You are on page 1of 11

Article

pubs.acs.org/EF

Oil Shale Pyrolysis: Conversion Dependence of Kinetic Parameters


Juliana Pedrilho Foltin,†,‡ Antonio Carlos Luz Lisboa,‡ and Arno de Klerk*,†

Department of Chemical and Materials Engineering, University of Alberta, 9211 116th Street, Edmonton, Alberta T6G 1H9, Canada

School of Chemical Engineering, The State University of Campinas (UNICAMP), Campinas - SP, Brazil

ABSTRACT: Oil can be recovered from kerogen in oil shale by pyrolysis. The devolatilization kinetics of the pyrolysis of oil
shale from the Irati Formation in Brazil was studied. Kinetic parameters were determined from dynamic thermogravimetric
analysis over the temperature range 323−1173 K, using different model-free methods. Evaluation and validation were performed
by pyrolysis at 673 K for 3 hours. It was found that the activation energy depended on the extent of conversion. Activation energy
increased over the range 215−255 kJ/mol for conversion in the range 0.15 ≤ α ≤ 0.55, where α = 1 for pyrolysis at 1173 K.
When the reaction rate was high, the conversion calculated using kinetic parameters derived by the Friedman method was more
accurate than those calculated from the Flynn−Wall−Ozawa and the Kissinger−Akahira−Sunose methods. The latter two
methods performed better when the reaction rate was lower, i.e., at higher conversion. Isothermal kerogen pyrolysis approached
an incomplete conversion limit that could be increased only by increasing the temperature; this type of behavior was predicted by
the conversion dependence of activation energy. The observed activation energy is an average of the different activation energies
of the individual compounds in kerogen. As conversion progresses, the compounds with lower activation energies are more
readily converted, so that the average activation energy of the compounds that remain increases with increasing conversion. The
work highlighted the importance of employing conversion-dependent kinetic parameters when modeling oil shale pyrolysis for
process design, especially when the process is designed for high kerogen conversion.

1. INTRODUCTION This investigation had two objectives. The first objective was
to develop a kinetic description of devolatilization by pyrolysis
The term “shale” is used to describe geologically quite different
of oil shale from the Irati Formation in Brazil. The second
geological formations, but the formations have some common
objective was to evaluate the performance of different
features. The organic component (kerogen) is geologically approaches to the kinetic modeling of oil shale pyrolysis for
immature, with low solubility in solvents, which can be practical application. The scope of the kinetic investigation was
converted to oil and gas by pyrolysis.1−3 Oil shale deposits restricted to the use of thermogravimetric analysis (TGA) as a
are widely distributed, with the most abundant deposits found tool for kinetic measurements, although the interpretation of
in the United States and Brazil, and significant deposits exist in the pyrolysis behavior was supplemented by characterization
Russia, Congo, Canada, Italy, China, Australia and Estonia.4,5 and spectroscopic analyses of the oil shale. TGA has been
Above-ground mining and in situ subsurface recovery methods successfully employed for the development of pyrolysis kinetics
are used for oil shale, but unlike oil sands, the organic with various oil shales.15−19 The data were modeled using
component cannot effectively be displaced at moderate model-free and model-fitting approaches in order to evaluate
temperatures by water or organic solvents; shale oil recovery the activation energy and the pre-exponential factor.20,21 This
from oil shale requires temperatures in excess of 300 °C.6 was followed by a critical evaluation of the kinetics and the
Potentially commercially attractive grades of oil shale contain at implications for practical use in the design of oil shale pyrolysis
least 0.1 m3/t (∼25 US gal/t) recoverable oil.4 processes.
When reviewing the technologies developed or employed for
shale oil production from oil shale,5,7−11 it is immediately 2. EXPERIMENTAL SECTION
apparent that pyrolysis plays a central role. The devolatilization 2.1. Materials. The oil shale used in this study was extracted from
kinetics is important in describing the rate of oil evolution as a the Irati Formation and supplied by Petrobras/SIX, located in São
function of pyrolysis time and temperature. This important first Mateus do Sul, Parana, Brazil. Two samples were employed in this
step in the development of process technology is the topic of study, with particle size ranges of 0.297−0.354 mm (48 mesh to 42
this work. Pyrolysis of the oil shale to liberate shale oil involves mesh) and 0.500−0.595 mm (32 mesh to 28 mesh). These size
fractions were obtained by sieving the material using Tyler screens, for
a complex reaction network, and it involves the breaking of example, recovering material that passed through a 42 mesh, but did
bonds with different activation energies within the organic not pass through a 48 mesh Tyler screen. The smaller particle range
matter and between the organic and mineral matter.12,13 There was employed for the kinetic study, while the larger particle range was
is consequently a distribution of activation energies, rather than employed to check that mass transfer was not limiting. The oil shale
a single activation energy, that must be considered during was characterized as part of the investigation (see section 3.1). The
pyrolysis. The actual time−temperature−gas environment also
affects the composition of the shale oil produced by oil shale Received: February 25, 2017
pyrolysis, which in turn affects downstream refining.14 Revised: June 6, 2017
However, this aspect was not studied.

