CASE_Jul09b - DSpace at ICSE

advertisement
Revised Quarterly Progress Report
Clean and Secure Energy from Coal, Oil Shale & Oil Sands
University of Utah
DE-NT0005015
August 13, 2009
Philip J. Smith (PI)
Project Period
April 1, 2009 to June 30, 2009
EXECUTIVE SUMMARY
The University of Utah Clean and Secure Energy (CASE) project is pursuing interdisciplinary, cradle-tograve research and development of energy for electric power generation and for liquid transportation fuels
from the abundant domestic resources of coal, oil sands, and oil shale. Its work is divided into three
programs: the Clean Coal Program, the Oil Shale and Sands Program (OSSP), and the Policy
Environment, and Economics Program (PEEP). Emphasis will be on minimizing the environmental
impacts associated with the development of these resources, including reducing the carbon footprint
through the use of CO2 capture for subsequent storage (sequestration).
During this quarter, the CASE team attended the project kickoff meeting for the Clean Coal Program,
recruited two students for the NETL internship program, and hosted the first Advisory Board meeting. In
the Clean Coal Program, investigators are continuing to integrate the experimental work with the
simulation efforts with an emphasis on the oxyfuel and gasification areas. Progress continues on the
simulation and experimental studies. Specifically, the simulation work is now being extended to provide
a series of oxy-coal and air-coal combustion cases. In addition, the Chemical Looping Combustion
(CLC) Team completed construction of their laboratory-scale reactor, and the Oxyfuel Team is upgrading
the oxyfuel combustor (OFC), repairing the drop-tube reactor and developing a coal burner for the
advanced diagnostics work. The Oxyfuel Team also obtained circulating fluidzed bed (CFB) results
during a brief test campaign, and the advanced diagnostics obtained an initial PIV dataset for a gas flame.
The Gasification Team continues to refine an equilibrium model of the 1 ton/day gasifier and operating
the 1 ton/day gasifier on isopropyl alcohol. They have also drafted papers on in situ and ex situ
measurements of ash emittance, thermal conductivity, sintering rates, and related properties under both
reducing and oxidizing conditions
Research efforts in the six tasks that are the Oil Shale and Sands Program (OSSP) continued this quarter.
The Task 17.0 team has completed building a 20-unit kerogen model and is proceeding to develop
additional models to further assess the pairwise distribution function obtained from X-ray scattering
experiments on a representative kerogen. The Task 18.0 team has completed an experiment to measure
the composition of products as they evolved during the pyrolysis process using thermal gravimetric
analysis/mass spectrometry. The Task 20.0 team has simulated a pyrolysis process followed by in-situ
combustion. Work by the Task 21.0 team this quarter focused on the development of a reliable tool for
creating reaction tables for oxy-fuel combustion application and is now preparing to work on a
demonstration simulation of an oxy-gas burner. To explore the fate of a chemical dissolved in water, the
Team 22.0 team solved the advective-dispersion equation using COMSOL. The problem was solved with
and without sorption and with various hydrogeological controls.
Also during this quarter, the Policy, Environment & Economics Program continued research and datagathering efforts on the Climate Change Legislation and Regulatory Gap, Oil Sands and Oil Shale
Resources, CO2 Emissions, Policy Analysis of the Canadian Oil Sands Experience and Policy Analysis of
Water Availability and Use Issues in the Context of Domestic Oil Shale and Sands Development projects.
The Task 23.0 team has largely completed the compilation of a database identifying roughly 500 industry
players involved in CCS implementation and advocacy and has finished a reviewing and analyzing
legislation generated by the last three Congresses in order to build a comprehensive list of possible
regulatory gaps. The Task 24.0 team met to better define production scenarios for heavy oil, oil shale, and
oil sands and identify potential sources of data for the economic impact analysis modeling. Progress also
has been made on the subtasks associated with Task 25.0, with approximately 280 documents added to
the repository, several new GIS datasets and functional enhancements of the map server developed this
past quarter, and development of a water geodatabase schema to support water-energy management in the
Uinta Basin.
RESULTS AND DISCUSSION
Task 1 – Project Management
The CASE team presented the project objectives and preliminary results at the project kickoff meeting in
mid-June.
Task 2 – Industrial Advisory Board
The Advisory Board met on June 22nd and 23rd to discuss CASE projects and organization (See Appendix
A for the agenda). The Board has drafted meeting notes and recommendations and is in the process of
finalizing these. CASE faculty will be meeting in August to address the Board’s recommendations.
Task 3 – Student Research Experience at DOE NETL
Two University of Utah students are participating in the NETL internship program. Jon Wilkey (student)
is working on air-quality monitoring of oil-and-gas extraction activities in Utah with Richard Hammack
(NETL sponsor). He has been assisting with the setup and operation of an air sampling trailer and
measuring methane in Allegheny National Forest (ANF) with this trailer. He has also been conducting a
literature survey on emissions from oil and gas exploration and development, and training on the use,
operation, and setup of the trailer’s instruments. NETL is planning on stationing the trailer in ANF once
it’s fully operational but has not selected a site for the trailer. To gather information for site selection
purposes, Jon led a team with two other interns to conduct a field survey of ANF for methane and light
hydrocarbons using a spectroscopic instrument from Apogee Scientific. The survey lasted five days and
covered approximately 800 miles of municipal and forest service roads. Jon and the other interns will be
processing the data from the field study and reporting their findings.
Michael Burton has been working in Dr. Schadle’s group to calibrate fiber optic probes for use in a
circulating fluidized bed. Mike has been performing a literature search to identify the best calibration
methods and has tested these methods. After testing, he analyzes the results. In addition, he is
determining the terminal velocity of the different types of bed material that NETL has been using within
the fluidized bed.
Task 4 – Oxy-coal Combustion Large Eddy Simulations
The ODT model has been developed to calculate coal combustion/gasification process. In previous work,
an oxy-coal combustion simulation was completed. This work is now being extended to provide a series
of oxy-coal and air-coal combustion cases. The cases will be valuable for extracting manifolds at
different operating conditions. The operation conditions for these eight cases are listed in Table 1. For
each case, an equilibrium solver was used to get the initial conditions for the co-flow region. The
velocities, temperature and gas composition of the jet are in-line with the experimental work carried out
by experimental group in Utah.
Table 1. Parameters for the eight cases.
The calculation domain of the ODT simulation is a 0.6m×1.0m rectangle. A typical diffusion step is on
the order of 1e-5m. Instantaneous results are acquired from one realization, which usually takes about five
hours on one processor. Average results are calculated by averaging 64 realizations.
A primary challenge in turbulent combustion modeling lies on the broad range of scales which are
inherently coupled. However, it is possible to achieve some degree of decoupling, relying on the
observation that many chemical timescales are significantly smaller than the fluid dynamic scales and that
the thermodynamic state of a reacting system relaxes onto a low dimensional manifold in chemical state
space (Chen et al. 2007). Based on the work of Tognotti et al. (2009) and Parente et al. (2009), a novel
methodology based on principal component analysis (PCA) was adopted to extract manifolds from the
dataset acquiring by ODT simulation for the oxycoal combustion process. The basic idea of PCA is to
reduce the dimensionality of a data set consisting of a large number of correlated variables while retaining
most of the variation present in the original data. In this work, a data set was generated as an m × n
matrix (m " n), where m rows represent m observations at different points in the simulation domain, n
represents all the variables (instant value or averaged value) considered in ODT simulation, which
including coordinate of each point, velocities, temperature, mixture fraction and species concentration, et
al. Different PCA method was used to find the potential manifolds. According to the time consuming
ODT simulation and complexity of the PCA method, this work is still on the way. Some interesting
results will be shown below.
Figure 1a shows the instant temperature profile for the air combustion cases. As the O2 concentration
increases, the flame temperature increases rapidly. At the same time, the beginning point of the flame
moves downward corresponding to a higher temperature. The flame is detached for the normal air
combustion case, while the highest O2 concentration case shows an ignition point very close to the jet.
For comparison, the difference in the temperature profile for the oxy-coal combustion cases and air cases
are shown in Figure 1b. Note that the temperature differences are not that great. However, the ignition
point moves downward as the O2 concentration increases.
(a) Air coal cases
(b) Oxy-coal cases.
Figure 1. Instant temperature for all eight cases.
Figure 2a and b show the average temperature profile for air-coal combustion cases and oxy-coal
combustion cases, based on averaging of 64 ODT realizations. The temperature profile of the air case
appears non-smooth. It appears that there is a stronger fluctuation for the air combustion flame and thus
more realizations would be required for computing the average. Comparing 2a with 2b, for each
equilibrium temperature (Case-1 vs Case-5 etc.), the flame region of the air-coal case is larger than the
oxy-coal case. In these particular cases, the volume flux of the oxidizer for the air case is much higher
than the oxy-coal case. In other words, the air-coal case has a higher Reynolds number and thus a stronger
turbulent mixing process.
(a) Air coal cases
(b) oxy coal cases
Figure 2. Average temperature profiles.
Based on a data set of m×n matrix acquired from the ODT simulation for case 8, principle component
analysis was carried out to find potential manifolds. It is found that the variable space may be dominated
by 2 major manifolds. However, the determination of manifolds depends on the selection of a model in
the PCA method. For example, when the data was scaled by ’VAST’ method, two main manifolds were
found. The effect of manifolds on the data set is shown in Figure 3. This figure illustrates that when the
entire dataset is predicted by the first manifold, the confidence level equals 85%. The confidence level is
raised to 95% when both manifolds are used. These initial results are promising in demonstrating the
ability to find manifolds in ODT data. The linear composition of mixture fraction of moisture, volatile,
char, primary oxidizer and secondary oxidizer plays an important role in the oxy-coal reacting system.
Figure 3. Manifold convergence.
At this stage, the development for the near-burner region of the oxy-coal burner is in line with the
development of the gasifier simulation. Current work is focused on increasing the complexity of the
physics for the particle phase by including more models in the DQMOM approach. Thus, progress for
this task and the LES simulation of near-nozzle gasifier injector overlap, and Subtask 8.1 contains the
results.
Task 5 – Experimental Studies of Oxy-coal Combustion
Subtask 5.1 – Near Field Aerodynamics of Oxy-Coal Flames
Work on design and construction of modifications to the upper furnace of the oxyfuel combustor (OFC)
continues. These include two potential designs:


One with interior cooling tubes embedded within refractory (described in part in previous quarterly
report).
The second with interior cooling tubes placed facing the flame, outside the refractory (described
here).
Although detailed drawings for Option 1 were drawn up for fabrication, we finally decided on Option 2 as
the preferred option. Option 2 entails modification of the current furnace rather than replacement (as in
Option 1). This will be less expensive and easier to complete, and should be equally effective. It also
allows other DOE work on the furnace to proceed without additional interruption.
With this option, the input and output of the tube heat exchanger are connected to the top plate burner,
which allows us to easily replacement. In order to calculate the area of the heat exchanger, we needed to
determine the radiation heat introduced by the flame and refractory to the tube. After some research, the
Oxyfuel team identified a model which describes the view factors inside the chamber between the flame,
tubes, and the refractory. According this model the view factor from flame to each tube is 0.022, and this
allows us to calculate all the view factors existing in the chamber. Figure 4 shows the schematic of the
heat transfer calculation required for design of the modified upper furnace section of the OFC under
Option 2.
Figure 4. Heat transfer calculation schematic for Option 2 design.
The Oxyfuel Team wrote a MATLAB® code to calculate the radiation heat for different sizes, and
numbers of the tubes. Results are shown on Figure 5 and Figure 6. The plots show the amount of heat
radiation to the tubes with different diameters, and also using an empirical correlation we could find the
convection heat transfer in the tubes. Using these plots, calculations, and considering safety factors, we
settled on eight ¾” tubes to provide cooling. The correlations and the formulas include:
Amount of heat by convection from water to the tubes ( W )
4
x 10
9
8
7
6
5
4
3
2
1
0
0.5
1
1.5
2
diameter of the tube ( cm )
2.5
3
Figure 5. Results for Option 2 calculations: heat flux inside tubes vs. tube diameter.
