You are on page 1of 9

2592 Ind. Eng. Chem. Res.

2000, 39, 2592-2600

Predictive Heat Model for Australian Oil Shale Drying and


Retorting

Adam J. Berkovich,† John H. Levy,‡ Brent R. Young,*,§ and S. James Schmidt|


Department of Chemistry, Materials and Forensic Science, University of Technology, Sydney, Sydney,
Australia, Division of Energy Technology, Commonwealth Scientific and Industrial Research Organisation,
Lucas Heights Science and Technology Centre, Menai, Australia, Department of Chemical and Petroleum
Engineering, University of Calgary, Calgary, Alberta, Canada, and Southern Pacific Petroleum (Development)
Pty Ltd., Brisbane, Queensland, Australia

The exploitation of Australia’s oil shale reserves has been the subject of much research over
several decades. The extraction of oil from shale is usually a thermal process and therefore oil
shale processing is dominated by heat-transfer and thermal reactions. To fully develop any oil
shale processing technology, knowledge of an oil shale’s thermodynamic properties is desirable.
This paper presents a novel approach for the determination of an oil shale’s thermodynamic
properties and subsequent development of a predictive heat model. This approach involves
experimental determination of the thermodynamic properties of an oil shale’s individual
components, in particular, the unique kerogen and clay minerals in Australian oil shale. The
experimental data was then used to develop a predictive heat model for Australian oil shale
based upon its kerogen and mineral, composition, and concentrations. The predictive heat model
allows investigation of heat capacity, enthalpy, and total enthalpy for Australian oil shale during
the process temperatures of drying and retorting.

Introduction nantly hydrocarbon vapor that is collected and trans-


ferred to the oil recovery section of the process. The
Australia has a vast reserve of oil shale deposited organic pyrolysis reactions are accompanied by a num-
throughout its northeastern regions as well as smaller ber of other thermal reactions due to the mineral
deposits located in its southeast. The current estimate components in the oil shale. These reactions include the
of Australia’s in situ shale oil reserve is approximately partial decomposition of pyrite, the dehydroxylation of
23 × 109 barrels, equivalent to 20 times Australia’s the clay minerals, and the decomposition of some
current crude oil reserves.1 In recent years there has carbonate minerals.3
been renewed interest in an Australian oil shale indus- The retorted oil shale is then transferred to the
try. combustion stage where the residual carbon from the
In Australia, the proposed processing technique is the oil shale is burnt at approximately 750 °C. The com-
AOSTRA-Taciuk continuous flow retort system, which busted oil shale is split into two streams. One stream
is based on rotary kiln technology.2 The processor is recycled through the retort to provide the necessary
consists of four compartments: drying/preheating of heat for retorting. The other stream is fed to a cooling
feed oil shale, retorting of dried oil shale, combustion zone where it transfers heat to the oil shale being fed
of retorted oil shale, and heat recovery from combusted into the preheat/drying stage.
oil shale.
The continued development of the AOSTRA-Taciuk
Oil shale is fed into the drying/preheating stage of processor for oil shale processing requires appropriate
the processor where it is heated to ≈250 °C. At the process models. In a process dominated by heat-transfer
preheat temperature, residual surface moisture as well and thermal reactions, knowledge of oil shale thermo-
as any water associated with the crystal structure of dynamic properties is important.
relevant minerals in the oil shale are liberated. The
preheated oil shale (at 250 °C) then passes into the The published examination of oil shale thermal
retort stage where it is mixed with recycled combusted characteristics related to processing dates back some
oil shale (at 750 °C). The mixing of the preheated and 80 years.4,5 This work examined the reaction heats and
combusted oil shale results in a retort temperature of heating requirements of whole Colorado oil shale and
≈500 °C. In the retort, kerogen is pyrolyzed to predomi- found that reaction heats and heating requirements
were dependent on the oil shale composition. Colorado
oil shale has also been the subject of much work on its
* To whom correspondence should be addressed. Depart- thermal characteristics using differential scanning cal-
ment of Chemical and Petroleum Engineering, University of orimetry and thermogravimetry.6-10 Comparisons be-
Calgary, 2500 University Drive N.W., Calgary, Alberta, Canada tween pyrolysis heats of Colorado and Devonian oil
T2N 1N4. Phone: (403) 220-8751. Fax: (403) 282-3945.
shale have also been made.7 The pyrolysis heats for
E-mail: byoung@ucalgary.ca.
† University of Technology, Sydney. Colorado oil shale were found to be lower because of the
‡ Commonwealth Scientific and Industrial Research Orga- differences in mineralogy between Colorado and Devo-
nization. nian oil shales, in particular, the influence of pyrite
§ University of Calgary. found in the Colorado shales. Other work has shown a
| Southern Pacific Petroleum (Development) Pty Ltd. correlation between total measured enthalpy and oil
10.1021/ie990942g CCC: $19.00 © 2000 American Chemical Society
Published on Web 06/13/2000
Ind. Eng. Chem. Res., Vol. 39, No. 7, 2000 2593