© XXXX American Chemical Society A DOI: 10.1021/acs.energyfuels.7b00578


Energy Fuels XXXX, XXX, XXX−XXX
Energy & Fuels Article

focus was on the kerogen and mineral matter components that may Table 1. Proximate Analysis of the Oil Shale from the Irati
affect pyrolysis and the kinetic investigation. X-ray diffraction data is Formation, Southern Brazil
also presented, but characterization did not include a full mineralogical
characterization, which was reported in the literature recently.22 description proximate analysis (wt %)
2.2. Equipment, Procedure, and Analyses. Oil shale pyrolysis Measured
was studied using thermogravimetric analysis with dynamic heating. In moisture 1.8
a typical experiment, around 9.5 mg of oil shale with particle size range
volatile matter 18.8
of 0.297−0.354 mm was weighed and put inside an alumina crucible,
with the exact weight being recorded using a Mettler Toledo XS105 fixed carbon 2.8
balance with 0.01 mg readability. The mass loss versus a defined ash 76.6
temperature measurement was performed using a Mettler Toledo Calculated
TGA/DSC1, LF 1100 furnace, with MX5 microbalance having a volatile matter, dry basis 19.1
measuring range of 5 g and 1 μg readability. The analyses were fixed carbon, dry basis 2.8
performed under constant nitrogen flow rate of 50 mL/min. The ash, dry basis 78.1
measurements were conducted at atmospheric pressure, which in
Edmonton, AB, Canada is on average 93 kPa because of its altitude. Table 2. CHNSO Elemental Analysis of the Oil Shale from
Mass loss was recorded over the temperature range 373−1173 K
(100−900 °C) at different heating rates (β): 2, 5, 10, 15, 20, 25, 40,
the Irati Formation, Southern Brazil
and 50 K/min. The final mass was considered to be mineral matter description ultimate analysis (wt %)
and unrecoverable organic matter, i.e., all material that did not form
carbon 14.0
vapor and gaseous products.
Two control experiments were also performed by TGA using 9.5 hydrogen 1.9
mg oil shale, one employing the 0.297−0.354 mm size range that was nitrogen 0.3
used for all TGA with dynamic heating, the other employing the sulfur 3.8
0.500−0.595 mm size range. The TGA of each sample was conducted oxygen 6.3
under constant nitrogen flow rate of 50 mL/min. The temperature mineralsa 73.7
program had three segments: dynamic heating at β = 50 K/min from a
Calculated by difference; material not converted to COx, H2O, N2,
323 to 673 K, then remaining isothermal at 673 K for 180 min, and and SO2 under pyrolytic and oxidative pyrolytic conditions.
finally dynamic heating at β = 50 K/min from 673 to 1173 K. The
objectives of these experiments were to provide isothermal kinetics for
validating the kinetic analysis performed by nonisothermal TGA, as
well as to determine whether mass transport meaningfully affected the
kinetic analysis. exclude the detection of those elements derived from mineral
In addition to the pyrolysis experiments performed by TGA, matter. The method of CHNS determination involves high-
pyrolysis was also studied by infrared spectroscopy. For this purpose, temperature pyrolytic oxidation with detection of CO2, H2O,
an ABB MB3000 Fourier transform infrared (FTIR) spectrometer was N2, and SO2. Residual water and minerals that can produce any
employed, which was equipped with a Pike DiffusIR attachment that of the aforementioned gases would cause an increase in the
has an environmental chamber. The environmental chamber is reported values for the corresponding CHNS elements. Oxygen
temperature-controlled and can be heated to 773 K. The oil shale content is determined separately and relies on the formation of
(3.5 mg) was mixed with potassium bromide (35.0 mg) and pressed CO so that any reduction of mineral matter with transfer of
into the sample holder. The environmental chamber was purged with
nitrogen, and the oil shale was heated to different temperatures at
oxygen to the organic matter will increase the reported oxygen
which the infrared spectra were recorded. Infrared spectra were content.
collected with the following parameters: resolution of 16 cm−1, average The oil shale was not porous, and the total surface area was
of 120 scans, 729 detector gain, and spectral range of 5000−500 cm−1. <1 m2/g. The oil shale was analyzed by X-ray diffraction
Oil shale characterization was performed using different techniques. (Figure 1), and the most prominent minerals were silicon
Proximate analysis was performed using a Leco TGA 701 in dioxide, iron oxide, and calcium sodium aluminum silicate.
accordance with ASTM Standard 7582 test method for coal.23 There was also evidence of what appears to be illite. Not all of
CHNSO elemental analysis of the organic matter was performed using
a Thermo Fisher Flash 2000 Organic Elemental Analyzer. Elemental
analysis of the elements heavier than Na in the shale oil was performed
with a Bruker S2 Ranger X-ray fluorescence (XRF) spectrometer. The
X-ray diffraction (XRD) data was collected using a Philips X’PERT
MPD diffractometer.

3. RESULTS
3.1. Characterization of Oil Shale. Proximate analysis of
oil shale is a convenient method to determine the relative
abundance of potentially recoverable shale oil. The proximate
analysis of the oil shale from the Irati Formation is presented in
Table 1. The proximate analysis was comparable to that for
mixed Irati oil shale reported by Carter et al.:24 2.2 wt %
moisture, 22.8 wt % organic matter, and 76.9 wt % ash. The
values reported by Petersen et al.13 were 1.9 wt % moisture and
72.6 wt % ash.
The CHNSO analysis of the oil shale is presented in Table 2. Figure 1. X-ray diffraction pattern of the Irati oil shale. The main
Although the intent of the analysis is to determine the CHNSO minerals identified were silicon dioxide (●), iron oxide (⧫), calcium
composition of the organic matter, the method does not sodium aluminum silicate (■), and possibly illite (▲).