4
Amount of heat by radiation to the tubes ( W )
5.5
x 10
5
4.5
4
3.5
3
2.5
2
0
0.5
1
1.5
2
diameter of the tube ( cm )
2.5
3
Figure 6. Results for Option 2 calculations: heat flux from flam to outer tube surface.
Subtask 5.2 – Ash Partitioning Mechanisms for Oxy-Coal Combustion with Varied Amounts of Flue Gas
Recycle
Work during the last quarter has focused on design and construction of a multi-fuel swirl burner that is
more suitable for maintaining a wide variety of stable pulverized coal flames for a range of coal types and
repairing the drop-tube furnace. Work has also proceeded on the specification and design of the flue gas
recycle system, to replace the once-through CO2 system used currently in the OFC.
Multi-fuel Burner. A multi-fuel burner has been constructed with swirl vanes that will allow for particle
tests on a stable and consistent flame in the OFC. The burner has been constructed to maintain primary
and secondary velocities as close as possible to industrial conditions working within the restrictions of
off-the-shelf pipe sizes. In addition to creating a swirl flow, the newly constructed burner will also allow
for quicker and safer changes between coal and natural gas fuels used to maintain wall and refractory
temperatures between tests. Since the new burner will be capable of utilizing both fuels, operators will no
longer have to manually remove and reinstall burners at high temperatures which will contribute to a safer
overall operation.
Recycling System. The recycle system we plan to use will comprise of the following components:
1. Baseline recycle system in which essentially all particles in the recycled flue gas will be removed by a
baghouse, after which the flue gas is transported by a blower.
2. A baghouse by-pass in which a portion of the particles will be removed by a small electro-static
precipitator, which, hopefully will remove enough of the large particles to prevent damage to the
blower, but still allow a representative amount of fine particles to pass through the blower and enter
the furnace, as would be the case in practice in the field.
A suitable bag house arrangement has been developed for the recycle system, and calculations have been
made to determine the necessary acid dew point and the volumetric flow rate of the recycled flue gas. A
key variable is the pressure drop required for the recycle system. This will be measured across the new
swirl burner as soon as the final welds and leak testing are completed. When the final volume, pressure,
and operating temperature are known, a blower will be selected and ordered.
Drop-tube tests. During this quarter, we have been working on modifications to the drop-tube furnace to
ensure an acceptable material balance and to repair aging components. Most of the modifications have
been completed. However, the top and bottom connecting flanges, which have recirculating water for
cooling, needed to be re-fabricated to meet the larger O.D. alumina muffle tube. The flanges were milled
to allow the outer alumina tube to move smoothly with the flanges. Unfortunately, after the flanges were
constructed, the bottom flange had three water leaks. After four repairs, the flange still leaked under high
water pressure conditions. Consequently, a new same flange was built to replace the broken one.
A new supporting flange was designed and built. The new flange is constructed from aluminum and not
water cooled (Figure 7). One stub flange is welded to the new supporting flange and then connected to
the other stub flange by O-ring and clamps. The new supporting flange provides more precise positioning
of the collecting probe, increasing collection efficiency.
Figure 7. Supporting flange and clamps.
Subtask 5.3 – Oxy-Coal Combustion in Circulating Fluidized Beds
Both bench-scale and pilot-scale fluid-bed experiments were performed during this quarter. A brief
summary of findings is provided below.
Bench-Scale Experiments. A study was carried out to compare air- versus oxy-combustion at three
different reactor temperatures and three different oxygen concentrations Table 2.
Table 2. Range of experimental conditions explored.
Combustion pattern
Air-combustion
Temperature
600℃
Oxygen concentration 10.5%
Oxy-combustion
800℃
21%
870℃
30%
Initial results illustrate CO release in air combustion versus oxy-combustion at two different temperatures.
As seen in Figure 8and Figure 9, the CO production is considerably higher for the oxy-combustion
conditions. Currently, some theoretical explanations for this observation are being developed.
Figure 8. CO release in air- and oxy-combustion for O2=21%,T=870℃.
Figure 9. CO release in air- and oxy-combustion for O2=21%,T=800℃.
Figure 10 and Figure 11 show a comparison between CO and CO2 release at different reactor
temperatures for air combustion only (21% O2 in nitrogen). When the reactor temperature is low
(T=600°C), the profiles for CO and CO2 release are similar. However, when the temperature is high
(T=800°C), the release of CO is delayed compared to the release of CO2. A model-based theoretical
explanation for these findings is currently being developed.
Figure 10. CO and CO2 release for combustion in air, T=600 °C.
Figure 11. CO and CO2 release for combustion in air, T=800℃.
Preliminary experiments indicated that sulfur evolution occurs prior to carbon evolution in this reactor, as
evidenced by a comparison of SO2 and CO2 evolution versus time. This observation implied that the
sulfur oxidation rate was more rapid than the carbon oxidation rate under these conditions. In order to
more clearly view this effect, experiments were carried out at a lower reactor temperature (600°C) (Figure
12). We also initiated a study of the relationship between oxygen concentration and the release of sulfur
and carbon for these conditions.
Figure 12. Sulfur and carbon release at an O2 concentration of 21%,T=600°C.
It is clear from Figure 12 that sulfur release is much faster than carbon release under these conditions, and
its peak concentration occurs after 15 to 30 seconds. Peak CO concentrations occur after approximately
120 seconds.
The CO2 evolution curve was also extracted from the FT-IR spectra and was found to track the CO
evolution curve almost identically when the reactor temperature is 600°C. Thus, we can conclude that the
rate of sulfur release is faster than carbon release in oxy-combustion at 600°C. Similar results were found
for air combustion.
Sulfur evolution at a reactor temperature of 870°C was also measured under both air- and oxy-combustion
conditions (oxygen = 21%), and the results show that sulfur release in air combution is very close to that
of oxy-combustion at this higher temperature (Figure 13).
Figure 13. Sufur and carbon release for an O2 = 21%, 870°C.
Preliminary experiments were carried out at a reactor temperature of 600°C on the effect of oxygen
concentration on sulfur evolution. Figure 14 shows that higher oxygen concentrations result in higher
peak SO2 concentrations in oxy combustion and a more rapid generation of sulfur overall. Note that the
SO2 concentration reaches its maximum at approximately the same time in the different oxygen
concentrations.
Figure 14. SO2 release for different oygen concentrations in oxy-combustion.
In order to verify the SO2 concentrations extracted from the FTIR, we made comparisons with a
Continuous Emission Monitor (CEM) designed for SO2. The hypothesis was that measurement of SO2
release by the gas analyzer (CEM) would be more accurate than extracting the SO2 information from the
FTIR spectra, due to the IR interference of CO2 and H2O. We selected one condition to use for the
comparison (T=870°C, O2 concentration=30%, oxy-combustion conditions). The results indicate that very
similar results are obtained using either the SO2 CEM or the FTIR.
Figure 15. Comparison of SO2 Measurements in FTIR and SO2 CEM.
Pilot-Scale Circulating Fluidized Bed Reactor. A brief testing CFB campaign was run during this past
quarter at two conditions (baseline air and oxy-combustion with 30% O2 and recirculated flue gas) to
characterize the carbon content of the bed at steady state conditions. Figure 16 shows the steady-state
temperature and emission levels. The 30% O2 condition was about 90oF hotter than the baseline
condition. NOx and SO2 emissions were relatively constant over the test, but CO emissions varied
widely. The tests were conducted with a 3% exit oxygen concentration.
Steady State Emissions and Temperatures
6.0
1800
5.0
1500
4.0
1200
3.0
900
2.0
600
1.0
300
0.0
0
Temp (F)
Conc. (lbs/MMBtu)
50 lbs/hr Utah Coal, 4010 Quartz Sand (300 mm)
600 F Primary Preheat, 80% Primary
FGR Test was with 30% O2
NOx (Air)
NOx (FGR/O2)
CO (Air)
CO (FGR/O2)
SO2 (Air)
SO2 (FGR/O2)
Bed Temp (Air)
Bed Temp (FGR/O2)
Figure 16. Emission and temperature information from pilot-scale CFB tests for air and oxy-combustion
conditions.
Solids samples were collected at eight different locations (bottom ash screw, one each in sections 1-5,
loop seal, and flyash collected in the baghouse), and the samples were taken at 2 hour increments to
determine their carbon content. In addition to these samples, other samples were taken in section 3 on 30
minute increments. All of these solid samples are currently being analyzed for carbon content and will be
summarized in the next quarterly report.
Task 6 – Advanced Diagnostics for Oxy-Coal Combustion
Gaseous flames consisting of ethylene, as the fuel, nitrogen and air at varying ratios were used to take
preliminary PIV data on gaseous flames in order to investigate the feasibility of the PIV setup. A
preliminary data set has been completed.
The experimental setup included:
•
Camera: Photron High Speed cam (1k x 1k format)
 Frame rate: 1500-3000 fps (1 K x 1K
 Shutter speed: 1/ frame rate (333-166 us)
•
Laser:
•
Lenses: (spherical and cylindrical) for laser sheet formation
•
532 nm band pass interference filter to reduce luminosity of flame
•
Seeder : Fluidized bed solid particle seeder (PB 100, Lavision)
•
Seeding particles: Aerosil (through air channel up to 2 bars)
•
PIV software: LaVision with Davis 7.2
•
Cross Correlation method for analysis, IA 32 x 32 Px
•
Time delay unit for camera and laser pulses: Labsmith (Software: Trigger)
PIV Q-switched double-headed Continuum laser
 Repetition rate: 15 Hz
 Energy per pulse: 25 mJ
 Time between PIV laser pulses: 666-166 us
There are many parameters in a typical PIV experiment that must be chosen properly in order to get
reasonable data. Among these parameters are the time delay Td between laser pulses (same as 1/frame
rate of the camera in this experimental setup) and the size of the interrogation area (IA). Figure 17 shows
the image maps and corresponding vector maps for the gaseous flame with camera framing rates of 1500
and 3000 fps. These correspond to Td of 666 and 333 ms, respectively. The seeding particles can be seen
in the flame image. As shown, using the 1500 fps results in underestimated results, with the vector maps
not representing the flow motion properly. However, using the higher framing rate of 3000 fps represents
the flame well, as expected.
1500 fps
3000 fps
Figure 1. Image map (top) and vector maps (bottom) of the gaseous flame at 1500 and 3000 fps.
Effect of Varying IA
Figure 17. Image map (top) and vector maps (bottom) of the gaseous flame at 1500 and 3000 fps.
The next step is to investigate the effect of the IA size on the PIV measurements. Figure 18 represents the
vector maps at two different IAs, namely 16 pix x 16 pix and 32 pix x 32 pix. In both cases, the camera
frame rate is kept at 3000 fps. The choice of the IA size depends on the resolution of the camera and the
range of measured velocities. The IA size needs to be larger if the velocity is high to accommodate for
the seeding particles in image pairs. Using a smaller IA produces overestimated velocity values.
16 pix x 16 pix
32 pix x 32 pix
Figure 18. PIV vector maps of the gaseous flame using two different IAs. Camera frame rate was 3000
fps.
The vertical velocity (the y-component of the velocity) was investigated at five different heights above
the burner for a gaseous flame. Figure 19 shows the symmetry around the burner center and the variation
in the vertical velocity, at different heights. Peak velocities continue to decrease with increasing height
until a height of 50 mm, where the velocity profile is nearly flat.
Figure 19. Vertical velocity, as measured by PIV, at five different heights above the burner. Camera
frame rate was 3000 fps, with an IA of 32 pix x 32 pix.
Now that preliminary data has been obtained with an ethylene flame, the next step will be to complete a
lab-scale pulverized coal (PC) burner and then adapt the PIV system to a PC flame.
