uses a diffusion coefficient that takes small
values when the temporal changes in the envelope are rapid. It acts to sharpen
the envelope peaks while maintaining their temporal position over many
iterations. The filter blurs small and slow variations that contribute to
inconsistency of timing results commonly occurring when dissipative materials
are tested. As a result , the timing of ultrasound signals based on the filtered
envelopes is expected to be more robust to the presence of measurement
noises and distortions. In the following, we put this expectation to the
experimental test.

Algorithm 3.2
3.3 Experiments and Results

Two different dissipative ultrasound waveguides were used in experiments. The


first sample was obtained by hot pressing high-purity alumina into 12"-long rod
with 1" diameter (ceramic sample). Four 3/32" diameter holes were drilled
radially along its length at 1”, 2”, 4”, and 6" from the distal end (see insert A in
Fig. 1). The second sample was obtained by form casting Portland Type I/II
cement-water mixture into a mold (2 in. I.D. PVC tubing). After casting a 1”-thick
layer, we allowed time for partial curing to occur; this followed by casting the
second 1" layer of identical composition and complete curing of this
cementitious sample (shown in insert B). Partial curing introduced enough
variation in material properties to create partial US reflections of the excitation
pulse from the internal interface between the two layers. The ultrasound
measurements were obtained using Panametrics pulser /receiver (model
5072PR) and contact transducer (model V609) which created an excitation pulse
with 5MHz central frequency. The echo waveforms were acquired using
Tektronix oscilloscope (model MSO 2024) interfaced to a computer. Custom
Matlab code was written to control the acquisition, visualize the waveforms,
and filter, analyze and interpret the data. Fig. 3.5I shows typical waveforms
collected with the alumina waveguide. Partial ultrasound reflections from the
drilled holes are marked as PR1, ..., PR4 echoes; the reflection from the distal
end of the waveguide is labeled as DE. The envelopes of these echo wave forms
are shown in Fig. 3.5II.Fig. 1III shows the echoes from the internal interface and
the distal end of the cementitious sample, along with the corresponding
envelopes. The importance of accurately measured time of flight of ultrasound
pulses is highlighted by the influence the timing errors have on ultrasound
measurements of temperature. Experimental characterization of c = f(T) for our
alumina waveguide indicated that at room temperature of 20◦C , a TOF
measurement error that overestimates the actual values by 0.1% implies that
the temperature of the wave guide is 61◦C–a temperature measurement error
of 41◦C. A measurement error corresponding to 1% overestimation in the TOF at
room temperature will result in the overestimation of the waveguide
temperature by over 300◦C.
When the ultrasound propagates through dissipative materials , noticeable
variations in the waveforms taken under apparently identical experimental
conditions are often present. Fig. 5.6II illustrates the variability in the echo
waveforms (blue traces) produced by an echogenic feature PR1 located in the
middle of the ceramic waveguide. Both waveforms were acquired at identical
experimental conditions but different times. The envelopes of the two
waveform (black lines) are also affected by this variability. For instance, note the
change in the position of the maximum envelope values indicated by green
triangles. If the maximum values are used to time the arrival of echoes, the
observed variability would result in TOF measurement errors and
inconsistencies. Specifically, a shift in the maximum value between the two
peaks in Fig. 5.6II introduces the variation of 0.4% in the TOF measurements,
which corresponds to a very significant (on the order of 150◦C) error in the
estimated waveguide temperature.

Panel IA inFig. 3.6 shows the original echowave form produced by the third
echogenic feature of the ceramic waveguide (PR3) and its envelope. As in the
case of echoes produced by PR1, the envelope of the original waveform is
characterized by two closely separated peaks. Experiments showed that echo
waveforms produced by PR3 and acquired at different times but identical
conditions can also have the maximum envelope value at either one of these
peaks. To address the observed inconsistencies and the corresponding timing
errors, the proposed anisotropic filter was applied to the envelopes of the echo
waveforms. The result for the ceramic waveguide is illustrated in Panels IB
through IE that show the filtered PR3 envelope as the number of iterations in
applying anisotropic diffusion is increasedfrom100to1500. With more
iterations , small variations are blurred, the filtered envelope retains a single
peak, and its position is stabilized. The position of this peak can now be used to
consistently time the arrival of ultrasound echoes. Once the filtered envelope is
obtained, other single-point and shape-matching timing techniques may also be
used.