B DOI: 10.1021/acs.energyfuels.7b00578
Energy Fuels XXXX, XXX, XXX−XXX
Energy & Fuels Article

the peaks in the diffraction pattern could be unambiguously Table 4. Onset Temperature (Tonset), Temperature at
assigned. Maximum Rate of Mass Loss (Tmax), and the Fraction of the
The elements heavier than Na were quantified using XRF Total Mass Loss at Tmax (αmax) in Figure 2 for Each Heating
(Table 3). The two most abundant elements were Si and Fe, Rate Recorded
with 38.1 and 28.2 wt %, respectively. This was consistent with region IIa region IIb
X-ray diffraction results presented in Figure 1.
heating rate, β Tonset Tmax αmax Tonset Tmax αmax
(K/min) (K) (K) (−) (K) (K) (−)
Table 3. Elements Heavier than Na in the Oil Shale from the
2 608 687 0.42 763 782 0.70
Irati Formation, Southern Brazil
5 638 706 0.45 784 794 0.71
description content (wt %)a 10 672 717 0.44 796 804 0.71
Al 7.2 15 677 720 0.43 804 810 0.74
Si 38.1 20 690 729 0.47 805 813 0.75
S 10.0 25 689 734 0.46 810 820 0.74
K 7.2 40 688 741 0.44 816 826 0.71
Ca 7.1 50 689 748 0.46 820 831 0.72
Ti 1.5
Mn 0.2 heating rates of above 20 K/min. The temperature at which the
Fe 28.2 maximum rate of mass loss was observed monotonically
Zn 0.1 increased with increasing heating rate, but the fraction of mass
Sr 0.2 lost up to that temperature remained within the range 0.42−
Zr 0.2 0.47.
a
Normalized values; only elements with 0.1 wt % and higher reported.
The onset temperature and temperature at maximum rate of
mass loss for region IIb increased monotonically with an
3.2. Pyrolysis of Oil Shale. The oil shale with size range increase in heating rate. Like region IIa, the fraction of mass lost
0.297−0.354 mm was pyrolyzed under N2 flow in a TGA at Tmax did not depend on heating rate and remained within the
instrument under dynamic heating conditions. The mass loss range 0.70−0.75.
was measured as a function of temperature at different heating Although the effect of heating rate on temperature at which
rates (Figure 2). This data was used as basis for the kinetic mass loss events are observed is noticeable from Figure 2, the
total mass loss at the highest pyrolysis temperature, 1173 K,
was between 19.0 and 20.5 wt %, irrespective of the heating rate
(Table 5). There was no specific trend with respect to total

Table 5. Mass Loss Recorded by TGA at Two Intermediate


Temperatures and the Maximum Temperature of Analysis
mass loss (wt %)
heating rate, β (K/min) 623 K 873 K 1173 K
2 2.86 16.0 20.5
5 2.38 16.1 20.1
10 2.24 15.5 19.4
15 2.29 16.0 19.7
20 2.08 16.0 19.2
25 2.09 15.7 19.6
40 2.04 15.9 20.4
Figure 2. TGA mass loss versus temperature data at different heating 50 1.80 14.8 19.0
rates.
mass loss and heating rate. This was an important observation,
modeling. The shape of the mass loss curves did not change at because heating rate affects the oil yield when pyrolysis is
different heating rates. Based on the temperature response of terminated at lower temperatures, which is more typical of
mass loss, three regions were defined in Figure 2, which is practical pyrolysis operation.14
explained in section 4.1. The two control experiments that were performed in the
The most important region for oil shale pyrolysis is region II. TGA, which evaluated pyrolysis of two samples of two different
The onset temperature of each region, as well as the particle size ranges, are presented in Figure 3. These control
temperature at which the maximum rate of mass loss is experiments provided TGA data on the isothermal pyrolysis of
observed, depends on the heating rate. The pyrolysis appears to kerogen at 673 K over an extended period. This data was not
take place in two stages, which will be referred to as regions IIa employed in the kinetic modeling but was used as a validation
and IIb. The onset temperature (Tonset), temperature at set to discriminate between the different methods. The oil shale
maximum rate of mass loss (Tmax), and the fraction of the with larger particle size range had a slightly higher overall mass
total mass loss found at Tmax (αmax) for each of these two stages loss. The total mass loss after pyrolysis to 1173 K (Figure 3)
are listed in Table 4. was 18.9 and 20.0 wt % for the 0.297−0.354 mm (48 mesh to
With increasing heating rate, the observed onset temperature 42 mesh) and 0.500−0.595 mm (32 mesh to 28 mesh) size
for region IIa increases, becoming near constant at 690 K at ranges, respectively.
C DOI: 10.1021/acs.energyfuels.7b00578
Energy Fuels XXXX, XXX, XXX−XXX
Energy & Fuels Article

place during each, and analogous descriptions can be found for


coal pyrolysis.26,27

Figure 3. TGA control experiments.