shale grade; that is, the higher the oil yield for the shale, differential scanning calorimetry (DSC)51 for thermo-
the higher the total measured enthalpy.8-10 dynamic measurements. Modulated DSC is a relatively
The application of thermal analysis to coal and oil new calorimetric technique that is capable of quantita-
shale has been reviewed.11 Thermogravimetry and dif- tively measuring heat capacity and enthalpy changes
ferential scanning calorimetry have been shown to be in a single experiment. This particular feature is unique
useful tools in determining the oil yield potential, to this technique and has important consequences for
reaction heats, and effect of minerals in different oil oil shale analysis.
shales.12-15 Drop calorimetry has also been used to Siroquant52,53 is a software package for quantitative
measure the heat of retorting for Green River oil shale analysis of minerals by X-ray diffraction (XRD). It
up to 773 K.16,17 Thermal analysis has also been coupled contains a specialized package for the analysis of clays.54
with spectroscopy to investigate the retorting behavior Siroquant has previously been applied successfully to
of carbonaceous oil shales from the Nagoorin and Australian oil shales55 and was further developed in this
Condor deposits in Australia.21 study to determine mineral concentrations in oil shales,
The thermal properties of Eastern and Green River kerogens, and isolated clay mixtures. The thermal,
oil shales from the United States have also been spectroscopic, and literature data of the individual
determined using a combination of thermodynamic data components were collated to develop a predictive heat
from the literature and the average concentrations of model of Stuart oil shale. The model allows prediction
oil shale components, to sum the heat effects for an of the thermodynamic properties of an oil shale based
average shale.18,19 The heat capacities of kerogen and on its mineral and organic composition. The model can
kerogen char were estimated using heat capacity data be directly applied or incorporated into oil shale process
of model compounds, such as anthracene, polymeric models and can be used to investigate the sensitivity of
organic materials, and graphite, and adjusting the heat thermodynamic properties related to an oil shale’s
capacities according to the hydrogen and carbon con- composition. The predictive heat model was designed
tent.18,19 to be interactive, allowing the user to select any tem-
Australian oil shales have a more complex mineralogy perature range in the drying and retorting process
than U.S. oil shales. For example, Australian oil shales temperature range.
from Stuart/Rundle are a mixture of kerogen, clays, The model can be used in the process design, optimi-
carbonate minerals, silica minerals, and pyrite. The zation, and control of industrial oil shale drying, retort-
thermal decomposition, pyrolysis, and dehydroxylation ing, and combustion plant. In process design, reliable
of these oil shale components overlap in the temperature estimates of the heat capacities and reaction enthalpies
range from 673 to 823 K (400-550 °C).27,3 As a result, of raw, dried, and pyrolyzed oil shales are required for
the overall mass loss, enthalpy, and heat capacity the determination of heat balances and the subsequent
change for oil shale is a combination of all components. mechanical design of process equipment. The model can
be directly incorporated into process models for process
Previous experimental work in oil shale thermal optimization by using the model to investigate the
characterization discussed above has treated oil shale sensitivity of a process related to an oil shale’s composi-
as a single material. Consequently, thermal character- tion. In a similar manner the model is useful in process
ization measurements were made on whole oil shale and control as it may be used to fine-tune the process
the data then mathematically treated (usually simul- operating conditions based on periodic laboratory oil
taneous linear regression) to determine the contribu- shale composition measurements.
tions of the individual components. This method is
usually satisfactory for oil shales that have a simple
mineralogy, such as the Green River oil shales, which Experimental Section
have a carbonate mineral base and only one dominant Modulated differential scanning calorimetric (MDSC)
clay mineral. However, this method cannot be applied analyses were performed on a TA Instruments MDSC
satisfactorily to Australian oil shales, in this case Stuart 2920 instrument, with a working temperature range of
oil shales, as the diverse mineralogy of Stuart oil shale 148 to 998 K. Samples were analyzed in aluminum pans
makes measurement of thermodynamic properties dif- with aluminum lids in an atmosphere of dry nitrogen
ficult. Each component in the oil shale has a different flowing at 120 mL min-1. Nominal sample mass was 5
concentration, different heat capacity, and different mg, which was measured to 1 µg with a microbalance.
reaction heats, and the temperature range of thermal Samples were equilibrated at 273 K for 5 min and then
reactions overlap each other.28 heated at 5 K min-1 to 873 K. The modulation ampli-
This paper reports on a new experimental approach tude was (1 K with a period of 60 s. Temperature and
to oil shale thermal characterization, which involves enthalpy were calibrated using In, Sn, and Zn stan-
treating oil shale as a mixture of different components dards. Heat capacity was calibrated with a 99.97%
that have individual thermal characteristics. In par- sapphire standard.
ticular, the unique components of Australian oil shale, A TA Instruments SDT 2960 instrument was used for
kerogen, smectite, kaolinite, and illite, are of the most thermogravimetry. The samples were analyzed in open
interest. By chemical and physical separation of these platinum pans at a heating rate of 5 K min-1 from
unique components from whole oil shale, the thermo- ambient to 873 K in an atmosphere of dry nitrogen
dynamics of these individual components can be deter- flowing at 120 mL min-1.
mined experimentally. The other more common mineral Two types of oil shale samples were selected for this
components found in Stuart oil shale, such as calcite, study, one set for kerogen isolation and the other set
quartz, siderite, and feldspar, have well-known ther- for clay mixture isolations. Samples with a high kerogen
modynamic properties. For these components, thermo- concentration were selected for kerogen isolation. Both
dynamic data were obtained from databases in the brown and carbonaceous kerogens were isolated. The
literature.29-50 high kerogen concentration required minimum chemical
This study also employed the use of modulated treatment, as those minerals not affected by the chemi-
2594 Ind. Eng. Chem. Res., Vol. 39, No. 7, 2000

Table 1. Combined Results for Siroquant and Chemical Table 2. Siroquant Analysis Results for Pyrolyzed
Analyses for Isolated Stuart Kerogens Kerogens (Including Organic Matter)

component concentration brown carbonaceous sample PK1 PK5


(% of kerogen) K1 K2 K3 K4 K5 pyrite (% of pyrolyzed kerogen) 10.0 11.4
pyrite 4.7 5.9 3.4 7.9 6.6 pyrrhotite (% of pyrolyzed kerogen) 9.6 7.3
anatase 0.7 0.4 0.7 0.3 0.4 anatase (% of pyrolyzed kerogen) 3.0 1.3
miscellaneous minerals 0.8 0.7 1.6 0.7 0.6 miscellaneous minerals (% of pyrolyzed kerogen) 6.7 3.3
organic matter 93.8 93.0 94.3 91.1 92.4 organic matter (% of pyrolyzed kerogen) 70.7 76.7

Table 3. Siroquant Results for Clay Minerals Isolated


cal treatment would be concentrated to the least extent. from Stuart Oil Shale
The clay minerals in Stuart oil shale usually exist as smectite kaolinite illite quartz albite cristobalite
mixtures of kaolinite, smectite, and illite in different sample (%) (%) (%) (%) (%) (%)
ratios. The three clay types, kaolinite, smectite, and
C1 31.0 19.9 28.1 12.0 2.9 6.0
illite, in Stuart oil shale cannot be separated from each C2 7.5 18.9 25.8 41.4 5.7 0.6
other. Therefore, a range of oil shales was selected for C3 11.6 38.2 20.1 17.8 4.0 8.3
clay isolation to ensure a greater diversity in the C4 0.4 11.7 20.7 58.6 8.6 0.0
composition for adequate characterization of the indi- C5 43.2 14.3 9.0 25.7 7.4 0.4
vidual clays. C6 41.3 17.4 8.8 20.7 9.4 2.5
Kerogen was isolated from oil shale using a HF/BF3 C7 48.9 10.1 28.7 8.9 2.4 1.0
C8 38.7 21.7 11.6 14.9 5.8 7.3
maceration technique56 where carbonate-type minerals C9 30.8 22.0 13.7 25.0 5.8 2.8
are destroyed with hydrofluoric acid and boron trifluo- C10 10.9 28.9 21.7 22.6 4.4 11.5
ride, generated from the reaction of hydrofluoric and C11 73.7 10.6 12.5 2.0 1.2 0
boric acids. Pyrolyzed samples of kerogen and oil shale C12 32.4 15.1 24.6 20.1 5.2 2.5
were also prepared by heating samples to 520 °C in an C13 23.6 20.2 15.5 31.0 6.4 3.3
atmosphere of flowing nitrogen and then quenching the C14 37.7 21.3 9.0 25.8 5.8 0.5
C15 42.6 21.9 8.8 21.1 5.2 0.5
sample at room temperature while the nitrogen atmo- C16 34.6 17.7 9.8 31.6 6.3 0.0
sphere was maintained.
The clay minerals were isolated from oil shale using Table 4. Thermogravimetric Mass Loss at 773 K for
buffered acetic acid and hydrogen peroxide.57 Acetic acid Isolated Kerogens
destroyed carbonate minerals and hydrogen peroxide kerogen initial mass final mass mass loss
destroyed the kerogen. The result was a mixture of clay sample (mg) (mg) (%)
minerals and silica-type minerals such as quartz and K1 10.21 2.22 78.2
feldspar. These heavier silica minerals were separated K2 10.42 2.30 78.0
from the lighter clay mixture by centrifugation. The clay K3 10.49 2.10 80.0
mixtures were comprised of kaolinite, smectite, and K4 19.74 5.23 73.5
illite, in varying ratios. K5 19.13 5.99 68.7
XRD patterns were collected using a Siemens Kri- Table 5. Dehydroxylation Mass Losses for Clays Isolated
stalloflex X-ray generator equipped with two powder from Stuart Oil Shale
cameras with Bragg-Brentano geometry. A Philips
dehydroxylation dehydroxylation
PW2276/20 X-ray tube was used at a power of 30 mA sample mass loss (%) sample mass loss (%)
and 45 kV to produce cobalt X-rays. Samples were
mounted in aluminum sample holders of dimensions 44- C1 10.9 C7 17.1
C2 14.2 C8 15.9
mm long (in the plane of the X-rays), 12.5-mm wide, C3 3.0 C9 17.6
and 1.7-mm deep. XRD patterns were collected from C4 8.6 C10 15.6
3.00° to 90.00°, at intervals of 0.04° 2θ using a count C5 5.4 C11 13.4
time of 10 s/interval. C6 5.5 C12 19.4