Panel II shows that small variations in the envelopes of PR1 echoes acquired at
identical conditions are filtered out after 3000 iterations and the single peak
value is stabilized at the same location within both envelopes. If the maximum
value of the filtered envelope is used to time the echoes, the identical
estimation of the time of flight will be produced despite the variations in the
original waveforms and their unfiltered envelopes. The influence of different
timing methods on the variability of the estimated TOF was investigated next.
Several timing methods were considered: (a) Timing based on the maximum of
the acquired echo waveform; (b) Timing by threshold value, which we selected
to be 1/3 of the maximum of the waveform value; (c) Cross-correlation between
the two waveforms; (d) Cross-correlation between their envelopes; (e)
Maximum value of the filtered envelope; and (f) Cross-correlation between
envelopes filtered using the anisotropic diffusion. The comparison was first
performed for ceramic waveguide based on the two PR1 echo waveforms
shown in Fig. 5.6II. For each selected method, the difference in timing results
when one or another waveform was used is given in Table 5.2. The same table
also lists the corresponding errors in the estimates of the waveguide
temperature. After applying 3000 iterations of the developed anisotropic
diffusion filter, the peaks of filtered envelopes were located in identical
positions for both waveforms, as illustrated in Fig. 5.6II. The cross-correlation
between the two filtered envelopes shows that a zero shift is needed for the
best match, indicating timing consistency when either one of the wave for
msisused. For all other methods , the difference intiming based on the two
waveforms was on the order of 25 nanoseconds. If we assume that this
difference is an increase in the TOF over the actual value acquired when the
waveguide is maintained at 20◦C, the timing errors would correspond to the
overestimation of the waveguide temperature by over 40◦C.
Figure3.6: (I)The original echo wave form , shown in blue , was produced by PR3
echogenic feature of the ceramic waveguide; its envelope is shown in A. The
application of the anisotropic diffusion filter for 100, 500, 1000, and 1500
iterations produced filtered envelopes respectively shown as B, C, D, and E.
Maximum values of filtered envelopes are marked with red triangles. (II) Two
echoes, shown as blue traces, were produced by the PR1 feature. They were
acquired at the same conditions but different times. Their envelopes show
variability in the position of the maximum value indicated by green triangles.
After applying 3000 iterations of anisotropic diffusion algorithm, the filtered
envelopes of both wave forms are shown in the middle of the panel. Red arrows
show that the peak values coincide after filtering.

Table 3.2: Timing errors in ceramic waveguide


Similar comparison of timing results produced by different methods was
repeated for the cementitious sample. The results, summarized in Table 5.3,
show a larger difference in the estimated TOF based on two different
waveforms (not shown), indicating stronger dissipation and attenuation of
ultrasound signals in a cementitious sample. The difference in the estimated
temperature , however, is smaller be cause of substantially slower speed of
sound in this sample. Specifically, a1% difference in TOF at a reference
temperature of 20◦C produced 26◦C difference in the estimated temperature of
the cementitious waveguide. As before , after anisotropic diffusion filtering of
the envelopes the consistent timing results were obtained based on the location
of the peak values (they coincide) and the temporal shift needed to maximize
the cross-correlation between the two filtered envelopes (zero shift needed).

The developed anisotropic diffusion filter was applied to the waveform data
collected during transient changes in the temperature distribution of the
alumina waveguide partially inserted into a high-temperature furnace (these
experiments were described in the preceding Chapter). Figure 3.7 compares the
results of these experiments obtained without the anisotropic filter (top graph)
and and after its application to the waveform data.

Table 5.3: Timing errors in cementitious waveguide


Furnace setpoint temperature is shown as a grey dashed line in Figure 5.7. The
thermocouple measurements at steady state are also included. After applying
the developed filter to the waveform envelope data, the obtained ultrasound
temperature measurements are consistent with the thermocouple
measurements, the setpoint trajectory, and unexpected excursions in the
estimated temperature distribution are eliminated.
Figure3.7: Real-time measurements of the temperature distribution in the sample during
transient changes in the furnace set point were obtained before (top plot) and after
applying the anisotropic diffusion filter. After the envelopes of the waveforms were filtered,
the excursion in the estimated temperature away from the setpoint trajectory were
eliminated.