3.3. Infrared Spectroscopy. Infrared spectra of the oil


shale were collected at different temperatures to determine in Figure 5. Temperature regions of oil shale mass loss as illustrated by
what way the changes affected composition. To illustrate the TGA data at β = 10 K/min.
changes, the infrared spectra at 623, 683, and 743 K are shown
(Figure 4). These represent the onset of region IIa, temperature In the first region, region I, the mass loss is attributed to
of high rate of mass loss, and close to the end of region IIa. evaporation of lighter organic species and moisture. It has been
reported that in this temperature region hydrogen bonding is
also disrupted,27 but such changes do not lead to mass loss,
unless it causes the liberation of a species that is volatile at the
temperature conditions. It should be noted that no lower
temperature limit for the decomposition of kerogen has been
established.10 The contribution of decomposition reactions
cannot be ruled out, but due to the time scale of the analysis,
their contribution to mass loss in region I is minor.
The start of the second region, region II, is at around 600 K
when pyrolysis of the kerogen becomes meaningful. Evapo-
ration of heavier species continues to contribute to the mass
loss, but the dominant contribution is that of pyrolysis of heavy
material to produce lighter material that can readily evaporate
at the temperature conditions. There appears to be two
distinctive subregions, as can be seen from the two “steps” in
the mass loss versus temperature relationship (Figure 5).
Region IIa is the normal pyrolysis regime, where most of the oil
Figure 4. Infrared spectra of oil shale heated to different temperatures: is evolved, and it is extensively described in the literature.14 The
623, 683, and 743 K.
Fischer assay that involves pyrolysis to 773 K10 provides a
standardized estimate of the recoverable kerogen in regions I
The presence of organic material (kerogen) in the oil shale is and IIa.
clearly seen from the C−H absorption bands due to Region IIb is the coking regime and starts at around 800 K.
unsaturated/aromatic compounds around 3150 and 3024 This two-step pyrolysis is commonly observed in studies where
cm−1 and saturated aliphatic C−H absorption bands around the temperature range of the investigation is extended beyond
2950 cm−1.24 Carbon dioxide absorption at 2353 and 671 cm−1 800 K.16−18,28−30 In this region pyrolysis continues, but the
was seen in all spectra; these absorption bands are reported at ability to form species that can be evaporated is restricted by
2349 and 667 cm−1,24,25 and the 4 cm−1 shift is likely due to the availability of transferable hydrogen within the remaining
resolution (16 cm−1) employed for spectral acquisition. organic material. Hydrogen disproportionation is central to the
The most significant changes in the infrared spectra with thermal cracking and coking reactions and determines to what
increasing temperature are the decrease in the 1589 cm−1 extent volatile products can be formed. Pyrolysis in region IIb is
absorption, increase in the 1110 and 617 cm−1 absorption reportedly also enhanced by mineral matter catalysis.29 Kerogen
bands, and the development of a new absorption band at 563 on its own does not exhibit a clear change in rate to distinguish
cm−1. There also appeared to be changes in absorption around this decomposition stage,29 and region II mineral-free kerogen
1720 and 1404 cm−1. does not have two separately identifiable subregions.
Aluminum silicates could have some Brønsted acidity that
4. DISCUSSION could facilitate cracking at high temperature, and aluminum
4.1. Changes during Oil Shale Pyrolysis. Three thermal silicates were found in the oil shale investigated (Figure 1),
regions can be identified based on the change in the rate of albeit in association with alkaline and alkaline earth metals. The
mass loss with increasing temperature (Figure 5). These presence of iron oxide (Figure 1), which would readily be
thermal regions have meaning based on the processes taking sulfided on release of H2S during pyrolysis, may also have
D DOI: 10.1021/acs.energyfuels.7b00578
Energy Fuels XXXX, XXX, XXX−XXX
Energy & Fuels Article