Results and Discussion correct measured thermodynamic data on isolated Stu-


art kerogen and pyrolyzed kerogen.
Initially, 66 oil shale samples from the Stuart deposit The X-ray patterns of the isolated clays also showed
were acquired and their XRD patterns measured. The low concentration of included minerals: quartz, cristo-
patterns were analyzed using Siroquant to determine balite, and albite. These silica-based minerals were
mineral concentrations in the oil shales. Samples to be unaffected by the chemical treatments and could not be
included in the study were selected using these data. completely separated from the clays by centrifugation.
Five samples were selected for kerogen isolation and 16 However, the concentrations of these minerals are
samples were selected for clay mineral isolation. relatively low and corrections for their effects could be
The purity of kerogen and clay minerals isolated from made in subsequent analyses. Siroquant results for the
Stuart oil shale was assessed using XRD and chemical isolated clays also show a diverse range of clay concen-
analyses. Siroquant analysis of the XRD patterns trations in each sample (Table 3).
determined the concentrations of any mineral inclusions Results of thermogravimetric analyses performed on
in the isolated materials. XRD patterns of isolated the isolated kerogen and clay samples are summarized
brown and carbonaceous kerogen showed peaks arising in Tables 4 and 5.
from anatase and pyrite: minerals which are inert to The raw heat capacity of pyrolyzed brown and car-
the chemical treatments used to isolate the kerogen. The bonaceous isolated kerogen was measured using modu-
combined results of the Siroquant and chemical analy- lated DSC over the temperature range 273-773 K (0-
ses are shown in Table 1, while Table 2 presents 500 °C). The heat capacity of pure pyrolyzed kerogen
Siroquant analysis results for pyrolyzed Stuart kero- (PK) was calculated by subtracting the heat capacities
gens. These concentrations were subsequently used to of the mineral matter components from the raw heat
Ind. Eng. Chem. Res., Vol. 39, No. 7, 2000 2595

Table 6. Heat Capacity Equation Coefficients (J g-1 K-1)


for Pyrolyzed Kerogens
sample
coefficients PK1 PK5
a 2.503 -7.490 × 10-2
b -1.781 × 10-2 7.535 × 10-3
c 6.690 × 10-5 -1.889 × 10-5
d -9.100 × 10-8 3.035 × 10-8
e 4.355 × 10-11 -1.906 × 10-11
R2 1.0000 0.99877
std. dev. 2.792 × 10-10 0.008

capacity data. Heat capacity data for R-alumina were


used for the miscellaneous minerals, which in any case
represent only a small proportion of the material. These
corrected data were fitted to a fourth-order polynomial
equation of the type Cp ) a + bT + cT2 + dT3 + eT4.
The coefficients are given in Table 6. Figure 1. Heat capacity curves for Stuart kerogen K1 (solid) and
The heat capacity of kerogen isolated from Stuart oil K5 (dash) between 273 and 773 K.
shale was measured using modulated DSC. However,
at temperatures above 623 K, pyrolysis of the kerogen
with the accompanying loss of mass began to influence ity data had been measured for the kerogen up to the
the heat capacity data and a different approach had to temperature prior to pyrolysis beginning at about 623
be used. The heat capacity of isolated kerogens was K. Heat capacity data for pyrolyzed kerogen have also
measured up to 673 K (the beginning of pyrolysis). been accurately measured across the full temperature
These data were corrected for moisture present in the range, to 773 K. It was reasonable to assume that the
samples, measured by TG/DTA, so that the heat capac- change in heat capacity between kerogen and pyrolyzed
ity determinations were reported on a dry basis. The kerogen would be proportional to the change in mass of
measured heat capacities were also corrected for min- the kerogen. This change was therefore modeled using
eral inclusions (Table 1). The adjusted data were fitted the thermogravimetric data obtained under the same
to the fourth-order polynomial Cp ) a + bT + cT2 + heating conditions for the same sample (Table 4). Thus,
dT3 + eT4 to give the coefficients in Table 7. The the heat capacity value of kerogen at 623 K was taken
temperature range used for fitting was 273-623 K as the initial value. The heat capacity value for the same
because the onset of pyrolysis began to influence the kerogen pyrolyzed at 773 K (in J gpyrolyzed kerogen-1) was
results in the temperature region above 623 K (350 °C). converted to J gkerogen-1 on the basis of mass loss
The initial attempts to measure heat capacities of observed by thermogravimetry (Table 4), corrected for
kerogens with the modulated DSC resulted in very low the included minerals and taken as the final value. The
values of heat capacity data after pyrolysis. These difference between the initial and final values was then
values were often negative and quite different from the multiplied by the fraction of mass loss occurring at 1 K
values for the heat capacity of the pyrolyzed kerogen intervals between 623 and 773 K, and each value was
as measured above, after allowance was made for the subtracted from the initial value to construct the model
change in mass. Further investigation showed that the curve. These data were poorly fitted with a fourth-order
modulated DSC heat flow curve base line after pyrolysis polynomial equation so an eighth-order polynomial
was wrongly positioned because the instrument could equation was used over this temperature range to give
not cope with the large mass losses of 69%-80% that the coefficients as shown in Table 8. The full heat
occur during kerogen pyrolysis.28 A possible solution was capacity curves for a kerogen sample K1 (brown) and
to dilute the kerogens with R-alumina and measure the K5 (carbonaceous) are shown in Figure 1.
heat capacities of the mixtures, such that the alumina Initial attempts to measure the enthalpy of pyrolysis
would help retain sufficient material in the sample pan for isolated Stuart kerogens using modulated DSC also
during and after pyrolysis to give good coupling of the gave unexpectedly low values, as found for the heat
sample with the pan and hence the thermocouple capacity measurements. These low results were at-
sensor. This was attempted, but because of the loss of tributed to the very large mass loss in the sample during
pyrolysis products, it was not possible to encapsulate pyrolysis and to the very endothermic nature of the
the samples properly, resulting in heat capacity data pyrolysis reaction.
that were still inconsistent with pyrolyzed kerogen data. To solve this problem, the kerogens were diluted with
The following method was employed to obtain con- alumina before the enthalpies of pyrolysis were mea-
sistent data. As stated previously, accurate heat capac- sured. Alumina is inert over this temperature range and