Design and fabrication of a laboratory coal burner. A coal burner has been designed and is currently
being fabricated for implementation with the PIV system. The design of this burner uses aspects of the
design of other bench-scale burners reported in the literature, as well as our own expertise in burner
design. The burner utilizes four concentric streams to control the ignition and flame shape (see Figure
20). In this design, the coal is fed via a gas stream through the central tube of the burner. Immediately
circling the central coal jet is a natural gas pilot whose purpose is to ignite the coal. Outside of the pilot
ring is an air feed that will provide air or oxygen/CO2 to both the pilot and coal flames and will allow for
control of the stoichiometric ratio during coal combustion. The final annular ring (outside of the air
stream) is a pre-mixed natural gas/air stream. The primary purpose of this stream is to heat the air stream,
facilitating ignition of the coal jet, and to reduce heat loss from the flame.
Figure 20. A computer-aided model of the coal burner for PIV measurements. The central tube is the
coal jet while the surrounding annulus is the air stream. The two pinhole rings are natural gas pilots.
The basic design parameters for the burner are listed in Table 3. The dominant design parameter for this
burner is the coal feed rate of 2 g/min. It is anticipated that this is the maximum necessary feed rate to
ensure a large enough flame for PIV measurements. The burner should be capable of operation at lower
flow rates if this coal feed rate proves to be too large. The velocity ratios were chosen to ensure proper
control of the combustion; coal ignition close to the burner face is desired, as well as adequate shear
mixing between the air stream and coal jet. It was also assumed that the pilot flame needed to provide
10% of the overall thermal input to assist the coal flame with ignition. Based upon the desired velocity
ratios and the necessary mass flow rates to achieve stoichiometric combustion of both the natural gas and
coal streams, it was possible to calculate the necessary outlet areas for each stream.
Table 3. Design parameters for the PIV coal burner system.
Parameter
Design
Actual
Coal Feed Rate (g/min)
2
2
Coal Jet Velocity (ft/s)
4
4.1
Coal/Pilot Vel. Ratio
1:1
1.03:1
Air/Pilot Vel. Ratio
2:1
2.18:1
Air/Pre-Mixed Vel. Ratio
1:1
1.09:1
Based upon the design calculations, the burner geometry was laid out as shown in Figure 20 . For
simplicity of construction, the burner was designed to be assembled from standard piping. To achieve the
necessary velocities from the low flow-rate gas streams, they were capped and vented with symmetrically
spaced holes. Using standard pipe sizes and the capping strategy, the burner comes reasonably close to
achieving the desired velocity ratios (see Table 3 ). Because the velocities have been specified to be an
order of magnitude larger than the largest flashback velocity and an order of magnitude smaller than the
draw velocity of the ventilation system in the combustion chamber, there should be a wide range of
adjustability in the flow rates should the current design be insufficient for maintaining a stable flame.
Task 7 – Fate and Control of Mercury in Coal-based Power Generation Systems with CO2
Capture
Using the high-quench temperature profile of 440 K/s (Figure 21), CHEMKIN was used to predict the
extent of homogeneous oxidation of mercury by bromine under standard and oxy-fuel conditions. Figure
22 shows that elemental mercury is almost completely converted to HgBr2 after 2.5 seconds for standard
and oxy-fuel conditions. The primary reaction product is HgBr2. The extents of oxidation obtained with
bromine and with chlorine are compared in Figure 23. With chlorine there is negligible oxidation. The
rate of oxidation with bromine is retarded by oxy-fuel conditions. The predicted profiles of Br, Br2, and
HBr do not offer any clues that explain this prediction.
Figure 21. Laboratory and industrial temperature profiles.
Figure 22. Mercury and HgBr2 concentration profiles for standard and oxy-fuel conditions using the high
quench profile temperature profile (440 K/s).
Figure 23. Fraction of mercury oxidized by bromine and chlorine for standard and oxy-fuel conditions
using the laboratory, high quench temperature profile (440 K/s).
Similar calculations were performed with the industrial boiler temperature profile. Figure 24 shows
predicted mercury and HgBr2 concentration profiles for conventional and oxy-fuel conditions. With the
industrial profile, complete mercury oxidation by bromine is observed, the extent of oxidation is
unaffected by the combustion conditions.
Figure 24. Mercury and HgBr concentration profiles for conventional and oxy-fuel conditions using the
industrial temperature profile.
The bromine data plotted in Figure 24 are also plotted in Figure 25 to show the fractional oxidation.
Figure 5 includes the fractional oxidation with chlorine. At all conditions examined, the extent of
oxidation by chlorine is negligible.
Figure 25. Fraction of mercury oxidized by bromine and chlorine for standard and oxy-fuel conditions
using the industrial temperature profile.
Task 8 – Entrained-Flow Coal Gasifier Simulation and Modeling
Subtask 8.1 – Large Eddy Simulations of the University of Utah Entrained-Flow Gasifer
Models adopted here are the same as those used in Task 4. An additional spatial consideration is
required for the gasifier in that the mass loading rate is much higher than the oxy-coal cases. In order to
get a convergence simulation, a smaller diffusion step will be required.
The Direct Quadrature Method of Moments has been implemented in the ARCHES LES code to model
the transport of the number density function of the coal particle phase. For LES simulation of the near
burner region, a number of three internal coordinates was chosen:
• the particle diameter dp
• the particle temperature Tp
• the mass of “raw coal” in the particle c
Here, we have introduced a new internal coordinate (Tp) that has not been included in previous work.
Previously, coal particles were assumed to be at the gas temperature at all locations in the flow. Now, the
particles will be allowed to heat up independently according to their history in the turbulent jet. In the
current approach, a homogeneous temperature distribution for each particle is assumed (Biot -> 0). Two
quadrature nodes were (α = 1, 2) chosen, i.e. the distribution of each internal coordinate is represented by
two values.
At this stage of the code development, the particle size is assumed to remain constant. The Kobayashi
model has been implemented to model the devolatilization of the raw coal and a heat transfer model
between the gas phase and the solid phase has also been implemented to compute the heating rate of the
particles.
A test simulation with fairly coarse resolution was performed and consists of a coaxial jet where a
mixture of air and particles is injected in the center and a mixture of air and fuel gas (i.e. methane) is
injected in the surrounding inlet. The fuel gas was added to be a source of ignition for the particles. In
the inlet, two quadrature nodes are initialized with particle sizes of 35 and 75 μm. The quadrature nodes
are initialized to the to temperatures of 298 and 357K for the small and large abscissa. The particles
require different temperatures at the inlet to avoid a singularity in the systems of equations. The
qualitative behavior of the simulation is good. Results show that the small particles heat up faster and
devolatilize faster than the large particles, as expected. A small amount of volatiles near the nozzle
released by the small particles is observed yet the majority of the volatiles are released further
downstream when larger particles are heated up. The coal gas mixture fraction is shown in Figure 26
along with the source of coal gas from the reacting particles in Figure 27.
Figure 26. Mixture fraction of the volatiles from the coal.
Figure 27. Rate of production of the mixture fraction of the volatiles from coal.
Task 8.2 – Sub-Grid Scale Models
Principal component analysis is a rigorous methodology to compress data, or identify a new basis that
best approximates the original system using a reduced number of variables. It is based on an eigenvalue
decomposition of the correlation matrix of the original data set. Details on the PCA approach were
presented in a previous quarterly report.
PCA is a linear transformation, which is attractive since if one applies it to variables such as temperature,
pressure and composition it allows one to define transport equations for the principal components. To
complete the modeling approach, one must regress source terms for the principal components onto the
principal components. However, it is not clear what basis functions to use. We have adopted a
multivariate, adaptive regression technique (MARS) as originally proposed by Friedman (1991). This
allows dynamic identification of basis functions to provide an optimal regression of data onto a desired
set of independent variables. Table 4 shows the R2 values for the original state variables as reconstructed
well by just two principal components. Furthermore, the nonlinear regression provides a marked
improvement in the ability to represent the data over the PCA approach. Analysis of source terms using
MARS is currently underway.
Table 4. R2 values for regression of the original state variables using PCA and MARS approaches.
n
T
H2
O2
O
OH H2O
H
HO2 H2O2 CO
CO2 HCO
1 0.907 0.614 0.000 0.419 0.589 0.868 0.473 0.014 0.001 0.487 0.704 0.059
PCA
0.9 0.97 0.99 0.41 0.59 0.94 0.52 0.18 0.21 0.99 0.78 0.12
2
86
1
0
9
9
1
6
4
9
3
8
5
1 0.932 0.779 0.343 0.518 0.842 0.892 0.554 0.220 0.219 0.653 0.772 0.147
MARS
0.9 0.98 0.99 0.89 0.97 0.97 0.79 0.73 0.63 0.99 0.86 0.35
2
89
8
7
2
3
9
7
9
5
7
3
5
The ODT model formulation developed as part of this project was outlined in the previous quarterly
report, where preliminary results were also shown. After performing additional verification and validation
studies on the code, several aspects of the model were improved. Figure 28 shows results for the
averaged mixture fraction profile across the jet at various points in time, and indicates remarkable
agreement between the three-dimensional DNS simulation and the one-dimensional ODT simulation.
Figure 28. Mixture fraction profiles as a function of time comparing DNS results
(points) with ODT results (lines).
Results for the velocity are also in very good agreement. This illustrates that ODT can reproduce the
correct mixing behavior observed in three-dimensional turbulent flows.
Figure 29 shows the average temperature (averaged in the streamwise direction) as a function of the
normalized distance across the jet. This also shows strong agreement between the ODT and DNS results,
Figure 29. Temperature profiles at various points in time for the DNS (points) and
ODT (lines) simulations.
2 - Temperature profiles
at various
points in time in
forODT.
the DNS
suggesting Figure
that the turbulence-chemistry
interactions
are also well-represented
Results for minor
(points)
andsimilar
ODTagreement
(lines) simulations.
and major species
show
between the DNS and ODT results.
Task 9 – Char and Soot Kinetics and Mechanisms
Investigation of Pressurized Pyrolysis and Char Conversion. Pyrolysis experiments at 2.5 atm were
conducted for Wyodak subbituminous coal in the previous quarter using the pressurized flat-flame burner
(PFFB) with a fuel-rich H2/CH4/O2/N2 flame. Ultimate and ICP analyses of these samples were
conducted in this quarter. Table 5 shows the ultimate analyses of the coal and chars. The ash and ICP
data used to calculate mass release are listed in Table 6. The mass release, apparent density and swelling
ratio from these experiments are presented in Table 7.
Predictions of the time-dependent pyrolysis behavior of this coal were performed using the CPD model
(Grant et al., 1989; Fletcher et al., 1990; Fletcher et al., 1992) and the measured gas temperature profile.
The CPD code predicted 62% mass release (67% measured) on a dry, ash-free basis (daf) from pyrolysis.
The additional mass release due to gasification was large, mainly occurring between 30 and 60 ms in this
experiment (Table 7). The 30 ms char was fully pyrolyzed, with perhaps a small extent of gasification.
Calculations of steam gasification rates indicate that the extent of gasification observed at 60 ms is
possible if film diffusion is the rate-limiting step.
Table 5. Ultimate Analyses (daf) of Wyodak Coal and Chars Produced at 2.5 atm and 1640 K.
Sample
C wt%
H wt%
N wt%
S wt%
O wt% (by diff)
Wyodak Coal
72.25
5.30
0.94
0.50
21.01
30 ms char
91.14
1.11
1.06
0.27
6.42
60 ms char
92.50
1.16
0.86
0.39
5.10
220 ms char
92.69
1.22
0.94
0.53
4.62
Table 6. Ash and ICP (dry basis) of Wyodak Coal and Chars Produced at 2.5 atm and 1640 K.
Sample
Ash wt%
Al wt%
Si wt%
Ti wt%
Wyodak Coal
5.62
0.38
0.56
0.043
30 ms char
15.16
1.15
1.87
0.122
60 ms char
32.5
2.84
4.52
0.309
220 ms char
43.52
3.44
6.62
0.382
Table 7. Characteristics of Wyodak Chars Produced at 2.5 atm and 1640 K.
