Measurement of Temperature Using Sound-1
Measurement of Temperature Using Sound-1
ON
Measurement of Temperature
Using Sound
[FOR THE PARTIAL FULFILMENT OF 6TH SEMESTER UG (PHYSICS) 2019-20]
Submitted by
Piyush
For the partial fulfilment of UG 3rd year PHYSICS, encompasses the results of
studies carried out by him/her.
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.
Signature of student
PREFACE
5) Conclusion
6) Bibliography
Acronyms
One-dimensional (1D)
Two-Dimensional (2D)
Three-Dimensional (3D)
Thermocouple (TC)
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)..
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
Delisle [°De]=(100[°C])× 3
⁄2 [°C]=100[°De]× 2
⁄3
1.3 Gasifier -
Indirect Measurement
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.
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
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
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 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:
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.
2.2.4 Scattering
Note that
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.
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.
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 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.
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, .
The envelope of the waveform s(t) is then calculated at the amplitude of its
analytic signal:
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.
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.
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.
3.1
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).
(3.2)
where f∗ is the complex conjugate of f and τ is the lag time between two
signals.
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.
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.
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.
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.
(3.4)
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.
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.
(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
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.
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.
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.
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.
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).
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.
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:
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.
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.
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.
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 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.
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
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.
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.
(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.
[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.
[11] G. Kychakoff, A.F. Hollingshead, and S.P. Boyd. Use of acoustic temperature
measurements in the cement manufacturing pyroprocess.
[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.