Chapter 4
Pilot Scale Testing
4.1 Pilot Scale Oxy-fuel Combustor
The pilot scale experiments were carried out on a down-fired oxy-fuel
combustor (OFC) schematically shown in Figure 4.1. This 100 kW pilot unit was
designed to allow for a systematic control of inlet gas flow rates and to maintain
refractory temperature high enough to simulate the self-sustaining combustion
conditions of full-scale units in terms of the temperatures, coal particle
concentrations, and mixing.

The OFC has three zones which are , from top to bottom: ignition
(0.61ID×0.91OD× 1.22mheight), radiation(0.27×0.61×2.60m), and
convection(0.15×0.15×3.66m) zone. The combustion products exit the
convection zone and pass through eight heat exchangers which cool the flue gas
prior to discharge. The ignition zone is normally surrounded by 3 ×8×840W
flanged ceramic-plate electrical heaters used to control the wall temperatures
of the furnace at the start-up. However, these heaters fell off from the interior
surface of the refractory wall during a previous test and our experiments
proceeded without using them. The burner for this unit is not equipped with an
ignition nozzle. As a result, natural gas must first be used initially to heat up the
furnace to a temperature high enough to ignite coal particles introduced in to
the ignition , at which point a self sustained combustion can be maintained.
Natural gas is also used to keep furnace temperature high at night. The
temperature during coal combustion is higher than achieved with natural gas ,
as visually apparent from Figure 4.2, which shows a quartz window into the
ignition zone. Nine pairs of ports provide access to the reaction zone inside the
OFC. All ports are positioned along the vertical section of the unit and are
marked as marked as P1–P9 in Figure 4.1. They can be used for sampling,
instrumentation insertion, and observations. Our refractory ultrasound
waveguide was mounted in the ignition zone inside port P3.
Figure 4.1: Pilot scale down-fired oxy-fuel combustor.
Figure4.2: Flame imaged through a quartz window during NG (left) and
coalcombustion.
Figure 4.3: Schematic depiction of US-MSTD system designed and installed
during the OFC pilot testing. Alumina wave guide was engineered with partial
ultrasound reflectors located along the 1×12- inch alumina rod. Their locations ,
marked as z1, z2 , z3 and z4, were 6 , 4 , 2 and 1 inches away from the hot distal
end (DE) of the rod. The data acquisition and interpretation systems provided
real time measurements and display of the temperature distribution across the
wave guide and the comparison with the independent thermocouple
measurements.

4.2 Components of the US-MSTD System: Design and Implementation

The overall US-MSTD system, that was designed to provide continuous real time
temperature measurements , consists of an ultrasound wave guide engineered
to provide multiple partial ultrasound reflection from known locations along its
propagation length, ultrasound transducer/receiver, and the data acquisition,
logging, interpretation, and the real-time display system. The schematic view of
the US-MSTD system in Figure 4.3. An alumina rod was selected as the
waveguide and the partial reflections were created by 4 small holes drilled along
its length at 1", 2", 4" and 6" from the hot distal (relative to transducer location)
end of the rod .

The overall design of the waveguide insertion system is shown in Figure 4.4. It
was
Figure 4.4: Port-mounted waveguide retention system.

designed to address several issues , including integrity of process containment ,


ultrasound coupling, and active cooling of the rod to insure that the
temperature of its cold end is below the Curie temperature of the ultrasound
(US) transducer. The photograph in Figure 4.5(B) shows the implementation of
the insertion part of the system prior to its installation.