Table 7. Heat Capacity Equation Coefficients (J g-1 K-1) from 273 to 623 K for Isolated Kerogens
kerogen sample
coefficients K1 K2 K3 K4 K5
a -5.185 -2.674 -7.623 -4.087 -5.260
b 5.781 × 10-2 2.356 × 10-2 7.334 × 10-2 4.069 × 10-2 5.109 × 10-2
c -1.809 × 10-4 -3.816 × 10-5 -2.346 × 10-4 -1.245 × 10-4 -1.408 × 10-4
d 2.632 × 10-7 2.950 × 10-8 3.361 × 10-7 1.753 × 10-7 1.887 × 10-7
e -1.423 × 10-10 -1.008 × 10-11 -1.777 × 10-11 -9.161 × 10-11 -9.840 × 10-11
R2 0.99418 0.99875 0.99858 0.99875 0.99779
std. dev. 0.027 0.014 0.012 0.011 0.017
2596 Ind. Eng. Chem. Res., Vol. 39, No. 7, 2000

Table 8. Heat Capacity Equation Coefficients (J g-1 K-1) During pyrolysis, part of the pyrite is converted to
from 623 to 773 K for Isolated Kerogens pyrrhotite, assumed to be by reduction with hydrogen
sample generated in the organic cracking process.18,60,61 The ∆H
coefficients K1 K5 for this reaction was calculated as 290.4 J g-1, at the
peak temperature of 748 K (475 °C), using Cp and ∆H
a -6.765 × 105 -4.948 × 105
b 5.360 × 103 4.002 × 103
data for pyrite and pyrrhotite. The analysis of the
c -1.671 × 101 -1.289 × 101 pyrolyzed kerogen (Siroquant data) shows that the
d 2.427 × 10-2 2.008 × 10-2 reaction does not proceed to completion, being only
e -1.149 × 10-5 -1.270 × 10-5 55.4% complete at 793 K (520 °C). When this same
f -1.134 × 10-8 -3.690 × 10-9 factor is used for kerogen from sample 65219K, ∆Hpyrite
g 1.680 × 10-11 9.444 × 10-12
h -6.762 × 10-15 -4.007 × 10-15
for 65219K was calculated as 160.9 J gpyrite analyzed-1.
i 5.326 × 10-19 2.599 × 10-19 Then, when the concentration of pyrite in the kerogen
R2 0.9995 0.9999 is taken into account (5.9%), ∆Hpyrite for 65219K2 was
std. dev. 0.021 0.009 7.56 J gkerogen-1. In addition, the dilution effect of the
pyrite, and for the anatase and miscellaneous minerals
Table 9. Enthalpies of Pyrolysis of Stuart K1 Kerogen
Diluted in r-Alumina
that were assumed inert during pyrolysis, also has to
be accounted for. The result is a ∆H of 364 J g-1 for the
concentration concentration enthalpy of pyrolysis of Stuart brown kerogen.
of kerogen (%) ∆H (J g-1) of kerogen (%) ∆H (J g-1)
Although heat capacities were measured on the clay
30 27.20 70 188.1 mixtures isolated from Stuart oil shale, the heat capaci-
50 121.8 80 267.0
60 179.0 ties of the individual clays were not calculated from the
experimental data. The errors associated with indi-
Table 10. Regression Line Parameters for Stuart 65219K vidual heat capacities (deconvoluted from heat capaci-
Kerogen in r-Alumina ties measured on the mixtures) were large compared to
parameter standard error the differences among the heat capacities of the indi-
slope 4.56 0.41 vidual clays. This occurred because the constitutions of
intercept -108.0 16.6 the various clays are all so similar: the chemical
R2 0.974 formulas of smectite, kaolinite, and illite all consist of
number of observations 5 aluminum, silica, and oxygen (that is, an alumino-
degrees of freedom 3 silicate-base structure) with minor contributions by
other substituents, such as potassium, calcium, or iron.
provided bulk to the sample, thereby reducing the mass
Therefore, the value of the heat capacity of the clay is
change occurring in the sample during the experiment
while sample contact with the sensor was maintained. influenced more by the alumino-silicate-base structure
Diluted kerogen samples were prepared in concentra- (which is similar for all clays) rather than by the
tions of 30%, 50%, 60%, 70%, 80% w/w kerogen. Heat substituents. This was evident from the literature heat
flow curves for the diluted samples were measured by capacity data for smectite, kaolinite, and illite,36,45,49
modulated DSC. The pyrolysis peak from the heat flow where capacity values for these clays are similar.
curve was integrated to calculate enthalpy of each Consequently, heat capacity data from the literature
diluted kerogen sample and these data are presented were used for the individual clays.
in Table 9. Heats of dehydroxylation were measured using modu-
A linear regression between measured enthalpy and lated DSC for mixtures of smectite, kaolinite, and illite
kerogen concentration was performed and the regres- isolated from Stuart oil shale. These results are sum-
sion parameters are shown in Table 10. From these marized in Table 11 which also shows the dehydroxy-
coefficients, a raw enthalpy value for a kerogen concen- lation enthalpies corrected for the mineral inclusion
tration of 100% was calculated as 348.2 J g-1. This value concentrations (Table 3) and the enthalpy change as-
must then be corrected for mineral inclusions in the sociated with the R to β transition of quartz. The
kerogen (Table 1). normalized concentrations of smectite, kaolinite and
Table 11. Dehydroxylation Enthalpies for Clays Isolated from Stuart Oil Shale
dehydroxylation x1, x2, x3, y,
peak normalized normalized normalized ∆Hdehyrox’n ∆Hdehyrox’n
temperature concentration concentration concentration (measured) (corrected)
sample (K) smectite kaolinite illite (J g-1) (J g-1)
C1 726.77 0.392 0.252 0.356 201.5 253.2
C2 738.03 0.144 0.362 0.494 139.3 258.4
C3 728.70 0.166 0.546 0.288 231.2 328.1
C4 766.98 0.012 0.357 0.631 120.5 348.8
C5 759.62 0.650 0.215 0.135 218.8 325.0
C6 761.20 0.612 0.258 0.130 247.4 363.7
C7 732.86 0.558 0.115 0.327 220.4 250.3
C8 734.89 0.538 0.301 0.161 178.4 245.6
C9 729.19 0.463 0.331 0.206 187.5 278.3
C10 718.89 0.177 0.470 0.353 233.7 376.2
C11 743.77 0.761 0.110 0.129 225.6 233.0
C12 729.32 0.449 0.209 0.341 219.8 300.7
C13 726.61 0.398 0.341 0.261 249.9 416.0
C14 788.18 0.554 0.313 0.132 240.0 349.3
C15 773.70 0.581 0.299 0.120 247.9 335.5
C16 732.19 0.557 0.285 0.158 249.1 395.8
Ind. Eng. Chem. Res., Vol. 39, No. 7, 2000 2597