Sample
Mass Release (wt% daf)
Apparent Density g/cc
Swelling Ratio (d/d0)
Wyodak coal
N/A
1.49
N/A
30 ms char
67.2
0.76
0.86
60 ms char
87.1
0.63
0.72
220 ms char
90.5
0.59
0.71
The swelling ratio was obtained using the mass release on an as received basis and the apparent density
ratio obtained from the tap density technique, according to the following equation:
m


m0  0
d 
 
 d0 
3
The changes in char physical structure observed between 30 and 60 ms show that both density and
particle diameter are decreasing. The particle structure does not change much at longer residence times in
this experiment. The SEM photos in Figure 30 indicate the development of a very porous structure in the
Wyodak particles with increasing gasification residence time. These data suggest that Zone III
gasification may occur in the near-burner zone, but much slower rates prevail thereafter where the
temperature is lower. An alternative explanation suggested by the significant change in density from 3060 ms is that O2 was present and Zone II oxidation and gasification occurred. If such is the case, it could
have been caused by channeling of fuel or oxidizer in the burner or poor calibration of the mass flow
controllers. The mass flow controller calibration is being checked and a convenient means to accurately
measure O2 in the post-flame region is currently being designed.
Modifications were completed during this quarter to allow the PFFB to safely use CO as a fuel. The use
of CO as a fuel offers several benefits. First, CH4 has been observed to produce soot at pressures above 2
atm and CO does not. The soot from CH4 would contaminate soot and char samples produced from coal
pyrolysis and gasification experiments. The method to avoid production of soot from CH4 was to use a
high percentage of H2. However, doing so increased the flame speed and preheated the burner surface
until it glowed, causing premature pyrolysis of coal to occur in the feeder tube. For bituminous coals the
premature pyrolysis caused the feeder tube to clog. For this reason only the Wyodak subbituminous coal
was used with the CH4/ H2 settings. It has now been demonstrated that the use of CO as the fuel
eliminated the heating of the burner surface and permitted experiments with bituminous coals without
clogging. Another benefit of using CO is that CO2-rich gasification studies can be conducted. In addition,
reactor start up time has been significantly reduced because the soot filters no longer need to be preheated
above 100o C to prevent condensation.
Temperature profiles were measured at 2.5 atm in the PFFB with CO as the primary fuel. During these
studies if the concentration of H2 is kept low, the burner surface did not glow, and bituminous coal can be
fed through the burner. However, preliminary experiments showed that no soot from coal pyrolysis was
found at an equivalence ratio of 1.1, and the mass release was near 90% at a very low residence time (~25
ms). For these reasons it is suspected that O2 may have been present along the centerline with both this
CO-fueled case and with the CH4/H2-fueled Wyodak settings. When the equivalence ratio was increased
to 1.3, soot from coal pyrolysis experiments was observed (as desired), and the observed mass release is
lower, indicating perhaps that no O2 was present in the more fuel-rich condition. The temperature profile
at this condition will be measured to make sure this is a desirable condition for experiments on coals at
2.5 atm.
Modifications are being made to use a smaller burner in the PFFB. This will increase the maximum
velocity of gases through the burner with the same flow control system. It will also help prevent a
turbulent mismatch between the velocities of the coal entrainment stream and the burner, especially at the
highest pressures. Further changes to allow steam to be added to the burner gases are being investigated.
This would allow H2O-rich gasification studies to be conducted without the detrimental effects of high
hydrogen concentration.
(a) Parent coal
(b) Char at 30 ms
(c) Char at 60 ms
(d) Char at 220 ms
Figure 30. Representative SEM photos of Wyodak coal and char particles produced at a peak
temperature of 1640 K and a pressure of 2.5 atm.
Investigation of Soot Formation during Gasification. Soot formation from coal tar has been
investigated through the use of two coals, a Wyodak subbituminous coal and an eastern bituminous coal.
These have been subjected to peak flame temperatures of 1150 K, 1300 K, and 1450 K in the atmospheric
flat-flame burner (FFB) at BYU. Data have been collected on the parent coal and all three tar/soot
samples of the Wyodak coal. Data have recently been obtained on the Eastern Bituminous coal and the
tar/soot sample obtained at 1450 K. The 1150 K and 1300 K samples are awaiting NMR instrument time
for data acquisition. The 13C-NMR spectra of the Wyodak subbituminous coal and the tar/soot produced
from it at 1450 K and 19 ms are shown in Figure 31. The corresponding spectra from the eastern
bituminous coal are shown in Figure 32. The aromaticity of the two coals is practically identical (6567%), but the shape of the aromatic peaks (~100-140 ppm) indicates that the structure of the aromatic
carbon is very different. Quantitatively, the carbon aromaticity was 83% for the Wyodak tar/soot and the
eastern bituminous tar/soot was approximately 100% aromatic. The aliphatic peaks (0-60 ppm) are
observed to decrease dramatically for the Wyodak sample and disappear entirely for the eastern
bituminous sample. The average number of carbons per cluster increased 70% (from 11.2 to 19) for the
Wyodak coal and 244% (from 16.8 to 41) for the eastern bituminous coal. This is an interesting
comparison. The two coals have similar carbon aromaticities, but the tar to soot transformation is quite
different. The two coals are of different rank, and have quite different chemical structure (as indicated by
the average number of carbons per cluster and the average molecular weight per side chain, M). This
finding is consistent with the recent work by Zeng and coworkers at Babcock and Wilcox (Zeng et al.,
2009).
fa = 0.83
fac = 0.05
Cav = 19
σ+1 = 5.1
B.L. = 4.1
fa = 0.65
S.C.
=1
fac = 0.09
MW = 318
Cav = 11.2
Mδ = 16
σ +1 = 3.2
B.L. = 0.6
S.C. = 2.6
MW =31.
33313C-NMR spectra and structural parameters for Wyodak coal and
Figure
tar.
Mδ = 61
fa = 1.0
Fac = -Cav ~ 41
σ +1 = -B.L. = -S.C. = -MW = 527
fa = 0.67
Mcδ = 0
Fa = 0.02
Cav = 16.8
σ +1 = 5.9
B.L. = 2.8
S.C. = 3.1
Figure 32.
13
C-NMR spectra and structural parameters for eastern bituminous coal and tar.
MW = 371
Mδ = 28Spectroscopy (FTIR) was used to identify gas species produced from the coals
Fourier Transform Infrared
during pyrolysis. The data indicate that 38.5% of the nitrogen in the Wyodak coal was converted to HCN
at this condition, and 30.7% of the nitrogen in the eastern bituminous was converted to HCN. No NH3
was detected in either case. For the Wyodak coal the mass yields of gaseous species were 3.0% CH4,
5.5% C2H2, and 0.6% C2H4 (on a dry basis). For the eastern bituminous coal the mass yields of gaseous
species were 3.0% CH4, 10.2% C2H2, and 2.8% C2H4 (on an as received basis).
Model Compound Studies. Several model compounds were selected for study as surrogates for tar/soot
formation. The initial surrogate selected was 2,6 dimethylnaphthalene. This was a structure that was
simple and provided valuable base line data for study of other surrogates. At a temperature of 1447 K
essentially all of the methyl groups have disappeared in the solid state C-13 NMR spectra. It appears that
new CH2 and/or CH groups have formed as bridges between the aromatic clusters which have an average
of approximately 32 carbons per aromatic cluster. As the temperature is raised to 1500 K the aromatic
cluster size grows to approximately 125 carbons/cluster.
Several other surrogates have been selected that are presently being evaluated. One of these model
compounds is 6-(5H)-phenanthridinone (PAD). The 13C NMR spectrum from the 1450 K, 19 ms
condition shows a cluster of peaks in the aromatic region, indicating that some of the carbonyl groups are
still present after pyrolysis at 1450 K and only modest changes are noted in the aromatic structure. For the
phenanthridinone, the mass yields of gaseous species were 1.5% CH4, 1.8% C2H2, and 1.2% HCN as
determined by FTIR. Experiments with the phenanthridinone at 1500 K and 1550 K have been completed
on the FFB but the tar/soot samples have not yet been analyzed. These higher temperature experiments
should produce compounds with 13C NMR spectra that exhibit a single peak in the aromatic region that
can be analyzed quantitatively.
Task 10 – Physical and Chemical Aspects of the Transformation from Char to Slag
During this quarter, ash deposition experiments were performed using a laminar entrained-flow reactor
with a specially designed deposition probe. The probe consists of an alumina tube approximately 1.25
inches diameter and an alumina plate roughly 1.25 inches on a side. A 60º (with respect to the crosssectional plane of the probe) V-shaped groove was cut on top of the alumina tube to hold the alumina
plate. The uncooled plate rests in this groove and allows gas to flow past and into the tube on both sides
of the plate. The probe is inserted into the bottom of the reactor and extends to a position such that the
plate is in the hot zone of the reactor. This configuration is meant to allow inertial impaction on a plate in
near-parallel flow. If particles are “sticky,” for example if there or molten or partly-molten mineral
inclusions on the surface, they will adhere to the plate. If the particles are below the condition of being
sticky, they will bounce off the plate and continue into the collection probe with the gas flow. The
experimental procedures and conditions are the same as in the char-slag transition experiments in our
previous report.
The ash stickiness, represented as the fraction of material passing the plate which sticks to the plate was
calculated and is presented in Figure 33. Generally, ash stickiness increased with carbon conversion at
both 1400 and 1500 ºC. In particular, the ash stickiness of 1400 ºC sample remained relatively low below
85.7% conversion and then increased dramatically as conversion proceeded to 89.1%. The ash stickiness
of 1500 ºC sample followed the same trend.
Figure 33. Evolution of ash particle stickiness as a function of carbon conversion.
The data in Figure 33 show that there is a critical conversion for the Illinois #6 coal at which the ash
stickiness increased dramatically. The sharp rise of ash stickiness coincides with the fast drop of internal
surface area of char particles described in the previous report. To explain this phenomenon, the microimages of the particles before and after the critical conversion were captured using a scanning electron
microscope and are presented in Figure 34. It can be seen that before the critical conversion, few
minerals appeared on the char surface. After the critical conversion, a lot of minerals attached on the
particle surface, making it sticky. This indicates that the transition from porous char to molten slag is
responsible for the dramatic change of ash stickiness at the critical conversion.
Figure 34. SEM images of the char/ash particles of different conversions: Note the dark carbon and the
bright mineral grains due to their different abilities in scattering electrons.
Task 11 – Slag Chemistry and Slag-Wall Interactions
During this quarter, we completed investigations and draft papers on in situ and ex situ measurements of
ash emittance, thermal conductivity, sintering rates, and related properties under both reducing and
oxidizing conditions. Among the most important findings, ash deposit structure, independent of porosity,
has a major impact on deposit thermal conductivity and sintering. Deposits tend to initially exhibit
tenuous connectedness among deposited ash. Fine particles, in the absence of disruptive large-particle
impacts, form filaments on the surface and exhibit primarily intra-filamentary, as opposed to interfilamentary, sintering. That is, particle necks along individual filaments thicken, but filament do not
sinter to one another. Deposit porosity decreases only modestly during this process, but thermal
conductivity changes from very low values to near the upper bound for a given porosity, largely because
of neck thickening driven by sintering. Subbituminous ash forms deposits with lower thermal
conductivity than bituminous ash. Thermal conductivities of both ashes are essentially independent of
local stoichiometry if all else (deposit composition and morphology, deposit thickness, and deposit
porosity) remain constant. However, reducing conditions favor sintering, especially for bituminous coals,
so functionally reducing conditions often increase conductivity. Deposit emittance is also significantly
lower for subbituminous ash than bituminous ash, in part because of differing composition (lower iron
content and higher calcium and sodium content) and in part because subbituminous ash has smaller
particle sizes. The effective total emittance (Planck-weighted values) decreases with increasing deposit
temperature for both types of deposits and for both reducing and operating conditions. Subbituminous
ash emittance is relatively independent of local stoichiometry while reducing conditions often increase
bituminous coal ash emittance, which is in any case higher than that of subbituminous coal ash. The
primary reason for the change with stoichiometry is a shift in oxidation state of iron (from +3 under
oxidizing conditions to +2 under reducing conditions). Slag exhibits uniformly higher emittance than
non-molten or non-sintered ash for all fuels, but the emittance retains some spectral character and also
decreases with increasing deposit temperature. The temperature dependence in emittance primarily arises
from a largely temperature-independent spectral emittance signature that is characteristically higher at
high wavelengths than at low wavelengths. Physical changes in deposit structure, including sintering/slag
formation, change both the spectral and total emittance values.