The alumina refractory rod with 5 segments was continuously used as the
ultrasound waveguide during pilot scale tests. Its segment extending from
transducer-waveguide interface to z1 served as a delay line and its temperature
distribution was not estimated. The combined thickness of the refractory and
the steel shell of the OFC is 6-1/2 inches. The length of the inserted alumina
waveguide was selected to place the distal end of the waveguide flush with the
OFC refractory (Figure 4.5(G)). The diameter of Port 3 is 3 inches. The inserted
waveguide was wrapped with fiber glass insulation to fill the gap between
ultrasound waveguide and OFC refractory. Omega Super type K
thermocouples were bent at 90◦ and inserted into the holes drilled into the
waveguide to provide ultrasound echoes. These thermocouples were used to
acquire independent point wise temperature measurements along the length of
the waveguide. In the described arrangement, the thermocouple measurements
at the location of the distal end of the waveguide are not available. In order to
obtain independent thermocouple measurements on the surface of the
refractory , Omega Super type B thermocouple was
inserted in Port 2 and its tip placed flush with the inside surface of the
refractory. Note, however, that the temperature at the internal refractory
surface at different elevations of ports P2 and P3 will generally be different.
Cooling water was run through the waveguide fixture to ensure that the
temperature of the rod’s cold side is within operating limits of the ultrasound
transducer. Figure4.5 (E) shows the flange fixture with the attached cooling
water , thermocouple leads, and the BNC connection to the ultrasound
instrumentation and the data acquisition system located on the ground level
within 20’ of P3 port. Figure 4.5(F) shows that we used a reflective shield to
minimize the effect of radiative heat transport on the ultrasound transducer.
Figure 4.5: (A) Waveguide was mounted in P3 port. (B) Engineered alumina waveguide is
shown mounted in the flange fixture. (C) Thermal insulation was wrapped around the
waveguide. (D) The fixture is being inserted into P3 port. (E) The mounted waveguide is
shownwiththeattachedultrasoundtransducerconnectedtotheBNCcableleadingtothe US data
acquisition system. Thermocouples are shown connected to DAQ. (F) Radiation shield is
now installed. (G) Distal end of the transducer is shown to be mounted flush with the
refractory surface. The picture was taken prior to the experiments with a camera inserted
inside the ignition zone. (H) The system during the experiments. (I) Computer screen shows
the ultrasound waveforms acquired in real time and the corresponding temperatures
obtained with ultrasound measurements. (J)Components of the ultrasound data acquisition
system. (K) The waveguide removed from the port at the completion of the experiments. (L)
The waveguide after removing thermal insulation. Note thermocouple wires inserted into
holes drilled to provide US echoes.
The transducer was mounted using a retaining plate with threaded ring, which
provided a secure and flexible way for its installation. The transducer was
coupled to the waveguide using silicone grease. Figure 4.5(J) shows components
of the ultrasound system and the data acquisition computer running custom
Matlab software written to control data acquisition, and interpret and visualize
ultrasound measurements, Figure 4.5(I).

In the described configuration, the sensitive electronic components are kept


away from harsh gasification environments and it is only required that the US
transducer be acoustically coupled to the cold side of the refractory ,
representing minimal modifications to the combustor.

The data acquisition system was kept the same as used during high-temperature
laboratory testing. A custom Matlab software was developed for on line data
acquisition, storage, interpretation and temperature profiling across the
refectory. The real time visualization of the temperature distribution based on
the measurements of the segmental time of flight used the combination of
envelope cross-correlation method and anisotropic diffusion filtering, discussed
in the previous Chapter. Using the segmental time of flight measurements
obtained from the analysis of the acquired ultrasound echo waveforms, the
real-time temperature distribution was obtained using the speed of sound vs.
temperature relationship obtained during laboratory experiments.

In order to process the ultrasound measurements at a high sampling rate (every


5 seconds), a piecewise constant parametrization of the temperatures
distribution in the four segments of the waveguide was used. The procedure for
online estimation of the temperature profile in the alumina waveguide is given
as Algorithm 4.1. The waveforms acquired during the pilot test have been
stored and later reinterpreted using piecewise linear parametrization of the
temperature distribution.
4.3 Experimental Conditions

Algorithm 4.1 Acquisition of Ultrasound Waveforms and estimation of the


Temperature distribution along the Waveguide

lb/hr; Secondary O2 was fed at 16.5 lb/hr. The internal pressure was elevated
by∼0.1" of water. During natural gas combustion ,a flow rate between 6 and
6.5 lb/hr was maintained. The US-MSTD system was used to monitor
temperature distribution across the refectory during several distinct regimes of
OFC operation, which included:

• Steady state combustion of natural gas and coal;

• Fuel transition from natural gas to coal; and

• Coal combustion at the different fuel flow rates.