Table 12. Simultaneous Linear Regression Results for Table 14. Sources of Error in the Predictive Heat Model
Isolated Clays
error source total error (%)
constant 0
literature heat capacity data 8
std error of y 56.4
literature enthalpy data 5
R2 0.2230
XRD/Siroquant 3
number of observations 19
MDSC heat capacity measurements 1-3
degrees of freedom 16
MDSC enthalpy measurements <1
x coefficients 265.3 535.2 146.0
MDSC temperature measurements <0.5
error of coefficients 42.8 105.2 92.4
polynomial fitting <1
Table 13. Dehydroxylation Enthalpies for Smectite,
Kaolinite, and Illite This value is approximately equal to the latent heat of
clay ∆Hdehydroxylation (J g-1) error (J g-1) % error (J g-1) vaporization for water, 2258 J g-1 at 110 °C.67
The data measured for kerogen and clay minerals
smectite 265 (43 16
kaolinite 535 (105 20
isolated from Stuart oil shale were combined with
illite 146 (92 63 literature data for the other mineral components in
Stuart oil shale and used to develop a predictive heat
illite used for the regression analysis described below model. The literature usually revealed several sources
are also shown in Table 11. of data for a particular mineral component. In such
The dehydroxylation enthalpies for the individual cases, an average data set was calculated for each
clays, smectite, kaolinite, and illite, were calculated component and these data were refitted to the same
using a simultaneous linear regression analysis of the fourth-order polynomial as the experimental data.
measured experimental data. The normalized concen- Primarily, the model allows prediction of an oil shale’s
trations of smectite, illite, and kaolinite were assigned thermochemical properties based upon its mineral
to regression ranges x1, x2, and x3, respectively. The composition. The model was written in a spreadsheet
corrected enthalpies for each sample from each experi- environment and has been designed for use by process
ment were assigned the regression range, y. A simul- engineers and others involved in oil shale processing;
taneous linear regression was performed to fit values therefore, the model is interactive and has different
of x1, x2, and x3 to y to calculate coefficients of x (the levels of complexity depending upon the user needs.
regression line was forced through the origin). The The model takes experimental- and literature-sourced
coefficients of x were equal to the dehydroxylation thermodynamic data for each oil shale component,
enthalpy for each individual clay. The results of the namely, heat capacity and reaction enthalpy, and
simultaneous linear regression analysis and subsequent combines them with the concentration of each compo-
dehydroxylation enthalpy values are shown in Tables nent determined from Siroquant to calculate a total
12 and 13. thermodynamic value for a given oil shale. To use the
Inspection of the regression data shows relatively high model, the user simply inputs concentration values for
errors for the fitted coefficients and subsequent dehy- each oil shale component and a temperature range. The
droxylation enthalpies. This is mirrored in the value of user is able to input a temperature range between 0 and
R2. There are a number of possible reasons for this. 500 °C (this covers the process temperature range of
First, errors have been introduced by the difficulty in drying and retorting). The model then calculates values
measuring the base line on the high-temperature side for heat capacity, sensible heat (integrated heat capac-
of the peak, resulting in low values for the illite, as it ity), and reaction enthalpy for the specified concentra-
was possible that part of this peak was beyond the upper tion of each oil shale component. The model then sums
temperature limit of the calorimeter. Indeed, the en- these thermodynamic data and calculates a total en-
thalpy for illite has the highest relative error. Second, thalpy (the sum of sensible heat and reaction enthalpies)
the error calculated for simultaneous linear regression for the whole oil shale. The interactive nature of the
analyses is dependent on the number of degrees of model allows the user not only to predict heat data for
freedom in the data set. It is possible that if more known oil shale feedstocks but also to investigate heat
samples were measured, there would be a reduction in data for different feedstock scenarios.
the calculated error. Finally, it is quite probable that Errors arise from the heat capacity and enthalpy data
the individual clays within the clay mixtures are not from literature sources, error in the concentration of
identical, and this leads to significant variation in the each oil shale component from Siroquant analyses, and
enthalpies, which are reported in the errors. For ex- the error from polynomial fitting of experimental and
ample, it is difficult to believe that all the smectite clays averaged literature data. Table 14 lists error sources
have identical formulas given that the degree of sub- and the approximate error associated with those sources.
stitution may vary within the deposit given that mate- The largest source of error in the database model is
rial was laid down at different geological times. Even the heat capacity data from the literature. Although
kaolinite, which has a relatively uniform chemical most heat capacity data for the minerals in oil shale
formula, has been shown to exhibit different enthalpies had a quoted accuracy of <1%, others had a quoted
of dehydroxylation dependent upon the crystallinity of accuracy of between 1% and 2%, and the total error for
the sample.62-66 Given these considerations, however, all components whose literature dataare used in the
this is the first time that the dehydroxylation enthalpies database model is 8%. The enthalpy data from the
have been measured for the clays in oil shales and literature showed a greater accuracy and the total error
represent a major step forward in our ability to be able for all oil shale mineral components was 5%. The errors
to calculate the thermochemical properties of oil shales. quoted for the modulated DSC measurements in this
The enthalpy associated with dehydration of oil shale study were determined from analysis of high-purity
was also determined using modulated DSC. An average (99.9%) mineral standards for enthalpy and heat capac-
value of 2196 J g-1 at a peak temperature of 110 °C for ity and high-purity (99.98%) metal standards melting
the loss of water was found for 12 oil shale samples. temperatures for temperature measurements.
2598 Ind. Eng. Chem. Res., Vol. 39, No. 7, 2000

Table 15. Comparison of Integrated Heats Determined


from Predictive Heat Model and Modulated DSC,
298-623 K (25-350 °C)
integrated integrated
heat-predictive heat-modulated
heat model DSC
sample (J g-1) (J g-1)
O1 536 515
O2 509 492
O3 348 356
O4 480 450
O5 456 420