contributed to catalysis in region IIb. However, no work was no MgCO3 was present in the mineral matter. Calcium was
done to distinguish between mineral-catalyzed and thermal present in the mineral matter (Table 3), but the decomposition
reactions. It should further be pointed out that the measured of CaCO3 becomes meaningful at a temperature >1000 K10 and
surface area of the oil shale was <1 m2/g, which would would not interfere with a TGA-based kinetic analysis of
inherently limit the contribution of mineral catalysis. It is more kerogen pyrolysis.
likely that the mineral matter acted as a nucleation site for Nevertheless, it was worthwhile to confirm whether
carbonaceous deposits,31 thereby facilitating the removal of carbonate minerals were present in meaningful quantities in
such material from the kerogen and indirectly affecting the the oil shale. Metal carbonates have two characteristic
reaction rate. absorption bands in the infrared spectrum, a strong absorption
In the third region, region III, further mass loss is due to at 1450−1410 cm−1 and a less intense absorption at 880−850
high-temperature decomposition of the remaining kerogen, as cm−1.35 A minor absorption band was found at 1443 cm−1, and
well as decomposition of some minerals when present, such as no associated absorption band in the 880−850 cm−1 was
dolomite and calcite.10 The pyrolysis process also causes detectable in the infrared spectra (Figure 4), which indicated a
physical changes in the mineral matter. The expansion caused low carbonate content. This was consistent with the literature
by kerogen during pyrolysis can fracture the mineral matter, for this particular oil shale deposit.21 Furthermore, no
which increases connectivity between the pores, providing carbonate minerals were identified (Figure 1).
additional pathways for escaping fluids and gases.28 Thus, Characterization of the minerals found at different depths
products that may have formed at a lower temperature, but that and locations in the Irati Formation indicated that quartz,
was trapped, can now escape and be measured as mass loss. albite, and smectite were common to all samples. Illite was also
The use of infrared spectroscopy (Figure 4) to study the found in all samples, but in one it was present in low quantity.
changes in the oil shale as it was heated through region IIa was Other minerals that were found in only some of the samples
instructive. The first observation is that with increasing were pyrite, kaolinite, chlorite, analcine and gypsum.21 Figure 1
temperature, there was a decrease in the strong asymmetrical indicated that silicon dioxide (e.g., quartz) and aluminum
carboxylate absorption band at 1589 cm−1, with changes in the silicates (e.g., albite and smectite) were present and that illite
weaker symmetrical carboxylate ion stretching being obscured was possibly present.
by the prominent absorption at 1404 cm−1. Concomitantly The possible presence of iron pyrite (FeS2), as identified in
there appeared to be an increase in absorption around 1720 some Irati oil shale samples,13,21 bore further investigation. In
cm−1, the carbonyl absorption region. These changes suggest the presence of organic matter at elevated temperature, iron
that kerogen present as carboxylic acids, bound to the mineral pyrite can result in transfer of sulfur to the organic matter,
matter as carboxylate salts, were liberated from the mineral thereby affecting the composition of the oil kerogen. Analysis of
matter by decomposition and ketonization. It is known from the oil shale indicated a high content of both Fe and S (Table
the literature that the organic material obtained from the Irati 3), but the molar ratio of Fe:S was 1:0.6; therefore, only some
oil shale Formation has a high heteroatom content,32 and of the Fe could be in the form of iron pyrite. The presence of
moreover, it is rich in ester-linkages.33 Plausible reaction steps iron pyrite could not be confirmed by infrared spectroscopy,
as temperature is increased include hydrolysis or decomposition because the main infrared absorption band of iron pyrite is
of the esters to produce carboxylic acids, reaction of the found at 350 cm−1.36 No evidence of iron pyrite was found by
carboxylic acids with basic mineral matter to produce metal X-ray diffraction analysis (Figure 1); in fact, iron oxide was one
carboxylates, and ultimately the decomposition of the metal of the main components on the oil shale. Even though some
carboxylates, as was observed by infrared spectroscopy (Figure iron pyrite might have been present, the transfer of sulfur is
4). slow, and it was unlikely that it would meaningfully interfere
Other changes appear to be related mainly to the mineral with the kinetic analysis. For example, it was found that iron
matter. The 1110 cm−1 absorption becomes increasingly pyrite decomposition at 673 K for 20 h affected a region of
prominent as more kerogen is removed from the oil shale by about 1 μm of organic matter in coal.37
pyrolysis. The absorption could be due to a sulfate (SO42−) 4.3. Deriving Kinetics from TGA Data. The use of TGA
group,25 and it was speculated that it was related to the to determine devolatilization kinetics during pyrolysis is one of
presence of gypsum in the Irati oil shale,21 but sulfate minerals the more reliable experimental approaches reported in the
were not specifically identified in the diffractogram of the oil literature.27 Its main drawbacks for kinetic analysis are the
shale employed for this study (Figure 1). It is more likely that comparatively slow heating rate and temperature limitations;27
the absorption around 1110 cm−1 is due to SiO2,34 which would the latter is not a relevant limitation in this study. Kinetics
be consistent with the high SiO2 content observed in Figure 1. parameters were determined by relating the mass loss versus
The increase in 617 cm−1 absorption and the appearance of temperature response at different heating rates to the oil shale
absorption at 563 cm−1 is tentatively assigned to absorption by pyrolysis kinetics. The kinetics of decomposition analysis are
various metal oxides,24 which in this study is likely dominated based on the Arrhenius equation (eq 1) and kerogen
by iron oxides. transformation rate for the volatile product (eq 2):
4.2. Impact of Minerals on Thermogravimetric Data.
k = k 0·e(−E / RT ) (1)
Carbonate minerals are usually the most troublesome for
kinetic studies, because the carbonates can decompose to dα
release CO2, which is recorded as a mass loss. Dolomite = k(T ) ·f (α)
dt (2)
(CaCO3·MgCO3) in particular can skew the TGA data for
kinetic analysis, because the half decomposition of MgCO3 In these equations, k is the rate constant, k0 the pre-
becomes meaningful at temperatures above 650 K.10 The XRF exponential factor, and E the activation energy. The conversion
analysis (Table 3) indicated that Mg, if present, was present at α is the standard form of the sample mass loss and is defined
<0.1 wt %. This was important because it suggested that little or according to eq 3, in which mf is the mass remaining after
E DOI: 10.1021/acs.energyfuels.7b00578
Energy Fuels XXXX, XXX, XXX−XXX
Energy & Fuels Article

pyrolysis to 1173 K (i.e., not conversion expressed in terms of


the Fischer assay procedure).10
m −m
α= o
mo − mf (3)
The combination of eqs 1 and 2 provides the fundamental
expression, eq 4, which is based on analytical methods to
calculate the kinetic parameters based on results of TGA.