4.4 Experiments and Results

4.4.1 Natural Gas Preheating

Figure4.6 shows the measured temperature change along the waveguide based
on piecewise constant assumption during initial preheating of the OFC by
natural gas. About half an hour after the heating started, there was an
unaccounted temperature drop in zone 4, which is the closest to the flame. The
temperature in Zone 4 eventually increased above the temperature of Zone 3
and remained at higher at higher temperature for the duration of 5-day-long
test. The electrical heaters, normally used at the startup of the unit, were not
available during this experimental campaign. It is possible that we are seeing a
temperature change due to a combination of changing flow rates of natural gas
and air, controlled manually and without the benefit of computer controlled
refractory temperature. The acquisition of the thermocouple measurements did
not start until 17:00. After 18:00, the drying of the ultrasound couplant caused
the deterioration in the strength of the ultrasound echoes. The smallest SNR is
for the last echo , which resulted in a higher variance in the estimated
temperature of the hottest segment of the refractory, seen in Figure 4.6. The
acquisition of the ultrasound signal was stopped at 18:30 and was not restarted
until the next day.

4.4.2 Steady Natural Gas Combustion

Temperature measurements using piecewise constant parametrization based on


the data acquired after the unit has reached steady operating conditions during
natural gas combustion are shown in Figure 4.7. Figure 4.8 gives the estimation
of the temperature distribution obtained using piecewise linear
parametrization. The comparison of pointwise thermocouple measurements at
the locations of the echogenic features z2 , z3, z4 with the ultrasound
measurements in the corresponding spatial locations extracted from the
estimated temperature distribution obtained under piecewise constant and
piecewise linear parameterizations are shown in Figure 4.9. Because piecewise
constant distribution is discontinuous at the location of the echogenic features,
the result in Figure 4.9 for this parametrization is obtained by averaging
temperature in the adjacent segments. For over 3 hours of continuous data, this
figure shows that the ultrasound and the thermocouple measurements are
consistent and for some locations (e.g., z4) are very close to each other.
Generally, the ultrasound measurements based on piecewise linear
parametrization are closer to the thermocouple measurements. In general, the
ultrasound measurements appear to be more sensitive to temperature changes.
The variance in the ultrasound results based on piecewise linear
parametrization is higher, reflecting higher sensitivity of the results to the
estimated slope of the linear profile of the segmental temperature.

Figure 4.7: Ultrasound measurements of segmental temperature during stable


natural gas combustion under piecewise constant parametrization.
Thermocouple measurements are shown for comparison.
Figure 4.8: Temperature distribution across the refractory during stable natural
gas combustion obtained based on piecewise linear parametrization.

Several factors contribute to the difference in ultrasound and thermocouple


measurements. While thermocouples provide pointwise measurements, the
ultrasound measurements are sensitive to the temperature distribution , which
we are estimating using different parameterizations. The second important
difference is the sensitivity of the thermocouple measurements to , primarily ,
surface temperatures , while the US measurements depend on the temperature
distribution inside the waveguide. Finally, the thermocouples were not potted
inside the holes , thus creating an addition thermal resistance that contributes
to the observed differences. In Figures 4.7and 4.9 , note straight line segments
in the ultrasound data starting at 14:45. During this time, the ultrasound
couplant was changed and no data were acquired.

Figure 4.9: Temperature evolution at the distal end of the waveguide and
feature locations zi during stable natural gas combustion. Ultrasound results
were obtained using piecewise constant and piecewise linear parametrization of
the segmental temperature. The thermocouple measurements at the location of
echogenic features are provided for comparison and show consistency with the
ultrasound measurements.
Figure 4.10: Ultrasound measurements of segmental temperature during
transition from natural gas to coal combustion under piecewise constant
parametrization . Thermocouple measurements are shown for comparison.
Ultrasound and thermocouple measurements are in agreement, but the
ultrasound response for temperature changes is faster.

4.4.3 Transition from Natural Gas to Coal

Figures 4.10 and 4.11 show the temperature change when the combustor was
switched from natural gas to coal under piecewise constant and piecewise
linear parameterizations, respectively. Thermocouple measurements are
included for comparison. Initially , just before 8AM the flow of natural gas was
reduced, leading to the decrease in the temperature. Note that, because the
ultrasound measurements are sensitive to the entire temperature distribution,
they responded to this change before TC1 thermocouple did. The feeding of
coal started just after 8AM . Again , the ultrasound measurements promptly
indicated that the change in fuel has occurred by showing the trend of rising
temperatures before this change was registered with thermocouples.
The temperature distribution change along the waveguide indicates the heat
conduction process from refractory hot face to vessel’s shell. The comparisons
of temperatures measured using two different parametrization in US-MSTD
method and the thermocouple measurements at hot distal end of the
waveguide and all echogenic features are shown in Figure 4.12. Compared to
thermocouples, the US measurements give both faster and more pronounced
indication that the operating conditions have changed.