Also during this quarter visitors at DOE in Albany and at Pittsburgh engaged in on-site refractory and
remote modeling interactions, respectively.
Task 12 – Characterization of Conditions in a 1 ton/day Entrained-Flow Gasifier for Simulation
Validation
Gasifier Model Development. During the second quarter of this year, progress was made on the
equilibrium model. Foremost, the energy balance has been much improved. The current version can now
predict the adiabatic flame temperature accurately to within a couple of degrees only given fuel’s
composition and lower heating value. The change in enthalpy of the system is calculated using
where LHV is the lower heating value of the fuel, the nmax terms indicate the maximum amount of the
indicated species possible during the reaction and NS is the total number of species in the system. This
equation compares the lower heating value of the fuel, but subtracts out the difference between the
maximum theoretical heat release and the heat release from the actual calculated products.
Once the change in enthalpy is known, the exiting temperature of the product stream can be calculated
using
The HLoss term accounts for any possible heat loss from the reactor.
Adiabatic flame temperatures offer a great opportunity to perform both verification and validation. The
literature values of flame temperatures can be used as a form of validation, since the literature values are
based on real world experiments. Verification can be performed by comparing the calculated values of
another equilibrium calculator. In this case, GasEQ is used to perform the verification.
Figure 35 shows that the new calculated adiabatic flame temperatures are within a couple of degrees of
both the literature values and those calculated by GasEQ. The small differences that are present are likely
caused by the difference in thermodynamic databases used by the two computer codes. GasEQ uses the
NASA database and equations while the simulation uses the database from NIST and the Shomate
equations.
Figure 35. Adiabatic flame temperatures determined from different methods.
Improving the energy balance portion of the model has resulted in improvements in the accuracy of
composition predictions as well.
Gasifier Testing. During this quarter the entrained-flow gasifier was operating using isopropyl alcohol
(IPA) as a fuel. IPA was chosen because it is well characterized, ignites easily and is easier to feed than
coal slurry. The primary motivation for the IPA testing was to confirm performance of the quench
system, reactor control and safety systems. The IPA was gasified at stoichiometric ratios of 0.3, 0.45 and
0.6 at atmospheric pressure. Temperatures in excess of 2500°F were recorded inside the reactor. These
tests confirmed that the quench system is able to operate stably when gasifying a liquid feed, that the
quench can efficiently cool the raw syngas, that the thermocouples, flow meters, etc. function properly
and that the cooled gas can be sampled and analyzed. This first round of IPA tests was conducted at low
pressure. The plan for the next quarter is to test system performance at high pressure and to finish the
prototype coal slurry feed system.
Task 13 – CLC Kinetics
While awaiting parts for the Thermax 500 pressure system, we were running experiments with the
Cu/CuO and Ni/NiO systems in the Department of Chemistry. The oxidation of Ni and Cu with air were
investigated (Figure 36). The CuO spontaneously decomposes under nitrogen atmosphere as shown in
Figure 37. The NiO, however, did not.
The oxidation and decomposition reaction of the Cu/Cu2O/CuO system were carried out repeatedly. Such
a cycling experiment is shown in Figure 38. Details of the same reaction are presented in Figure 39. For
this cycling experiment data sets were collected at different temperatures and various gas exposure times.
The plot in Figure 40 is our longest experiment of 201 cycles. The oxygen yield was small, and the
reaction was remarkably repeatable. The evaluation of the collected data is in progress.
125
Theoretical yield: 125.2%
115
110
1000
Temperature (°C)
800
105
600
400
200
0
100
0
20
40
60
80
100
Time (min)
0
20
40
60
80
100
120
140
Time [minute]
Figure 36. Oxidation of Cu with air.
1000
100
800
Relative weight
Temperature
600
95
400
200
90
Theoretical yields: 89.9%
79.9%
Cu2O
Cu
0
0
50
100
150
200
250
300
Time (min)
Figure 37. Decomposition of CuO in nitrogen.
Temperature (°C)
Relative weight (%)
Relative Weight [%]
120
104
Air
Relative weight (%)
102
100
98
96
N2
94
92
90
0
500
1000
1500
2000
2500
2000
2500
Time (minute)
1000
Temperature (°C)
800
33.5 cycles of air (30 min) and N 2 (40 min)
600
Sample: CuO
Temperature: 850 °C
Notice the evolution of the bottom!
400
200
0
0
500
1000
1500
Time (minute)
Figure 38. Decomposition and oxidation of CuO.
Relative weight (%)
102
96
90
240
300
360
420
480
1980
2040
Time (minute)
Relative weight (%)
102
96
90
1800
1860
1920
Time (minute)
Figure 39. Details of the Cu cycles.
Relative weight (%)
100
99
98
97
96
0
500
1000
1500
2000
2500
3000
Time (min)
Figure 40. Long looping of the copper with nitrogen and air.
Task 14 – Laboratory-Scale CLC Studies
During this quarter, construction of the lab-scale chemical looping fluidized bed system was completed.
A photo of the system is shown in Figure 41. The reactor/furnace, valves, rotameters, furnace controller,
electrical box, and filter have been mounted to a sheet of plywood which has been hung on a steel frame
secured to a wall. The fluidized bed reactor is constructed of quartz in order to withstand temperatures
upwards of 1100°C. Type K thermocouples are being used to monitor the temperature about 1 inch below
the distributor grid and within the bed itself.
Figure 41. Lab-scale chemical looping reactor setup. The quartz fluidized bed reactor is shown (and
difficult to see) in the open clamshell furnace. The analyzer for CO, CO2, O2 and CH4 is in the lower
right corner. Solenoid valves and all gas lines for the system are hidden behind the plywood panel.
For an oxygen carrier, 99.9% pure copper particles have been purchased with a mesh maximum (which
corresponds to a maximum diameter of about 150 microns). The material has been sieved to separate size
fractions. Initial testing will take place using particles in the range 100-150 microns.
The gas supply system is configured to allow either air or nitrogen to be fed to the bed. The gas flow to
the reactor will be switched from one gas to another at a pre-determined time interval using solenoid
valves on the gas lines. The signal from the gas analyzer outputs will be interpreted and gas
concentrations will be logged using LabView hardware and software.
Task 15 – Process Modeling and Economics
Material and energy balance calculations were performed on a spreadsheet and ASPEN PLUS for a
preliminary conceptual design for the chemical looping combustion of coal (Figure 42). Oxygen for
combustion is supplied by copper (II) oxide via the mechanism of chemical looping with oxygen
uncoupling as proposed by researchers at Chalmers.
Cu2O + ⅟2 O2→ 2CuO
(air reactor)
2CuO → Cu2O + ⅟2 O2 (fuel Reactor)
Figure 42. Results of material and energy balance calculations for the conceptualized process.
The calculations were based on a coal feed flow rate of 1000 tonne/day of Illinois #6, with a net heating
value of 13,730 Btu/lb (or 31.987 MJ/kg determined on a maf basis (Table 8). The temperature of the air
and the fuel reactor was determined based on assumed equilibrium O2 concentrations, as outlined in the
design assumptions below.
Table 8. Properties of Illinois#6 coal.
The major steps for the combustion process considered were the devolatilization and subsequent volatile
oxidation of coal, char oxidation, and carbon burnout. The following design assumptions were taken into
consideration:

Coal was modeled as a chemical compound with an empirical formula of CaHbOc .

Ash was considered an inert and its effect on the energy balance was assumed to be the heating
value of coal on a moisture-free basis.

An exit O2 concentration was 2.1 mol%, which yielded a temperature of 916⁰C.

It is assumed that 40% of the coal burns in the volatile combustion section having a net heating
value of coal on a moisture and ash-free basis. The outlet gas consists of 10 mol% O2, which
yielded a temperature of 994⁰C.

90% of the carbon burns in the char combustion section with a net heating value of coal on a
moisture-free basis. The outlet gas consists of 10 mol% O2.

It was assumed that the remaining carbon burns in the final carbon burnout section with a net
heating value of coal on a moisture-free basis. The outlet gas consists of 10 mol% O2.
The results of preliminary material and energy balance calculations (Figure 42) emphasize the importance
of the oxygen carrier. The oxidation of the oxygen carrier in the air reactor is a significant contributor to
the processes’s heat output. An important outcome of the preliminary study is the need to heat the air
entering the air reactor. The option should not result in a significant loss of energy obtained from
oxidation of oxygen carrier. Attempts are also being made for determining reactor volumes by
incorporating appropriate chemical reaction kinetics for the air and conceptualized fuel reactor sections.
Task 16 - Carbon Sequestration
During this quarter, the Sequestration Team considered typical kinetic rates published in the literature and
variations in mineral compositions of ions in important sequestration reactions. Geo-chemical
simulations were carried out with the different initial mineral concentrations and brine chemistries using
Geo-Chemists Work Bench 7.0.3 (GWB). GWB has its own thermodynamic database and requires kinetic
data as input from the user. The rate equation is
Rate =
dni
  Ea
 KAmin exp 
dt
 RT

 Q
 1

  K eq 
where K is the rate constant (mol/cm2), Amin is the reactive surface area (cm2), Ea is the activation energy
(J/mol), R is the gas constant (J/kmol), T is absolute temperature (K), Q is the activity product and K eq is
the equilibrium constant. The kinetic rate constants used here are from the literature and are measured in
the laboratory. These can differ from the original weathering or precipitation rates of the minerals in the
field by several orders of magnitude. The equilibrium constants are calculated by the GWB from its
internal database by interpolation. Surface areas used are also from the literature. They can be measured
by experiments or estimated by geometric approximations as suggested by Knauss et al. (2005).
The kinetic data entered by the user plays a significant part in measuring the changes in the rock
chemistry and the corresponding changes in the brine chemistry. The rate at which the principal ions are
liberated into the brine depends on the rate of weathering of their corresponding minerals. Hence the
kinetic rate constants and the surface area are the governing parameters in geochemical simulations.
The kinetic data from the literature i.e., kinetic rate constants and surface area vary to a considerable
degree depending on the source. Hence a comparison between two sets of kinetic data from White et al.
(2005) and Knauss et al. (2005) is described below (Table 10 Table 11
Fig
ure 43 Figure 44).
Table 9. Initial basis for the simulation.
H20
Ca++
CO2
H+
Na+
ClSiO2 (aq)
Mg++
Al+++
K+
Fe++
Time
Temperature
0.03 kg solvent
3 mg/l
113.08 bar fugacity
pH 6
23032 molal
58525 molal
.4 mg/l
1 mg/l
172 mg/l
5 mg/l
24 mg/l
134 days
100 C
Table 10. Kinetic data from White et al. (2005).
Mineral
Kinetic rate constant (mol/cm2sec)
surface area(cm2/g)
Calcite
Quartz
Feldspar
Clinochlore
Dolomite
671
686
711
1130
635
3.16E-14
1.86E-16
1.60E-11
2.34E-16
4.17E-12
7.20E-12
7.10E-12
7.00E-12
6.90E-12
6.80E-12
6.70E-12
Time (Days)
0.03
0.02
0.01
Time (Days)
134
127
121
114
107
101
93.8
87.1
80.4
73.7
67
60.3
53.6
46.9
40.2
33.5
26.8
20.1
13.4
0
134
127
121
114
107
101
93.8
87.1
80.4
8
6
4
2
0
6.7
13
.4
20
.1
26
.8
33
.5
40
.2
46
.9
53
.6
60
.3
67
73
.7
80
.4
87
.1
93
.
10 8
0.5
10
7.