= k 0·f (α) ·e(−E / RT )
dt (4)
The function f(α), eq 5, is inserted into eq 4, producing eq 6,
wherein the parameter β = dT/dt, the heating rate, is used, and
for a first-order reaction n = 1:
f (α) = (1 − α)n (5) Figure 6. Kinetic analysis of TGA data using the Friedman method at
different heating rates, with β = 2, 5, 10, 15, 20, 25, 40, and 50 K/min.
dα k
= 0 ·(1 − α)n ·e(−E / RT )
dT β (6)
Four different methods were employed to calculate the
pyrolysis kinetics from the TGA data, namely, Friedman (eq 7),
Flynn−Wall−Ozawa (eq 8), Kissinger−Akahira−Sunose (eq
9), and Coats−Redfern (eq 10). Except for the Coats−Redfern
method, the other methods were associated with model-free
descriptions of the kinetics.
⎛ dα ⎞ E
ln⎜β ⎟ = ln[k 0·f (α)] −
⎝ dT ⎠ RT (7)

⎛ k ·E ⎞ E
ln(βi ) = ln⎜ 0 ⎟ − 5.331 − 1.052
⎝ R · g (α ) ⎠ RT (8)
Figure 7. Kinetic analysis of TGA data using the Flynn−Wall−Ozawa
⎛ β⎞ ⎛ k ·R ⎞ E (FWO) method at different heating rates, with β = 2, 5, 10, 15, 20, 25,
ln⎜ 2 ⎟ = ln⎜ 0 ⎟−
⎝T ⎠ ⎝ E ·g (α) ⎠ RT (9)
40, and 50 K/min.

⎛ g (α ) ⎞ ⎛ k ·R ⎞ E
ln⎜ 2 ⎟ = ln⎜ 0 ⎟ −
⎝ T ⎠ ⎝ β · E ⎠ RT (10)
As explained before, the mass loss that occurred in region I
was mainly due to evaporation of lighter material in the
kerogen and not due to pyrolysis. The fraction of the total mass
loss that took place in region I was in the range α = 0.10−0.15.
The kinetic analysis is therefore valid only for values of α ≥
0.15, and the calculated kinetics at lower conversion has no real
meaning. Although region IIb represented pyrolysis, the
literature indicated that this was due to mineral matter
catalysis.29 The conversion at the start of region IIb was
around 65%, and the kinetics analysis of oil shale kerogen
pyrolysis was consequently limited to α ≤ 0.65.
The methods evaluated (eqs 7−10) are all in the format of a Figure 8. Kinetic analysis of TGA data using the Kissinger−Akahira−
Sunose (KAS) method at different heating rates, with β = 2, 5, 10, 15,
linear relationship with respect to reciprocal temperature (1/ 20, 25, 40, and 50 K/min.
T). These relationships are shown for different levels of
conversion for each of the models in Figures 6−9.
The pre-exponential factor and activation energy for the first- thermal cracking (region IIa) to mineral matter assisted
order reaction (n = 1) derived at different levels of conversion cracking (region IIb) was approached. The Friedman method
using the model-free methods are listed in Table 6. It should be was particularly sensitive to the change. A significant increase in
pointed out that each value of the activation energy reported the activation energy at 0.60 and 0.65 conversion is indicated by
actually represents a weighted average of a distribution of the Friedman method (Table 6), whereas this change is
activation energies12,13 and does not suggest that the activation detected only at α = 0.65 by the other two model-free methods.
energy has a single value only. The calculated correlation Even so, lower activation energies were calculated for Flynn−
coefficients for linear regression were generally r2 > 0.97 in the Wall−Ozawa and Kissinger−Akahira−Sunose methods com-
range 0.15 ≤ α ≤ 0.55 but became worse as the transition from pared to the Friedman method.
F DOI: 10.1021/acs.energyfuels.7b00578
Energy Fuels XXXX, XXX, XXX−XXX
Energy & Fuels Article

Figure 9. Kinetic analysis of TGA data using the Coats−Redfern Figure 10. Comparison of calculated reaction rates by different
method at different heating rates, with β = 2, 5, 10, 15, 20, 25, 40, and methods with experimental reaction rate determined during isothermal
50 K/min. pyrolysis at 673 K.