Figure4.11: Temperature distribution across the refractory during transition


from natural gas to coal combustion obtained based on piecewise linear
parametrization.

4.4.4 Response to Decreased Feed Rate of Coal

Figure 4.13 shows an initial increase in temperature after switching from natural
gas to coal on one of the test days. This transition occurred before 8AM and was
not captured due to miscommunication with operating personnel. After rising,
the temperature stabilized, at which point, at approximately 9:20 AM, the flow
rate of coal was reduced. The measured temperature decreased in response
and stabilized in the lower range. The corresponding temperature distribution
obtained using piecewise linear parametrization is shown in Figure 4.14.
Compared to thermocouples, the US measurements give, both, faster and a
more pronounced indication that the operating conditions have changed. Note
that the temperature in segments 2 and 1 was only marginally affected.

The comparisons of temperatures measured using two different


parametrization in USMSTD method and the thermocouple measurements at
hot distal end of the waveguide and all echogenic features are shown in Figure
4.15, and indicate an excellent agreement.

4.4.5 Stable Coal Combustion

The last captured process during this pilot testing campaign was the stable coal
combustion. The temperature distribution estimated using US-MSTD method is
shown in Figures 4.16 and 4.17 under piecewise constant and piecewise linear
parameterizations, respectively. The comparisons of temperatures of obtained
using both parameterizations and the thermocouple measurements at hot distal
end and all echogenic features are shown in Figure 4.18. The refractory
temperatures during coal combustion are higher than during the natural gas
combustion, especially in zone 3 and 4, closest to the flame. Significant
reduction in the strength of the ultrasound echoes was observed as the pilot
testing progressed, especially for the echo produced by the feature closest to
the combustion zone (1 inch away from the distal end). Several intervals when
ultrasound measurements flat lined correspond to times when the signal had
deteriorated enough to motivate the application of fresh ultrasound couplant at
the transducer - waveguide interface. Subsequent investigation of the the
alumina waveguide revealed that a more likely reason for the deterioration of
the ultrasound signals is the change in the material properties , mainly the
growth of the grain boundaries, as discussed later in this chapter.
Figure 4.12: Temperature evolution at the distal end of the waveguide and
feature locations zi during transition from natural gas to coal combustion.
Ultrasound results were obtained using piecewise constant and piecewise linear
parametrization of the segmental temperature. The thermocouple
measurements at the location of echogenic features are provided for
comparison and show consistency with the ultrasound measurements .
Figure 4.13: Ultrasound measurements of segmental temperature during the
reduction in the feed rate of coal under piecewise constant parametrization.
Thermocouple measurements are shown for comparison and indicate the same
trend in temperature changes.

Figure 4.14: Temperature distribution across the refractory during the reduction
in the feed rate of coal obtained based on piecewise linear parametrization.
Figure 4.15: Temperature evolution at the distal end of the wave guide and
feature locations zi during the reduction in the feed rate of coal. Ultrasound
results were obtained using piecewise constant and piecewise linear
parametrization of the segmental temperature. The thermocouple
measurements at the location of echogenic features are provided for
comparison and show consistency with the ultrasound measurements.
Figure 4.16: Ultrasound measurements of segmental temperature during stable
coal combustion under piecewise constant parametrization. Thermocouple
measurements are shown for comparison and indicate the same trend in
temperature changes.

Figure4.17: Temperature distribution across the refractory during stable coal


combustion obtained based on piecewise linear parametrization.
4.4.6 Surface Temperature of the Refractory

The distal end of the waveguide was aligned with the hot refractory surface.
Therefore, the temperature of the distal end measured by US-MSTD method
gives the direct characterization of the temperature on the surface of the
refractory. Figure 6.19 shows the ultrasound measurements of the distal end
temperature under two parametrizations of the temperature distribution
obtained during the process transitions and different operating conditions
described above. For comparison, this figure also shows the temperature
measurements obtained with the thermocouple inserted into a different ports
(port P2) and aligned with the surface of the refractory. For all operating
conditions , the ultrasound measurements are consistent with the independent
thermocouple measurements and capture similar temperature trends. This
demonstrates on the pilot scale the capability of the developed US-MSTD
method to measure surface temperature inside containments of extreme
environments.