11 2
3.9
12
0.6
12
7.3
13
4
0.04
10
0
Concentration (mg/kg)
0.05
0
67
K
0.06
6.7
73.7
Time (Days)
Mg
Concentration (mg/kg)
60.3
53.6
46.9
40.2
33.5
26.8
20.1
13.4
0
6.60E-12
6.7
131.32
124.62
117.92
111.22
97.82
104.52
91.12
84.42
77.72
71.02
64.32
57.62
50.92
44.22
37.52
30.82
24.12
17.42
4.02
10.72
Concentration (mg/kg)
Al
1.63E-04
1.63E-04
1.63E-04
1.63E-04
1.63E-04
1.62E-04
1.62E-04
Time
Concentration (mg/kg)
Ca
Time (Days)
Figure 43. Change in the concentrations of principal ions Ca, Al, Mg, K based on White et al (2005).
Table 11. Kinetic data from Knauss et al. (2005).
Mineral
Kinetic rate constant
(mol/cm2sec)
Surface area(cm2/g)
Calcite
Quartz
Feldspar
Clinochlore
Dolomite
6.45E-11
1.25E-16
3.45E-10
2.50E-14
3.90E-12
671
686
711
1130
635
131
123
115
107
99.2
75
83.1
83.08
91.1
67
73.7
59
51.2
25
20
15
10
129.98
120.6
111.22
92.46
Time (Days)
101.84
54.94
45.8369…
38.4711…
32.7185…
27.6070…
22.0089…
16.3593…
10.7616…
0
5.11185…
5
0
131.32
123.28
115.24
107.20
99.16
91.12
83.08
75.04
67.00
58.96
51.20
43.16
37.56
32.72
28.41
23.62
18.78
13.94
9.15
4.30
Concentration (mg/kg)
K
0.06
0.05
0.04
0.03
0.02
0.01
0
0.00
43.2
Time (Days)
Mg
Concentration (mg/kg)
64.32
Time (Days)
37.6
32.7
28.4
133
125
117
109
101
92.5
84.4
76.4
68.3
60.3
52.5
44.5
38.5
33.5
29.2
24.4
19.6
14.7
9.95
5.11
1.58E-04
23.6
1.59E-04
18.8
1.60E-04
13.9
1.61E-04
9.15
1.62E-04
7.20E-12
7.00E-12
6.80E-12
6.60E-12
6.40E-12
6.20E-12
6.00E-12
5.80E-12
0
1.63E-04
4.3
Concentration (mg/kg)
Al
1.64E-04
0.27
Concentration (mg/kg)
Ca
Time (Days)
Figure 44. Change in the concentrations of principal ions Ca, Al, Mg, K based on Knauss et al. (2005).
As observed from the kinetic data from the two sources, calcite dissolution rate constant from Knauss
et al. is about 3 orders of magnitude greater than that from White et al. The dolomite dissolution rates are
similar. Hence the difference in the trend in Ca ion concentration can be attributed to the difference in the
kinetic parameters. The same can be said about Al and Mg where the kinetic parameters of their
corresponding minerals differ. K ion concentration follows the same trend but its concentration in brine
in both cases are relatively high because the K-feldspar dissociation rates are the highest and silicate
weathering reactions are the most sensitive to pH.
Ca
Al
concentration in ug/l
Concentration in ug/l
500
400
300
200
100
200
150
100
50
0
0
0
14
27
42
62
94
0
134
14
27
42
Time (days)
94
134
K
Mg
160
200
Concentration in mg/l
Concentration in ug/l
62
Time (days)
150
100
50
0
0
14
27
42
Time (days)
62
94
134
140
120
100
80
60
40
20
0
0
14
27
42
62
94
134
Time (days)
Figure 45. Change in the concentrations of principal ions Ca, Al, Mg, K based experimental
measurements.
The measured concentrations of the principal ions using inductively coupled plasma-mass spectrometry
(ICP-MS) are lower compared to those from the simulation (Figure 45). This may be attributed to two
factors.
1. The rates from the literature are at least 2-3 magnitudes higher than the measured rates.
2. The concentrations calculated when a mineral assemblage was reacted with brine and CO2 rather
than individual minerals. The role of competitive kinetics among the participating minerals,
especially minerals with common principal ions like calcite and dolomite, cannot be ruled out.
Task 17.0 – Atomistic Modeling of Oil Shale Kerogen and Asphaltenes
The goal of this project is to establish a representative three-dimensional model of the Green River oil
shale kerogen by performing energy minimizations of proposed two-dimensional Siskin structures and its
interaction with the inorganic matter of the rock for the purpose of identifying suitable means of isolating
kerogen from oil shales. The model is being evaluated by comparing the pairwise distribution function
(PDF) obtained from X-ray scattering experiments on a representative kerogen.
The Task 17.0 team has constructed a 20-unit kerogen model (20402 atoms) that approximately fits a
rectangular box to mimic the experimental density of this kerogen (0.98 g/cm3). The presence of “dead”
spaces around the corners of our box reduces the model’s density to 0.80 g/cm3 and introduces problems
when comparing the PDFs. To minimize this error, a volume correction is being added to address the void
spaces. By correcting the models for the shape and density, good agreement is observed between the
model and experimental density-corrected PDF (the G(r) representation).
The team is currently building other models with the same stoichiometry as the Siskin model but with
very different structural features in order to investigate the sensitivity of the PDF. They are also in the
early stages of studying the mineral-kerogen interaction by putting the kerogen model between two finite
slabs of inorganic mineral such as illite. Further plan includes performing energy minimization of the
mineral-kerogen model and studying the swelling properties of kerogen with different organic solvents.
Task 18.0 - Multiscale Thermal Processing (Pyrolysis) of Shale
The Task 18.0 team completed an experiment to measure the composition of products as they evolved
during the pyrolysis process using Thermal Gravimetric Analysis/Mass Spectrometry (TGA-MS). A
heating rate of 5°C/minute was used. The weight loss curve is shown in Figure 46. The shape of the
weight loss curve is essentially identical to those obtained by TGA alone. The derivative peak also occurs
near the same temperature. Features of the second weight loss peak are also consistent with previous
observations.
usin
g TGA-MS.
Figure 46. The weight loss curve and the derivative obtained.
The evolution of different components with time is recorded in Figure 47. Some of the prominent
molecular ions are shown in the figure. The majority of the components evolve between 400°C (80
minutes) and 450°C (90 minutes). There is no distinct water peak in the TGA signature; however, the
molecular ion for water (18) is seen over the entire temperature range. Some secondary decomposition to
light gases (methane, propane) is seen at high temperatures. This study provides insight into the
formation of hydrocarbon species from kerogen.
Figure 2: Evolution of different pyrolysis components as a function of time.
Figure 47. Evolution of different pyrolysis components as a function of time.
Task 19.0 – Pore Scale Analysis of Oil Sands/Oil Shale Pyrolysis by X-ray Micro CT
and LB Simulation
The main thrusts of Task 19.0 include 1) CT characterization of the pore network structure for selected oil
sand/oil shale resources, 2) LB simulation of flow through pore network structures to predict transport
properties, such as permeability, and 3) CT analysis of pore network structure during pyrolysis reactions
at different temperatures. The drill core samples of Mahogany oil shale and the coke products after
pyrolysis were provided by Professor M. Deo from Task 18.0 of this research program.
The pore structure and the connectivity of the pore space are important features which determine fluid
flow in oil shale during pyrolysis. In this regard, the X-ray micro/nano tomography (XMT/XNT)
technique is the best non-invasive, non-destructive method available today to characterize complex pore
structures. XMT/XNT allows the capture of the 3D shape and connectivity of the void space in oil shale
after pyrolysis. From image digitalization of the oil shale sample, the research team is able to obtain the
porous network structure that evolved during pyrolysis. Computer simulation can then be used to
calculate macro variables of the flow. The Lattice Boltzmann Model (LBM) is an emergent mathematical
technique that is able to handle the complex boundary conditions of flow in the porous structures of an
actual oil shale sample and to simulate the process in a reasonable amount of time.
Preliminary results are used to evaluate the utility of the 3D LBM in predicting saturated permeability for
oil shale during pyrolysis. First, the pore network structure of oil shale and the residual products after
pyrolysis are characterized using XNT with 50 nm resolution. Then, the reconstructed 3D image is used
for LBM flow simulation to calculate the expected permeability based on the porous structure of the
sample. Figure 48 illustrates the 3-D view of the LB simulation for saturated flow through the pore
spaces of residual products of oil shale after pyrolysis (sample MD-8). The left-hand side of Figure 48
shows the porous structure of the MD-8 sample. After removing the solid phases, the right-hand side of
Figure 48 shows the nature of the flow channels. The estimated permeability from LB simulation of oil
shale after pyrolysis is found to be about 0.00363 μm2 or 0.363 mD (millidarcy).
Figure 48. 3D views of LB simulated flow through the reconstructed CT image of the oil shale sample
Figure
3: 3D
views (MD-8).
of LB simulated
flowphase
through
the reconstructed
CT image solid
of thephase
oil shale
sample
after
pyrolysis
(Left) Solid
is white.
(Right) A transparent
reveals
flowafter
pyrolysis (MD-8). (Left) Solid phase is white. (Right)
A
transparent
solid
phase
reveals
flow
channels.
channels.
Task 20.0 – Basin-wide Characterization of Oil Shale Resource in Utah and Examination of In-situ
Production Models
A major inefficiency when producing oil from oil shale by pyrolysis alone is the enormous heat
requirement to decompose the kerogen into products. Also, the temperature of the rock near the heating
wells is extreme in order to provide a large enough temperature gradient for sufficiently fast heat transfer
through the reservoir. Another potential problem with pyrolysis alone is excessive coking in some areas
of the reservoir due to fast heating rates and high temperatures. In-situ combustion could be included in
an in-situ process strategy to supply heat to the reservoir, thus reducing the external heating requirement.
A pyrolysis/combustion combined process may be an attractive strategy to utilize the strengths of each
process alone. Pyrolysis is attractive because the reservoir can be heated without permeability, and
combustion is attractive because the heat is supplied by the reservoir itself. Using the rigorous material
and elemental balance stoichiometry and assuming component lumping, the Task 20.0 team simulated a
pyrolysis process followed by in-situ combustion. In the simulated process strategy, heat was supplied to
the reservoir approximately 900 days, and then the heaters were shut off and air was injected for coke
combustion.
Oil shale source rock is essentially impermeable, so air injection can potentially be a problem. While in
the past oil shale reservoirs were rubbalized to create permeability, most current in-situ processing
strategies do not include the rubbalization stage. Instead, the assumption is made that permeability is
created as the solid kerogen in the rock decomposes to fluid products. The CMG simulator STARS used
in this work has a model for permeability creation based on the initial permeability and fluid porosity and
a multiplier factor.
Simulation results are quite sensitive to changes in gas saturations and pressure. After investigation, it is
hypothesized that this sensitivity is due to rapid changes in gas saturations because gas flows more freely
than other components through the low permeability reservoir. When these changes due to flow are
coupled with gas saturation changes due to reaction, the time step in the simulation reduces significantly.
This sensitivity was overcome by maintaining a back pressure that matched the initial pressure of the
reservoir (1000 psi) initially, and then altering the back pressure to 900 psi at 900 days so that minimal
products are consumed during the combustion stage of the process.
For this hybrid simulation, the pyrolysis reactions were the same as those in the pyrolysis-only simulation
with the exception that the “rigorous” mass and elemental balanced stoichiometry was used. The heating
rate was reduced over time to avoid the extreme high temperatures near the heating wells that were seen
with pyrolysis-only simulations. The hybrid simulations include reactions for kerogen, heavy oil, light
oil, char, and coke combustion. Parameters for these combustion reactions were estimated based on
STARS template files and data. The product distribution from this simulation is shown in Figure 49.
Fig
methane) produced duriolysis-combustion process.
Figure 49. Total amount of products (heavy oil, light oil, non-methane gas and methane)
produced during the hybrid pyrolysis-combustion process.
There are some notable differences between this process and the base case pyrolysis-only process. First,
this process shows significantly less oil production but more gas production. This result could be due to
the process or to the more “rigorous” pyrolysis stoichiometry. The temperature histories of the blocks in
the hybrid process are somewhat more complex than those in the pyrolysis-only process. The
temperatures are not as extreme, although the temperature of the block nearest the heater still reaches a
maximum of about 1,750 R. Simulation results show that the combustion process distributes heat in the
reservoir more efficiently to blocks far from heaters (as expected since the heaters are stationary), and the
combustion front moves throughout the reservoir.