The predictive heat model was validated by compar-


ing the total enthalpy calculated by the model with an
experimentally determined total enthalpy using modu-
lated DSC. Five oil shales from the sample suite, O1,
O2, O3, O4 and O5, were selected for comparison
because these samples showed suitable differences in Figure 2. Oil shale grade versus total enthalpy.
their mineralogy, which ensured that the selection was
representative of oil shale from the Stuart deposit. The heat capacity from modulated DSC was fitted to
The comparison was performed in the following way. a fourth-order polynomial and the total enthalpy cal-
The concentration of the mineral components in each culated. Table 16, shows the heat capacity coefficients
oil shale was entered into the database model and the for retorted O2 shale. The total enthalpy for O2 was
total enthalpy calculated for the temperature range determined to be 551 J g-1, compared to the database
298-623 K (25-350 °C). This temperature range was model calculation of 516 J g-1 for retorted O2 oil shale.
selected so that the mass loss interferences described Again, excellent agreement is obtained between experi-
above would not affect the integrated heat determina- mental and model results.
tion. The integrated heat was determined by measuring As is apparent from the above, the predictive heat
the heat capacity of the oil shales for the same temper- model cannot be validated against experimental data
ature range above using modulated DSC. The modu- over the full range of process temperatures (298-773
lated DSC data was then plotted versus temperature K) for whole oil shale. This is where the predictive heat
and fitted to a fourth-order polynomial, Cp ) a + bT + model supersedes experimental techniques. The data-
cT2 + dT3 + eT4. This expression was then integrated base model was developed using critically reviewed
and the integrated heat calculated. Table 15 shows the literature data as well as thermodynamic data mea-
comparison between integrated heats calculated from sured on the unique components isolated from Stuart
the predictive heat model and the integrated heats oil shale. The complex nature of Australian oil shale
determined from modulated DSC experiments. mineralogy was the reason that thermodynamic data
The above results show good agreement between the could not be determined on whole oil shale directly and
database model and experimental results, with the therefore required this approach. That is, a suitable
predicted values being within 5% of the experimental predictive heat model was developed so that the thermal
values, which is well within the experimental error. characteristics of whole oil shale can be explored over
The database model can also be used to calculate total the full range of processing temperatures.
enthalpy of retorted oil shale. Oil shale sample O2 was It has been shown8,18 that a linear relationship exists
used as an example. Estimates were made of the between U.S. oil shale grade (in liters of oil per ton of
concentrations of components in a retorted oil shale on shale at 0% moisture (LTOM)) and total enthalpy. The
the basis of mass loss (TG) data. The heat capacities of predictive heat model was used to calculate the total
pyrolyzed kerogen and dehydroxylated clay minerals enthalpy for the oil shale samples in the sample suite.
measured by modulated DSC were incorporated into the The total enthalpy was then plotted versus LTOM and
database model so that the total enthalpy of retorted displayed in Figure 2.
oil shale could be calculated. The concentrations of Inspection of Figure 2 shows at best an approximate
decomposition products of minerals in the original oil linear relationship with some substantial deviations. In
shale were calculated from the mole ratios of the a similar attempt27 at correlating total enthalpy versus
decomposition reactions of each component. oil shale grade for a series of Australian oil shales from
A sample of retorted oil shale was prepared from a different deposits, no clear correlation was found. The
sample of O2 oil shale in a manner similar to the reason for the differences in correlating oil shale grade
preparation of pyrolyzed kerogen. The retorted shale with total enthalpy can be attributed to differences in
was then analyzed by modulated DSC to determine the mineralogy of the oil shales being compared.
heat capacity for the temperature range 298-773 K The linear correlation found previously8,10,18 was for
(25-500 °C). Further validation of the model was Green River and Eastern oil shales from the United
performed by comparing the total enthalpy calculated States. Both of these oil shales have relatively simple
by the database model and that determined by modu- mineralogies compared to those of Australian oil shales.
lated DSC for retorted oil shale. The Green River oil shales have a simple predominantly

Table 16. Heat Capacity Coefficients from Modulated DSC of Retorted O2 Oil Shale for the Temperature Range 298-773
K (25-500 °C)
sample a b c d e
O2 0.2350 0.002 123 1.0255 × 10-6 -4.708 × 10-9 2.958 × 10-12
Ind. Eng. Chem. Res., Vol. 39, No. 7, 2000 2599