Kinetic parameters determined by the Coats−Redfern


It was reported in the literature that the Friedman method is method were insensitive to heating rate, which agreed with
considered the most accurate of the methods, because it did not previous observations.18 Using the Coats−Redfern method
employ approximations.18 However, the Friedman method was activation energies in the range of 75−80 kJ/mol (average 76.4
also more sensitive to noise in the data and can become kJ/mol, sample standard deviation 5.1 kJ/mol) were found,
numerically unstable. Noise in the data can erode the potential compared to activation energies of >210 kJ/mol for the other
gain in accuracy and the claim that the Friedman method is methods (Table 6). The values obtained from the Coats−
more accurate does not always hold true.17 Redfern method were too low to be realistic.
In order to discriminate between the methods, the kinetic When the pyrolysis of oil shale from the Irati Formation was
parameters in Table 6 were used to calculate the reaction rate studied using electron spin resonance (ESR) spectroscopy,38
for oil shale pyrolysis at 673 K, and these values were compared two first-order reactions occurring in parallel were identified as
with the experimental reaction rate obtained in the control the main free radical producing reactions. The activation
experiments (Figure 10). Numerically the Friedman method energies for these two reactions were 226 ± 7 and 150 ± 40 kJ/
predicted reaction rates that were lower than those predicted mol respectively, with associated pre-exponential factors of 0.9
based on the Flynn−Wall−Ozawa and Kissinger−Akahira− × 1017 and 2.3 × 109 h−1 (or for comparison with Table 6, 1.5
Sunose methods. × 1015 and 3.8 × 107 min−1).
The Friedman method performed better at lower conversion, As another point of reference, Petersen et al.13 reported an
when the overall reaction rate was higher and the signal-to- activation energy distribution for Irati oil shale. There was a
noise ratio of the data was higher. Generally, the Friedman minor contribution of activation energies in the ranges 167−
method resulted in kinetic parameters that under-predicted the 192 and 247−260 kJ/mol, but most of the material had
observed reaction rate at 673 K. The Flynn−Wall−Ozawa and activation energies in the range 201−238 kJ/mol. The
Kissinger−Akahira−Sunose methods performed better at activation energy distribution centered around 218−222 kJ/
higher conversion and tended to overestimate the reaction rate. mol, which represented 63% of the kerogen. The average pre-

Table 6. Calculated Pre-exponential Factor (k0, min−1) and Activation Energy (E, kJ/mol) for the Pyrolysis Reaction Using
Different Model-Free Methods
Friedman Flynn−Wall−Ozawa Kissinger−Akahira−Sunose
α E k0 r2 E k0 r2 E k0 r2
0.05a 143 2.7 × 1012 0.773 121 1.0 × 1011 0.895 119 5.3 × 1010 0.881
0.10a 141 2.1 × 1010 0.955 180 2.7 × 1014 0.895 179 2.3 × 1014 0.884
0.15 221 8.1 × 1015 0.982 216 1.7 × 1016 0.977 216 1.7 × 1016 0.975
0.20 217 2.1 × 1015 0.995 219 1.0 × 1016 0.993 219 1.0 × 1016 0.992
0.25 232 2.5 × 1016 0.999 211 1.8 × 1015 0.990 211 1.6 × 1015 0.989
0.30 234 3.0 × 1016 0.998 218 4.6 × 1015 0.996 218 4.3 × 1015 0.995
0.35 228 9.4 × 1015 0.995 240 1.5 × 1017 0.995 240 1.6 × 1017 0.995
0.40 249 2.4 × 1017 0.994 239 1.1 × 1017 0.998 240 1.2 × 1017 0.998
0.45 240 4.4 × 1016 0.995 246 3.0 × 1017 0.992 247 3.3 × 1017 0.991
0.50 253 3.0 × 1017 0.988 237 4.8 × 1016 0.997 237 4.8 × 1016 0.996
0.55 269 3.0 × 1018 0.969 245 1.7 × 1017 0.992 246 1.8 × 1017 0.992
0.60 300 2.0 × 1020 0.931 255 5.6 × 1017 0.982 256 6.3 × 1017 0.981
0.65 328 4.6 × 1021 0.703 301 5.6 × 1020 0.868 304 8.3 × 1020 0.858

a
Region I, kinetic analysis invalid because of significant contribution of evaporation without pyrolysis.

G DOI: 10.1021/acs.energyfuels.7b00578
Energy Fuels XXXX, XXX, XXX−XXX
Energy & Fuels Article