The energy loss/gain profile is shown in Figure 50. The Shell in Situ Conversion Process pilot-scale
project reported a 3:1 ratio for energy gained over energy supplied using resistive heating for producing
oil from oil shale. The simulation results from the pyrolysis-only base case gave a ratio of 4.8 units of
energy gained in oil and gas per unit of energy supplied by the heaters. This estimate does not include
any electricity generation inefficiency. The hybrid process that was simulated gave a ratio of 8.0 units of
energy gained in oil and gas per unit of energy supplied by the heaters. This increase in efficiency is
significant, though fewer barrels of oil are produced.
Figure 50. The energy loss/gain profile of the hybrid process.
Task 21.0 – Simulation, Validation, and Uncertainty Quantification of Oxy-Gas Combustion for
CO2 Capture
After attending several conferences this past quarter and interacting with various companies in the process
heater industry, the project team has determined that a focus on oxy-fuel flameless combustion represents
an opportunity for research work in a technology of great interest to industry. After the Industrial
Advisory Board meeting in June 2009, one of the board members approached the project team with an
oxy-gas flameless combustion application. While the dataset is of great interest, the application has some
complex features that may require extensive code modification before simulations can be performed. The
team is evaluating whether to pursue this application or a dataset from CANMET involving an oxy-gas
burner.
Work this quarter focused on the development of a reliable tool for creating reaction tables for oxy-fuel
combustion application. To create a reaction table, a detailed chemical kinetics model is run in a
simplified reactor such as an opposed jet diffusion flame (e.g. flamelets). The resulting species and
temperature information is then tabulated as a function of a few parameters (e.g. heat loss, mixture
fraction, etc.). Preliminary simulations of an oxy-gas burner indicated a problem existed with the reaction
table. A new tool has been created that uses a commercial software package, DARS, to create flamelets
tables. The information in these tables is then reformatted for use with an LES simulation tool. Work can
now proceed on a demonstration simulation of an oxy-gas burner.
Task 22.0 – Effect of Oil Shale Processing on Water Compositions
It is recognized that thousands of compounds are formed when kerogen is converted to shale oil. In the
last report, the Task 22.0 research team identified a number of polar/organic compounds formed using gas
chromatography/mass spectrometry. A few of the compounds and their water solubilities and sorption
coefficients at room temperature are shown in Table 12.
Table 12. Solubilities and sorption coefficients of some of the compounds in shale oil.
me of the compounds in shale
To explore the fate of a chemical dissolved in water, the research team solved the advective-dispersion
equation using COMSOL. The problem was solved with and without sorption and with various
hydrogeological controls. The hydraulic conductivity was 10 m/day. Some example solutions after 10
days are provided in Figure 51. In this figure, the concentration of a pollutant (e.g. benzene) in a subdomain source (continuous source) is compared. Solutions with and without sorption (Freundlich
isotherm with n=7) are shown in Figure 51. The pollution potential of this source is significant as
observed in the figure. With sorption, the plume is more spread out, while it is concentrated without the
sorption.
contaminant transport equation showing pollutant concentration profiles: (left) with a Freundlich sorption
Figure 51. COMSOL solutions of the contaminant transport equation.
Subtask 23.0 – Climate Change Legislation and Regulatory Gap Assessment
During the past quarter, the Task 23.0 team has continued to focus on establishing the framework and
factual foundation for assessing the substance of what regulations will be necessary to facilitate
commercial-scale CCS deployment. In this regard, we have made progress in three primary areas. First,
we have largely completed our compilation of the database identifying roughly 500 industry players
involved in CCS implementation and advocacy (including electric power companies, mining companies,
major oil and gas companies, researchers, and think tanks). Second, we have reviewed and analyzed
legislation spanning the last three Congresses, to build a comprehensive list of possible regulatory gaps.
We are now in the process of combining these two efforts to create a standardized survey we can use to
contact industry and other CCS players concerning their views on regulatory needs. Finally, we have
begun expanding our review of governmental work on CCS to examine what steps agencies have begun
taking, and how those steps fit with pending and prior legislation. And we continue to keenly monitor
Congress' and the Obama administration's efforts and signals on CCS, as the new administration begins
implementing its environmental policies.
Task 24.0 – Market Assessment of Heavy Oil, Oil Sands and Oil Shale Resources
The Task 24.0 team met to better define production scenarios for heavy oil, oil shale, and oil sands and
identify potential sources of data for the economic impact analysis modeling. Heavy oil economic data is
readily available due to decades of heavy oil production in California, but there will be some
modifications to make when applying this data to production from as yet undeveloped Alaskan heavy oil
fields. For oil shale development, three types of production are being researched: mining, modified in
situ and in situ. There are companies with pilot scale facilities for each of these technology types. We
realize that this type of information is proprietary, but we are focusing our efforts on combing available
documents and on consulting contacts in the industry for any potentially useful information. Some
members of the team are traveling to the Athabasca region of Canada in August 2009 to visit various oil
sands operations. One purpose of the trip is to better understand how Canadian oil sands economic data
might be applied to Utah, where the size of the resource, while substantial, is several orders of magnitude
less than that of the Canadian resource.
Michael Hogue, the team economist, envisions the following format for the economic impact analysis.
First, for each region-resource combination and for each of several plausible production scenarios,
estimate the total addition to regional levels of employment, personal income, and government revenue.
Consider potential development from unconventional resources for the following combinations of regions
and resources:
Uinta and Piceance Basin - oil shale
Utah (Uinta and Paradox basins) - oil sands
North Slope area of Alaska - heavy oil
Second, for each region, produce a number of plausible production scenarios. From models of these
processes, obtain a list of the required inputs (e.g. x units of steel, y units of rubber hosing, . . .) for each
year of two phases of the process: (1) construction, (2) operations and maintenance. Third, estimate
economic impacts by preparing the data obtained from the scenarios, preparing economic profiles for the
regions, preparing the regional accounts, applying the stimulus, and devising an effective manner of
presenting the results.
Subtask 24.1 - CO2 Emissions
This past quarter, the subtask 24.1 team expanded its literature data on greenhouse gas (GHG) emissions
to include data for oil sands and shale. This information provides a more complete picture of the
differences in emissions from various production methods such as the contrast between in situ and aboveground retorting of oil shale. In addition, the team contacted experts in oil sands and shale and gained
valuable published resources and a better understanding of which processes are in commercial use. In
addition, we are expecting a copy of a draft chapter from an upcoming book, Preliminary Estimates of
Greenhouse Gas from Shale Oil Production, Dr. Jerry Boak of the Center for Oil Shale Technology and
Research at the Colorado School of Mines.
Specifically, further literature was gathered on the Alberta Taciuk Procesor (ATP), an above-ground oil
shale retort (Brandt 2007). The ATP is claimed as one of the most promising above-ground retorts. As
presented in Figure 52, two primary cases, low and high, were based on energy consumption from large,
open-pit mines. The low case is modeled after a coal mine, and the high case is modeled after an oil
sands operation. The next two cases, small scale and low carbon emissions, are bounding cases
demonstrating potential efficiencies of the ATP process.
Figure 52. Emissions by process stage for converting Green River oil shale to liquid fuels with the ATP
(Brandt, 2007).
Figure 53 shows a comparison of GHG emissions estimates for production of synthetic crude from oil
shale using by various processes. Upstream emissions in Figure 54 include emissions from extraction,
processing, and upgrading to crude oil or syncrude, and the refining emissions are shown separately.
Final combustion emissions are excluded. The Oil Sands Exploration Company (OSEC) uses an ATP
process similar to the small-scale case in Figure 52. The Greenpeace values are for the retorting of
Australian Stuart oil shales in a Southern Pacific Petroleum retort. Emissions from the conversion of
Green River Shale are presented for the Shell in situ conversion process (ICP) and the ATP process.
These data present more realistic emissions range, about 12 to 20 gCeq/MJ, compared to older studies
where complete carbonate decomposition and combustion of organic material were assumed. For
example, Sundquist and Miller (1980) provide a CO2 emissions range of 14 to 57 gCeq/MJ.
Figure 55. Comparison of GHG emissions estimates for production of synthetic crude from oil shale
using the Shell in-situ conversion and Alberta Taciuk processes (Brandt 2007, 2008).
Subtask 24.2 – Policy Analysis of the Canadian Oil Sands Experience
During the past quarter, the subtask 24.2 team began gathering available literature relevant to
understanding of the economic context that facilitated development of the Canadian oil sands industry.
The subtask 24.2 team has continued research related to air quality issues associated with domestic oil
sands production and to the differences in water issues associated with Canadian and domestic oil sands
production. Additionally the subtask 24.2 team has prepared interview questions and scheduled visits with
various members of government, academia and industry in Edmonton and Calgary in preparation for a
tour of the oil sands in Canada scheduled for August 3-6, 2009.
Subtask 24.3 – Policy Analysis of Water Availability and Use Issues in the Context of Domestic Oil Shale
and Sands Development
During the past quarter the subtask 24.3 team focused on research in the following areas: regulation of
produced water; regulations applicable to impoundment construction; and regulation of groundwater.
Subtask 24.3 team efforts also involved outlining, drafting, and editing a paper for publication that
addresses interstate river allocation and Indian reserved rights. Preliminary findings were presented at the
Critical Intersections for Energy and Water Law conference in Calgary, Alberta.
Task 25.0 – Repository of Data, Information and Software
Subtask 25.1 – Addition of New Materials to the Repository
Approximately 200 additional full-text documents have been uploaded to the repository during the last
quarter, as well as approximately 80 abstract-only documents for which full-text permission could not be
obtained. ICSE’s computer professional has begun working with the Institute librarian to develop
enhanced document management queues within the DSpace platform. The Institute librarian is pursuing
ongoing communications with various publishers to seek full-text permission on documents remaining in
the workflow and abstracting documents where appropriate. Work has also begun on acquiring
permission to post various theses in the repository workflow.
Subtask 25.2 – Improvement of Map Server Interface to Repository
During the last quarter, efforts have focused on improving the ICSE interactive map by updating the map
as well as the web application that displays the map. This new web map will be deployed as a 'beta'
version in August 2009, and then is scheduled to replace the existing web map in October 2009.
The subtask 25.2 team also developed a new and improved prototype web mapping application including:
a new table of contents to include collapsible folders and improve the end user's browsing experience; an
updated layout of interface revamped to be more functional and appealing for the end user
; and
improved performance (speed) of the map. Michelle Kline at the Energy and Geoscience Institute at the
University of Utah further prepared for planned upgrading of our existing ArcIMS web mapping software
to ArcGIS
Server by taking a two day training course for the new software.
New GIS datasets have been obtained from multiple sources, and those datatsets were evaluated for
integrity and usefulness. The new datasets include water resources information, wilderness and sensitive
areas information, and
detailed oil shale study information. Higher resolution basemap data for the
interactive map was also obtained during this quarter.
Subtask 25.3 – Water Solutions for Future Unconventional Fuel Development
To address the first phase of this task, a graduate student revisited the previously compiled GIS datasets
and acquired and created new metadata. Metadata for 21 GIS datasets was provided to the interactive map
team. In addition, the student worked with Michelle Kline (interactive map developer) to answer
questions about data and identify new datasets to obtain. Next, new GIS datasets were obtained from the
USGS seamless data server to provide input to a related project to develop a water management model for
the White River. These datasets and their metadata will also be delivered to the interactive map developer.
Finally, as part of this first phase, oil shale reports and publications specifically related to water
management and water quality were obtained and delivered to the repository librarian.
The second phase of this task involves developing a geodatabase schema and geodatabase to support
water-energy management in the Uinta Basin. The geodatabase schema was prepared. Data will be added
as the next step.