carbonate mineral base, of calcite and dolomite, which (4) McKee, R. H.; Lyder, E. E. The Thermal Decomposition of
do not decompose at retorting temperatures. Smectite Oil Shales. II Determination of the Heat of Reaction Involved in
has also been found in Green River oil shales, though Their Thermal Decomposition. Ind. Eng. Chem. 1921, 13, 678.
(5) Sohns, H. W.; Mitchell, L. E.; Cox, R. J.; Barnett, W. I.;
only in very small amounts (<10%). In essence, these Murphy, W. I. R. Heat Requirements for Retorting Oil Shale. Ind.
minerals simply act as a diluent for kerogen that leads Eng. Chem. 1951, 43, 33.
to a linear relationship between enthalpy and grade. (6) Cook, E. W. Thermal Analysis of Oil Shales. Q. Colo. Sch.
Indeed, the results presented in this study for pyrolysis Mines 1970, 65, 33.
enthalpies of kerogen diluted in alumina showed an (7) Wen, C. S.; Yen, T. F. A Comparison between the Properties
almost perfect linear relationship. However, the oil of Devonian Shale and Green River Oil Shale via Thermal
Analysis. Therm. Hydrocarbon Chem. 1979, 20, 343.
shales studied here have a diverse range of carbonate (8) Rajeshwar, K.; Nottenburg, R. N.; DuBow, J. B. Reviews
minerals and clays that have a tremendous effect on thermophysical Properties of Oil Shales. J. Mater. Sci. 1979, 14,
the heat requirements during processing (total en- 2025.
thalpy), depending upon their respective concentrations. (9) Jones, D. B.; Rajeshwar, K.; DuBow, J. B. Specific Heats of
These effects have also been observed for Rundle oil Colorado Oil Shales. A Differential Scanning Calorimetry Study.
shale,16 where it was found that a significant amount Ind. Eng. Chem. Prod. Res. Dev. 1980, 19, 125.
(10) Rajeshwar, K.; Jones, D. B.; DuBow, J. B. Characterisation
of extra heat was required for the retorting of Rundle of Oil Shales by Differential Scanning Calorimetry. Anal. Chem.
shale compared to that of Green River oil shale. The 1981, 53, 121.
Rundle shale examined in that study is akin to Stuart (11) Warne, S. St. J.; Dubrawski, J. V. Applications of DTA and
oil shale because it is from a shale deposit that is DSC to Coal and Oil Shale Evaluation. J. Therm. Anal. 1989, 35,
contiguous with Stuart. 219.
(12) Earnest, C. M. Thermogravimetry of Selected American
The nature of Australian oil shales requires a more and Australian Oil Shales in Inert Atmospheres. Thermochim.
complex approach to predicting their thermal charac- Acta 1982, 58, 271.
teristics rather than simple correlations that have been (13) Rajeshwar, K.; Rosenvold, R. J.; DuBow, J. B. Thermo-
used for U.S. oil shales. This only increases the calibre gravimetric Assay of Oil Shale. Thermochim. Acta 1983, 66, 373.
of the predictive heat model developed here as a tool (14) Levy, M.; Kramer, R. Comparative TGA and DSC Studies
for investigating the thermal characteristics of Austra- of Oil Shales. Thermochim. Acta 1988, 134, 327.
(15) Corino, G. L.; Turnbull, A. G. Calorimetric Studies of Oil
lian oil shales.
Shales and Shale Products. Proceedings Second Australian Work-
shop on Oil Shale, St. Lucia, Australia, Dec 6 and 7; 1984; pp 97-
Conclusions 102.
(16) Mraw, S. C.; Keweshan, C. F. New Calorimetric Method
The continued development of oil shale processing for Vaporisation Processes at High Temperatures: Liquid Frac-
technologies and the re-emerging Australian oil shale tions and Raw Oil Shale. Ind. Eng. Chem. Fundam. 1985, 24, 269.
industry have provided the impetus for accurate deter- (17) Mraw, S. C.; Keweshan, C. F. Calorimetric Determination
of the Heat of Retorting Oil Shales to 773K. Fuel 1986, 65, 54.
mination of oil shale thermodynamic properties. The (18) Camp, D. W. Oil Shale Heat Capacity Relations and Heats
complexity of Australian oil shale mineralogy warranted of Pyrolysis and Dehydration. Proceedings 20th Oil Shale Sym-
a new approach for the determination of these proper- posium, Golden, CO, 1987.
ties. The new approach involved the chemical and (19) Camp, D. W. Enthalpy Relations for Eastern Oil Shale.
physical separation of the unique components of Aus- Proceedings Eastern Oil Shale Symposium, Lexington, KY, 1987.
tralian oil shale and then thermal analysis of these (20) Jeong, K. M.; Patzer, J. F., II. Indigenous Mineral Matter
Effects in Pyrolysis of Green River Oil Shale. Geochem. Chem. Oil
materials to measure heat capacity, reaction enthalpy, Shale 1983, 30, 529.
and sensible heats. The isolated components and their (21) Stuart, W. I.; Levy, J. H. Thermal Properties of Carbon-
original oil shales were also analyzed by X-ray diffrac- aceous Oil Shales from Nagoorin and Condor Deposits. Fuel 1987,
tion and Siroquant to determine mineral concentrations. 66, 493.
The experimental thermodynamic data, literature- (22) Warne, S. St. J.; French, D. H. The Application of Simul-
sourced data, and Siroquant data were then combined taneous DTA and TG to Some Aspects of Oil Shale Mineralogy.
Thermochim. Acta 1984, 76, 179.
to develop a predictive heat model that allows heat (23) French, D. H.; Warne, S. St. J. Calcite Peak Anomalies in
capacity, enthalpy, and total enthalpy to be predicted the DTA of Oil Shales Determined in Flowing Carbon Dioxide
during the processes of drying and retorting, on the Compared to Nitrogen. Thermochim. Acta 1985, 84, 197.
basis of an oil shale’s composition. (24) Levy, J. H. Effect of Water Vapour Pressure on the
Dehydration and Dehydroxylation of Kaolinite and Smectite
Isolated from Some Australian Tertiary Oil Shales. Energy Fuels
Acknowledgment 1990, 4, 146.
(25) Hurst, H. J.; Levy, J. H.; Patterson, J. H. Siderite
The authors gratefully acknowledge the support of Decomposition in Retorting Atmospheres. Fuel 1993, 72, 885.
Southern Pacific Petroleum (Development) Pty Ltd. and (26) Patterson, J. H.; Hurst, H. J.; Levy, J. H. Relevance of
the Australian Research Council for support during this Carbonate Minerals in the Processing of Australian Tertiary Oil
research project. Shales. Fuel 1991, 70, 1252.
(27) Berkovich, A. J.; Young, B. R.; Ray, A.; Schmidt, S. J. The
Effect of Minerals on Retorting Enthalpies of Some Australian
Literature Cited Tertiary Oil Shales. Fuel 1998, 77, 987.
(28) Berkovich, A. J.; Young, B. R.; Schmidt, S. J.; Ray, A.
(1) Gannon, A. J.; Wright, B. C. Progress in Continuing Oil Thermal Characterisation of Australian Oil Shales. J. Therm.
Shale Research. In Proceedings of the Fifth Australian Workshop Anal. 1997, 49, 737.
on Oil Shale; CSIRO Division of Fuel Technology: Sydney, 1989; (29) Barin, I.; Knacke, O.; Kubaschewski, O. Thermodynamic
pp 3-8. Properties of Inorganic Substances; Springer-Verlag: NewYork,
(2) Taciuk, W.; Turner, L. R. Development Status of Australian 1973, supplement, 1977.
Oil Shale Processing Utilising the Taciuk Processor. Fuel 1988, (30) Barin, I. Thermochemical Data of Pure Substances; VCH:
67, 1405. Germany, 1989; Part I and II.
(3) Patterson, J. H.; Hurst, H. J.; Levy, J. H.; Killingley, J. S. (31) Barron, T. H. K.; Berg, W. T.; Morrison, J. A. On the Heat
Mineral Reactions in the Processing of Tertiary Oil Shales. Fuel Capacity of Magnesium Oxide. Proc. R. Soc. London 1959, 250A,
1990, 69, 1119. 70.
2600 Ind. Eng. Chem. Res., Vol. 39, No. 7, 2000