4.5 Discussion

The US-MSTD system was successfully tested on the pilot scale OFC oxy-fuel
combustor during ∼120 hours of continuous operation. Real time temperature
distributions along the waveguide were captured during all relevant process
changes. The comparison of temperature results based on US-MSTD method
and thermocouple measurements exhibit excellent consistency. Several issues
were revealed during the test.

4.5.1 Couplant

The continuous real-time temperature monitoring using US-MSTD method


requires permanent coupling which is essential for providing the acoustic bond
between the transducer and the waveguide. Couplants come in different forms,
such as gels, oils, grease, pastes and others. After laboratory experimentations,
for pilot scale studies we selected silicon grease because of its close match of
acoustic impedance to that of alumina, high viscosity, relatively slow
drying/evaporation rate compared to other gels, and low sensitivity of its
properties to environmental changes. Despite these attractive attributes, we
found that it is still necessary to change our couplant every 2–3 hours during
pilot-scale experiments. Flexible bonding agents, such as silicon rubber
compounds, or rigid bonding with adhesives or epoxies, such as cyanoacrylates,
may provide good ultrasound transmission (comparable with what we achieved
with “wet” couplants) and a more permanent and a strong bond to the
waveguide. Clearly, with all alternatives, it is important to insure that an
adhesive layer does not contain voids nor air bubbles. Further studies are
needed to characterize the ultrasound performance of the alternative coupling
solutions and their long term stability in revelent applications.
Figure 4.18: Temperature evolution at the distal end of the waveguide and
feature locations zi during stable coal combustion. Ultrasound results were
obtained using piecewise constant and piecewise linear parametrization of the
segmental temperature. Thethermocouple measurements at the location of
echogenic features are provided for comparison and show consistency with the
Ultrasound measurement

Figure 4.19: Temperature evolution at the distal end of the waveguide under
different operating conditions. The thermocouple measurements were obtained
at the different elevation inside the flame zone, in Port P2.

4.5.2 Alumina Refractory Waveguide

The alumina waveguide had no visual damage after 5 days nonstop use in a high
temperature environment of OFC (Figure 4.5(K) and (L)). Only a limited ash
deposit was found on its distal end and no surface corrosion was observed. At
the same time,the thermocouple at the position z4 (echogenic feature closest to
the flame zone) has lost its functionality after the test. Its outer sheathing was
burned, became brittle, and the inner wires were exposed with any subsequent
handling.

Dispute the normal appearance of the alumina waveguide, a significant and


irreversible deterioration in the strength of ultrasound echoes was revealed.
This indicates that alumina experienced property changes caused by a
prolonged exposure to high temperatures. We attribute this change to micro-
structure changes, mainly the growth in alumina grains. Micro-structural
changes due to an increase in size of grains (crystallites) in a material at high
temperature is commonly seen in metals , minerals , and ceramics. This happens
when recovery and recrystallization are complete and further reduction in the
internal energy can only be achieved by reducing the total area of grain
boundary.

Grain boundaries have associated macroscopic and microscopic degrees of


freedom, which play an essential role in controlling the dynamic
growth/depletion of grains under specific thermal conditions. Excess free
energy of grain boundaries is a driving force towards the reduction in total area
of grain boundaries, which drives the growth of grain sizes and the reduction of
the number of grains per unit volume. Mullins investigated the kinetics of the
grain growth. The rate of change of the mean grain size, dr/dt, must be related
to the migration rate of boundaries in the system. The mechanism for ideal
grain growth and a quantitative relationship between the velocity of a grain
boundary, υ, and the size of the grains may be described as

(4.1)

where M is grain boundary mobility, ζ is the grain boundary energy, and κ is the
sum of the two principal surface curvatures of the grain. The integration of this
equation gives:

(4.2)
Figure 4.20: Alumina grain size distribution in the waveguide material as a function of
thermal treatment is images with over 13000x magnification. (A) shows grains prior to heat
treatment ; the sample imaged in (B) was subjected to short term treatment ; the effect of
the long term exposure during pilot scale testing on the grain sizes is shown in (C).

where r is the final grain size and rt=0 is the initial grain size. Grain growth is
irreversible and strongly temperature dependent. High temperature
environment accelerates the growth. To mitigate this process in the refractory
martials that must operate at high temperature for a long time, different
dopants, such as MgO, CaO, and SiO2, are often added to stabilize the
boundaries and inhibit the growth of grains.