Leveraged with this project is another project developing the White River water management model. The
model has been prepared in cooperation with the Utah State Division of Water Resources. The most
recent leveraged effort was the acquisition of a reservoir location dataset to add those components to the
water management model. In addition, two proposed reservoirs (one in Colorado and one in Utah) have
been incorporated into the model. The model is being applied to study the feasibility of both the Colorado
and Utah reservoirs being constructed and the implications of future uncertain climate variability. This
application is being used as the guide for the GIS dataset collection and geodatabase development
activities for this task.
MILESTONE STATUS
Critical path milestones. The project team completed the construction of the laboratory-scale CLC system
as part of Task 14. Because the drop-tube furnace has required unexpected repairs, the planned
completion date for the preliminary drop-tube experiments will be delayed until September 2009.
Table 13 summarizes the critical-path milestones.
Milestone
Project management plan
Oxy-coal combustion LES
tool
Validation data for oxy-fuel
combustor
Construction of the labscale CLC system
Drop-tube experiments
Comparison of LES results
and validation data
Assessment report
Reservoir models
Expanded repository
Policy reports
Table 13. Milestone status.
Planned Completion Date Actual Completion
Date
October 2009
October 2009
March 2010
Notes
March 2010
June 2009
August 2009
March 2010
June 2009
Expected Sept. 2009
March 2010
March 2010
March 2010
March 2010
In the Oil Sands and Shale Program, four tasks had milestones that in the past two quarters (none were
reported last quarter). For Task 18.0, the two milestones were the development of experimental systems
and the obtaining of analytical measurements of oils produced under various conditions. These tasks are
now complete and have been briefly reported in quarterly reports submitted for the first three quarters of
this project. For Task 20.0, the two milestones scheduled for completion are an evaluation of in situ
processes and the basin-wide characterization of Green River oil shale. The first milestone was completed
and reported in two previous quarterly reports. The basin-wide characterization is 60% complete. The
research team has joined forces with UGS to complete this task in a timely fashion. For task 21.0, the two
milestones to be completed by this quarter were the identification of available experimental datasets and
the demonstration of an oxy-natural gas burner simulation. After the Industrial Advisory Board meeting
in June 2009, one of the board members approached the project team with an oxy-gas flameless
combustion application. While the dataset is of great interest, the application has some complex features
that may require extensive code modification before simulations can be performed. The team is still
evaluating whether to pursue this application or a dataset from CANMET involving an oxy-gas burner. A
decision will be made in the next quarter. Progress toward achieving a demonstration simulation of an
oxy-natural gas burner was hampered by the lack of a reliable reaction table to represent the chemistry
occurring in an oxy-fuel flame. The problem has been solved and a simulation will be performed in the
next quarter. For Task 22.0, an experimental design for studying the effect of oil shale processing on
water compositions was scheduled for completion at the end of the last quarter. This task 80% complete
and will be reported next quarter.
ACCOMPLISHMENTS
During this quarter, two University of Utah students began participating in the DOE NETL research
experience this summer, and the CASE advisory board met on June 23 – 24th.
PROBLEMS OR DELAYS
A graduate student is needed to perform some of the work for Task 21.0. It is hoped that this
open position will be filled when the new academic year begins (third quarter of 2009).
REFERENCES
A. Brandt, Converting Green River oil shale to liquid fuels with the Alberta Taciuk Processor: energy
inputs and greenhouse gas emissions, University of California Berkeley (2007).
A. Brandt, Converting Oil Shale to Liquid Fuels: Energy Inputs and Greenhouse Gas Emissions of the
Shell in Situ Conversion Process, Environ. Sci. Technol., 2008, 42 (19), pp 7489–7495.
J.H. Chen J.C. Sutherland, P.J. Smith. A quantitative method for a priori evaluation of combustion
reaction models. Combustion Theory and Modeling, 11:287–303, 2007.
Fletcher, T. H., A. R. Kerstein, R. J. Pugmire and D. M. Grant, "Chemical percolation model for
devolatilization. 2. Temperature and heating rate effects on product yields," Energy & Fuels, 4(1), 5460 (1990).
Fletcher, T. H., A. R. Kerstein, R. J. Pugmire, M. S. Solum and D. M. Grant, "Chemical percolation
model for devolatilization. 3. Direct use of carbon-13 NMR data to predict effects of coal type,"
Energy & Fuels, 6(4), 414-31 (1992).
J.H. Friedman, “Multivariate Adaptive Regression Splines,” The Annals of Statistics, vol. 19, 1991, pp. 167.
Grant, D. M., R. J. Pugmire, T. H. Fletcher and A. R. Kerstein, "Chemical model of coal devolatilization
using percolation lattice statistics," Energy & Fuels, 3(2), 175-86 (1989).
K. G. Knauss, James W. Johnson, Carl I. Steefel., 2005, Evaluation of the impact of CO2 , co-contaminant
gas, aqueous fluid and reservoir rock interactions on the geologic sequestration of CO2 . Chemical
geology 217 (339-350)
A. Parente, J.C. Sutherland. Combustion modeling using principal component analysis. In Elsevier,
editor, Proceedings of the Combustion Institute, volume 32, pages 1563–1570. the combustion
institute, 2009.
E. T. Sundquist, G.A. Miller, Oil Shales and Carbon Dioxide, Science, 1980, 740-741.
L. Tognotti P.J. Smith A. Parente, J. Sutherland. Identification of lowdimensional manifolds in turbulent
flames. In Elsevier, editor, Proceedings of the Combustion Institute, volume 32, pages 1579–1586.
the combustion institute, Elsevier, 2009.
S.P. White, R.G. Allis, J. Moore, T. Chidsey, C. Morgan,W. Gwynn, M. Adams, 2005, Simulation of
reactive transport of injected CO2 on the Colorado Plateau, Utah, USA , Chemical geology 217 (387405).
Zeng, D., S. Hu, A. N. Sayre, and T. H. Fletcher, “Mechanisms of Coal Secondary Pyrolysis and Soot
Formation,” presented at the 34th International Technical Conference on Clean Coal & Fuel Systems,
Clearwater, FL (May 31-June 4, 2009).
RECENT AND UPCOMING PRESENTATIONS/PUBLICATIONS
Jake Bauman and Milind Deo, “Interaction Between Reactivity and Flow in the In-situ Production of Oil
from Oil Shale,” 29th Oil Shale Symposium, Colorado School of Mines, Golden, CO, October 19-23,
2009.
R. Baraki, G. Konya, E. Eyring “Investigation of the reaction kinetics of oxygen carriers in chemical
looping combustion." 26th Annual International Pittsburgh Coal Conference, September 20 - 23,
2009.
E.G. Eddings, A. Sanchez, L. Wang, F. Mondragon “Pollutant Formation During Oxy-Coal Combustion”,
Presentation at the Clearwater Coal Conference, June 1 - 8, 2008, Clearwater, FL.
Burian, S. Jones, E., Hong, A., Goel, R., Li, L., Cha, Z., Dudley-Murphy, E., and Nash, G. Oil Shale
Development in Western U.S.: Water Resources Challenges and Solutions. American Water
Resources Association Spring Specialty Conference, Anchorage, Alaska, May 4-6, 2009.
Chung-kan Huang, Jake Bauman, and Milind Deo, “Modeling of the In-situ Production of Oil from Oil
Shale,” book chapter in review, to be published by ACS in 2009.
C. L. Lin, J.D. Miller, C. H. Hsieh, P. Tiwari and M. D. Deo, “Pore-scale Analysis of Pyrolyzed Oil Shale
Cores,” 29th Oil Shale Symposium, Colorado School of Mines, Golden, CO, October 19-23, 2009.
Suhui Li, Kevin J. Whitty, Investigation of Coal Char-Slag Transition during Oxidation: Effect of
Temperature and Residual Carbon, Energy & Fuels 2009, 23, 1998–2005.
Will Morris, Dunxi Yu and Jost O. L. Wendt, Comparison of size segregated composition of coal ash
aerosol under air-fired and oxy-fired conditions, Presented at 11th International Congress on
Combustion By-Products and their Health Effects, Research Triangle Park, NC 27711, June 1-3,
2009.
J. Pedel, J. Thornock, P.J. Smith, “ Modeling Coal Devolatization in Large Eddy Simulation of Oxy-Coal
Combustion, Presentation will be made at the Annual Clearwater Coal Conference, June 1 - 8, 2008,
Clearwater, FL.
N. K. Punati and J. C. Sutherland, “Application of an Eulerian One Dimensional Turbulence model to
Simulation of Turbulent Jets,” 6th US National Combustion Meeting, Ann Arbor, MI, May 2009.
N. K. Punati, J. C. Sutherland, A. R. Kerstein, E. R. Hawkes, and J. H. Chen, Western States Section of
the Combustion Institute Meeting, San Diego, CA, October 2009.
C. Reid; J. Thornock; P. Smith, Temporally and Spatially Resolved Calculations of a Reacting Coal Jet
Using Large Eddy Simulations (LES) and Direct Quadrature Method of Moments (DQMOM), 26th
Annual International Pittsburgh Coal Conference, September 20 - 23, 2009.
John Ruple and Robert Keiter, “Unresolved Water Allocations in the Uinta and Piceance Basins.”
Presentation will be given at “Critical Intersections for Energy & Water Law: New Challenges and
Opportunities,” Calgary, Alberta, Canada, May 20-21, 2009.
A. Sanchez, Fanor Mondragon, Eric G. Eddings "Fuel-Nitrogen Evolution During Fluidized Bed OxyCoal Combustion," presentation at the 20th International Conference on Fluidized Bed Combustion,
Xi'an, China, May 18-20, 2009.
Philip Smith; J. Thrornock; J. Pedel; C. Reid, Large Eddy Simulation of Near-Nozzle Region in an
Entrained Coal Gasifier, 26th Annual International Pittsburgh Coal Conference, September 20 - 23,
2009.
A.F. Sarofim, P.J. Smith, R.J. Pugmire, K. E. Kelly, Near and longer term solutions to carbon capture and
sequestration, 26th Annual International Pittsburgh Coal Conference, September 20 - 23, 2009.
R. Shurtz; T. Fletcher; R. Pugmire, the Use of 2,6-Dimethylnapthalene as a surrogate for studying soot
formation from coal tar, 26th Annual International Pittsburgh Coal Conference, September 20 - 23,
2009.
Randy Shurtz; Tom Fletcher, Pyrolysis and gasification of a sub-bituminous coal at high heating rates.
26th Annual International Pittsburgh Coal Conference, September 20 - 23, 2009.
Jennifer Spinti, Milind Deo, and Philip Smith, “Technology Advances in Oil Shale Producion: Reducing
the Environmental Footprint,” book chapter in review, to be published by ACS in 2009.
J. C. Sutherland and A. Parente, “Modeling Combustion using Principal Component Analysis: Dealing
with Source Terms,” 6th US National Combustion Meeting, Ann Arbor, MI, May 2009.
Pankaj Tiwari, Kyeongseok Oh and Milind Deo, “Oil Shale Pyrolysis: Characterization and
Compositional Analyses,” 29th Oil Shale Symposium, Colorado School of Mines, Golden, CO,
October 19-23, 2009.
Kirsten Uchitel, Robert Keiter, and John Ruple, “Environmental and Policy Issues Relevant to Oil Shale
Development Under the Energy Policy Act of 2005,” book chapter in review, to be published by ACS
in 2009.
Dunxi Yu; Will Morris; Jost Wendt, Partitioning of Alkali and Alkaline Earth Metals in Fine Ash from
Oxy-coal Combustion by Individual Particle Analysis, 26th Annual International Pittsburgh Coal
Conference, September 20 - 23, 2009.
Jingwei Zhang, Kerry Kelly, Eric G. Eddings, and Jost O.L. Wendt, Oxy-coal Combustion: effects of
PO2 on Coal Jet Stability in O2/CO2 Environments, International Flame Research Foundation, 16th
IFRF Members’ Conference Boston, MA, June 8th -10th, 2009.
J. Zhang; K. Kelly; E. Eddings; J.O.L. Wendt, Effects of O2 Partial Pressure on Flame Stability in oxycoal combustion, 26th Annual International Pittsburgh Coal Conference, September 20 - 23, 2009.
Download