(32) Chopelas, A. Thermal Expansion, Heat Capacity and (50) Watanabe, H. Thermal Constants for Ni, NiO, MgO and
Entropy of MgO at Mantle Pressures. Phys. Chem. Miner. 1990, CoO at Low Temperatures. Thermochim. Acta 1993, 218, 365.
17, 142. (51) Reading, M.; Luget, A.; Wilson, R. Modulated Differential
(33) Gronvold, F.; Westrum, E. F., Jr. Heat Capacities of Iron Scanning Calorimetry. Thermochim. Acta 1994, 238, 295.
Disulfides. Thermodynamics of Marcasite from 5 to 700K, Pyrite (52) Taylor, J. C. Computer Programs for Standardless Quan-
from 300 to 780K, and the Transformation of Marcasite to Pyrite. titative Analysis of Minerals Using the Full Powder Diffraction
J. Chem. Thermodyn. 1976, 8, 1039. Profile. Powder Diffr. 1991, 4, 2.
(34) Gronvold, F.; Stolen, S.; Svendsen, S. R. Heat Capacity of (53) Taylor, J. C.; Clapp, R. A. New Features and Advanced
R-Quartz from 298.15 to 847.3K, and of β-quartz from 847.3 to Applications of SIROQUANT: A Personal Computer XRD Full
1000KsTransition Behaviour and Re-evaluation of the Thermo- Profile Quantitative Analysis Software Package. Adv. X-ray Anal.
dynamic Properties. Thermochim. Acta 1989, 139, 225. 1992, 35, 49.
(35) Helgeson, H. C. Summary and Critique of the Thermo- (54) Taylor, J. C.; Matulis, C. E. 1994 A New Method for
dynamic Properties of the Rock-Forming Minerals. Am. J. Sci. Rietveld Clay Analysis. Part 1. Use of a Universal Measured
1978, 278A, 225. Standard Profile for Rietveld Quantification of Montmorillonite.
(36) Hemingway, B. S.; Robie, R. A.; Kittrick, J. A. Revised Powder Diffr. 1994, 9, 119-123.
Values for the Gibbs Free Energy of Formation of [Al(OH)4aq], (55) Mandile, A. J.; Hutton, A. C. Quantitative X-ray Diffraction
Diaspore, Boehmite, Bayerite at 298.15K and 1 bar, the Thermo- Analysis of Mineral and Organic Phases in Organic-Rich Rocks.
dynamic Properties of Kaolinite to 800K and 1 bar, and the Heats Int. J. Coal Geol. 1995, 28, 51.
of Solution of Several Gibbsite Samples. Geochim. Cosmochim. (56) Robl, T. L.; Davis, B. H. Comparison of the HF-HCl and
Acta 1978, 42, 1533. HF-BF3 Maceration Techniques and the Chemistry of Resultant
(37) Hemingway, B. S.; Krupa, K. M.; Robie, R. A. Heat Organic Concentrates. Org. Geochem. 1993, 20, 249.
Capacities of the Alkali Feldspars between 350 and 1000K from (57) Jackson, M. L. Soil Chemical AnalysissAdvanced Course,
Differential Scanning Calorimetry, the Thermodynamic Functions A Manual of Methods Useful for Instruction and Research in Soil
of the Alkali Feldspars from 298.15 to 1400K, and the Reaction Chemistry, Physical Chemistry of Soils, Soil Fertility, and Soil
Quartz + Jadeite ) Analbite. Am. Miner. 1981, 66, 1202. Genesis, 2nd ed.; University of Wisconsin: Madison, WI, 1973.
(38) Hemingway, B. S. Quartz: Heat Capacities from 340 to (58) Khorasani, G. K. Characterization of Organic Matter in
1000 K and Revised Values for the Thermodynamic Properties. Oil Shales: Application of X-ray Diffraction Technique and
Am. Miner. 1987, 72, 273. Electron Microscopy. Proceedings Second Australian Workshop on
(39) Holland, T. J. B.; Powell, R. An Enlarged and Updated Oil Shale, St. Lucia, Australia, Dec 6 and 7, 1984; pp 62-66.
Internally Consistent Thermodynamic Dataset with Uncertainties (59) Lee, G. S. H.; Berkovich, A. J.; Levy, J. H.; Young, B. R.;
and Correlations: the System K2O-Na2O-CaO-MgO-MnO- Wilson, M. A. Reduction of Molecular Motion of Polymethylene in
FeO-Fe2O3-Al2O3-TiO2-SiO2-C-H2-O2. J. Metamorph. Geol. Oil Shales by Mineral Matter. Energy Fuels 1998, 12, 262.
1990, 8, 89. (60) Howatson, J.; Clark, J. A.; Poulson, R. E. Inorganic Sulfur
(40) JANAF Thermochemical Tables; Stull, D.R, Chase, M. W., Species in Retorted Oil Shale. 17th Oil Shale Symposium, Golden,
Curnutt, J. L., Downey, R. A., McDonald, A. N., Syverud, E. A., CO, 1984; Colorado School of Mines Press: Golden, CO, 1984; pp
Valenzuela et al. U.S. National Bureau of Standards, U.S. Govern- 376-383.
ment Printing Office: Washington, DC, 1971; 1974-1982. (61) Williamson, D. L.; Melchior, D. C.; Wildeman, T. R.
(41) Jacobs, G. K.; Kerrick, D. M.; Krupka, K. M. The High Changes in Iron Minerals during Oil-Shale Retorting. 13th Oil
Temperature Heat Capacity of Natural Calcite (CaCO3). Phys. Shale Symposium, Golden, CO, 1980; Colorado School of Mines
Chem. Miner. 1981, 7, 55. Press: Golden, CO, 1980; pp 337-350.
(42) Krupka, K. M.; Hemingway, B. S.; Robie, R. A.; Kerrick, (62) Nicholson, P. S.; Fulrath, R. M. Differential Thermal
D. M. High-Temperature Heat Capacities and Derived Thermo- Calorimetric Estimation of Kaolinite-Metakaolinte Endothermic
dynamic Properties of Anthophyllite, Diopside, Dolomite, Ensta- Enthalpy. J. Am. Ceram. Soc. 1970, 53.
tite, Bronzite, Talc, Tremolite, and Wollastonite. Am. Miner. 1985, (63) Davies, T. W.; Hooper, R. M. Structural Changes in
70, 261. Kaolinite Caused by Rapid Dehydroxylation. J. Mater. Sci. Lett.
(43) Pankratz, L. B. Thermodynamic Properties of Elements and 1985, 4 (1), 83.
Oxides; Bulletin 672; U.S. Bureau of Mines, Supt. of Docs., (64) Murat, M.; Chbihi, M.; El, M.; Mathurin, D. Enthalpie de
Washington, DC, 1982. Dissolution de Differentes Kaolinites et Metakaolinites dans
(44) Pisarczyk, K. Manganese Compounds. In Kirk-Othmer l’Acide Fluorhydrique Influence des Carateristiques Cristallo-
Encyclopedia of Chemical Technology, 4th ed.; Wiley: New York, Chemiques. Ind. Ceram. 1987, 822, 799.
1995; Vol. 15, p 1003. (65) Warne, S. St. J.; Dubrawski, J. V. Potential for Coal
(45) Ransom, B.; Helgeson, H. C. Estimation of the Standard Calorific Value Corrections, Dependent on the Carbonate Mineral
Molal Heat Capacities, Entropies, and Volumes of 2:1 Clay Type Present. J. Therm. Anal. 1987, 33 (2), 35.
Minerals. Geochim. Cosmochim. Acta 1994, 58 (21), 4537. (66) Meinhold, R. H.; Salvador, S.; Davies, T. W.; Slade, R. C.
(46) Robie, R. A.; Haselton, H. T., Jr.; Hemingway, B. S. Heat T. Comparison of the Kinetics of Flash Calcination of Kaolinite in
Capacities and Entropies of Rhodochrosite (MnCO3) and Siderite Different Calciners. Chem. Eng. Res. Des. Part A: Trans. Inst.
(FeCO3) between 5 and 600K. Am. Miner. 1984, 69, 349. Chem. Eng. 1994, 72 (A1), 105.
(47) Robie, R. A.; Russell-Robinson, S.; Hemingway, B. S. Heat (67) CRC Handbook of Chemistry and Physics, 73rd ed.; Lide,
Capacities and Entropies from 8 to 1000 K of Langbeinite (K2- D. R., Ed.; CRC Press: Boca Raton, Fl, 1992-1993.
Mg2(SO4)3), Anhydrite (CaSO4) and of Gypsum (CaSO4‚2H2O) to
325 K. Thermochim. Acta 1989, 139, 67.
(48) Skauge, A.; Fuller, N.; Hepler, L. G. Specific Heats of Clay
Received for review December 30, 1999
Minerals: Sodium and Calcium Kaolinites, Sodium and Calcium
Montmorillonites, Illite and Attapulgite. Thermochim. Acta 1983,
Revised manuscript received April 24, 2000
61, 139. Accepted April 24, 2000
(49) Skauge, A.; Fuller, N.; Yan, H.-K.; Cassis, R.; Srinivasan,
N. S.; Hepler, L. G Specific Heat Capacities of Minerals from Oil
Sands and Heavy Oil Deposits. Thermochim. Acta. 1983, 68, 291. IE990942G

You might also like