The growth of grain sizes caused exposure to high temperature was


characterized by conventional and electron microscopy. For example, SEM
images shown in Figures 4.20 demonstrate the effect of the exposure to high
temperatures on the grains of the alumina wave guide used by us. Figure4.20(A)
shows that green alumina ( sample without thermal treatment) has a relative
uniform particle size distribution , between 1 and 5µm . After short term
exposure to high temperatures of over 800◦C for less than 100 hours, the
symptoms of the grain growth are observed in Figure 4.20(B), which shows a
higher percentage of grains of larger sizes. Figure 4.20(C) reveals a substantial
increase in grain sizes of the alumina sample heat treatment up to 1300◦C for
over a thousand of hours to the point that a single grain has grown to the point
of occupying almost the entire field of view.
When the mean grain diameter of our alumina ceramic waveguide increases,
the inhomogeneity causes stronger scattering and absorption of ultrasonic
energy and increases the attenuation of the ultrasound echoes significantly. This
observation indicates that we can monitor certain mechanical properties of the
refractory material and detect their changes prior to visible degradation and
damage to the refractory.
Chapter 5
Conclusion
This Report summarizes the concept, the successful development, and the
experimental validation of the method for ultrasound measurements of
segmental temperature distribution in solids. This project solved a myriad of
technical problems to make the proposed concept a reality and to successfully
test the performance of this new method on a pilot scale. As a result of this
project, this noninvasive method becomes a new alternative for solving a
challenging problem of long-term temperature measurements in extreme
environments , such as though encountered inside gasifiers , combustors , and
other energy conversion processes.
Bibliography -
[1] In-Situ Acoustic Measurements Of Temperature Profile In Extreme
Environments

[2] Google

[3] C. Gault, M. Huger, J.M. Auvray, J. Soro, and E. Yeugo Fogaing. Contribution
of high temperature ultrasonic measurements to investigations of thermo
mechanical behavior of refractories . In 10 the International Conference and
Exhibition of the European Ceramic Society, 2007.

[4] A Wolfenden. Measurement and analysis of elastic and anelastic properties


of alumina and silicon carbide. Journal of materials science.

[5] S Chakraborty , S Sakar ,S Gupt a, and A Ray. Damage monitoring of


refractory wall in a generic entrained-bed slagging gasification system.
Proceedings of the Institution of Mechanical Engineers, Part A: Journal of Power
and Energy

[6] Ihsan Sarigul. Model-based estimation of adiabatic flame temperature during


coal gasification. Master’s thesis.

[7] G Pickrell, Y Zhang, and A Wang. Development of a temperature measure-


ment system for use in coalgasifiers . In 19th Annual International Pittsburgh
Coal Conference, 2002.

[8] Yunlu Jia and Mikhail Skliar. Ultrasound measurements of temperature


distribution and heat fluxes in solids. submitted, 2015.
[9] Brian Michael Lempriere. Ultrasound and elastic waves: frequently asked
questions. Academic Press, 2003.

[10] A Badidi Bouda, S Lebaili, and A Benchaala. Grain size influence on


ultrasonic velocities and attenuation. NDT & E International.

[11] G. Kychakoff, A.F. Hollingshead, and S.P. Boyd. Use of acoustic temperature
measurements in the cement manufacturing pyroprocess.

[12] Krishnan Balasubramaniam, Vimal V Shah, R Daniel Costley, Gary


Boudreaux, and Jagdish P Singh. High temperature ultrasonic sensor for the
simultaneous measurement of viscosity and temperature of melts.

[13] R.K. Ramalingam and H. Neumann. Fiber bragg grating-based temperature


distribution evaluation of multilayer insulations between 300 k-77 k.

[14] M.Parrilla, J.J.Anaya , and C.Fritsch. Digital signal processing techniques for
high accuracy ultrasonic range measurements.

[15] M.J. Black, G. Sapiro, D.H. Marimont, and D. Heeger. Robust anisotropic
diffusion.

You might also like

pFad - Phonifier reborn

Pfad - The Proxy pFad of © 2024 Garber Painting. All rights reserved.

Note: This service is not intended for secure transactions such as banking, social media, email, or purchasing. Use at your own risk. We assume no liability whatsoever for broken pages.


Alternative Proxies:

Alternative Proxy

pFad Proxy

pFad v3 Proxy

pFad v4 Proxy