exponential factor was 4.4 × 1013 s−1 (or for comparison with determine the kinetics, but that too would be an over-
Table 6, 2.6 × 1015 min−1). These values are close to the values simplification.
the conversion range 0.15 ≤ α ≤ 0.30 reported in Table 6. However, in the literature it is not uncommon to find that
Considerable variation in activation energies reported for investigators report a single average value for activation energy,
different types of oil shale is found, e.g. see refs 12, 13, and 39, because it enables an analytical solution for the reaction
but the activation energy values for Irati oil shale calculated by kinetics. For conversion in the region 0.15 ≤ α ≤ 0.55, the
the Coats−Redfern method were clearly outside of the range of values for activation energy are mainly in the range 215−255
realistic values. The failure of the Coats−Redfern method to kJ/mol−1. However, using a single value for activation energy is
yield realistic results may be related to the model used. A large an oversimplification, which not only averages the activation
number of reaction models have been proposed.18 A systematic energy distribution at a specific level of conversion but also
evaluation of all the possible reaction models was not averages the activation energies over different levels of
undertaken, because all of the models are based on the conversion.
assumption that there is a single reaction pathway. However, Kerogen is not a single substance. The measurement of the
the ESR investigation of Irati oil shale pyrolysis showed two overall pyrolysis kinetics is a measurement of the contribution
reaction pathways,38 and the same is reported in more general of numerous different reactions in parallel. Each of these
overviews.14 pyrolysis reactions has its own activation energy, and the
4.4. Conversion Dependence of Activation Energy. activation energies for pyrolysis of different species can be quite
The activation energy for the pyrolysis reaction changes with different. The activation energies can be represented as a
conversion (Table 6). This can be more clearly seen when distribution curve of fraction of kerogen in relation to activation
plotting the activation energy as a function of conversion energy. For example, this was done in the development work
(Figure 11). The values derived from the kinetic analysis performed by ExxonMobil (normal distribution with activation
energies ranging from 209−234 kJ/mol)40 and by Petersen et
al. (distributions of activation energies that were not necessarily
normal distributions).13
It is convenient to lump the different species together and
express the activation energy as a single value, because it makes
the analytical solution of the rate equation possible. However,
that does not make the actual system homogeneous, and it is
questionable whether this type of mathematical simplification is
necessary. The rate at which the species with low pyrolysis
activation energy is depleted is not the same as the rate at
which the species with high pyrolysis activation energy are
depleted. Because all of the reactions are pyrolysis-type
reactions, it is likely that the reactions requiring low activation
energy would proceed more rapidly and start proceeding at
lower temperature than those requiring high activation energy,
despite potential differences in their pre-exponential factor.
Figure 11. Activation energies calculated by model-free methods as a With an increase in conversion there will be change in the ratio
function of conversion. of pyrolysis reactions requiring low activation energy compared
to pyrolysis reactions requiring high activation energy. The
represent weighted averages of activation energy distribu- relative contribution of different activation energies will
tions.12,13 With increasing conversion it is also possible that consequently change to indicate that on average higher
there is a change in the shape of the distribution of activation activation energy is required for pyrolysis as conversion is
energies of the kerogen. The kinetic analysis did not determine increased (Figure 11). This is not a consequence of the increase
the change in the shape of the activation energy distribution at in temperature, but a consequence of the extent of conversion
different conversion levels, only that there was a progressive that changes the chemical nature of the kerogen.
increase in average activation energy with increasing con- 4.5. Conversion Dependence of Observed Kinetics. A
version. known but often ignored drawback of TGA as a technique for
Burnham14 makes the point clearly that the model determining oil shale pyrolysis kinetics is that the detection of
description of kerogen → bitumen → oil and gas is not conversion is related to the detection of mass loss. Reactions
supported by experimental evidence, despite its persistence in leading to the formation of oil products that are not volatile at
the literature. There are at least two paths to oil and gas. The the measurement temperature will not be recorded as a mass
existence of at least two major pathways to oil and gas during loss by TGA. Likewise, there are reactions, leading to coke-like
pyrolysis is also supported by quantitative ESR analysis.38 Any residue remaining on the mineral matter,10 that cannot be
change in the ratio of these two pathways would lead to a observed by TGA either, because the product is not volatile.
change in the observed activation energy. Because it is likely At lower pyrolysis temperatures and lower conversion levels,
that the contribution of the energetically more demanding there is a higher chance of reactions of the type kerogen →
pathway would increase with temperature, the observed nonvolatile product + volatile product, instead of kerogen →
increase in activation energy with conversion is also reflective volatile products. The impact of this can be seen from the
of the increase in temperature at which the conversion was isothermal pyrolysis of the oil shale data at 673 K in the control
found during TGA. This observation might suggest that the experiments. When the data in Figure 3 is converted into
reported increase in activation energy with conversion (Table conversion rate (dα/dt) and plotted as a function of reaction
6) is a consequence of using TGA with dynamic heating to time (Figure 12), the isothermal conversion rate passes through
H DOI: 10.1021/acs.energyfuels.7b00578
Energy Fuels XXXX, XXX, XXX−XXX
Energy & Fuels Article

E = 1990·α 3 − 1820·α 2 + 600·α + 160 (11)


This was an approximation only, with a correlation
coefficient of r2 = 0.975, and should be considered empirical
with no fundamental basis. Alternative approaches to represent
conversion dependent activation energy can be found in the
literature, e.g., see ref 45. The same approach as in eq 11 was
followed for the pre-exponential factor. To do so, the rate
constant k was calculated for each data point in Table 6 at 673
K; using the E value determined by eq 11, for each k, the
matching k0 was calculated. The k0 values were then regressed
(eq 12), and the correlation coefficient of the regression was r2
= 0.999. The good correlation between k0 and E, both as
functions of α, is indicative of what Janković17 calls the kinetic
compensation effect.
Figure 12. Reaction rate during isothermal conversion of oil shale at ln k 0 = 230·α 3 − 200·α 2 + 63.5·α + 29.9 (12)
673 K, with t = 0 defined as the time at which 673 K was reached.
When the conversion dependence of the kinetics is
considered, the time required in order to reach a specific
level of conversion is different compared to an approximation
a maximum. This is not related to mass transport, because both
using only a single value for the rate constant. This is illustrated
size ranges of oil shale resulted in very similar reaction rates
by Figure 13, which compares the reaction time required for
based on TGA mass loss.
The initially lower reaction rate is likely due to the formation
of pyrolysis products that are not volatile and therefore not
detected. This type of pyrolysis behavior appears similar to that
observed during the thermal cracking of waxes, which develops
a bimodal carbon number distribution during pyrolysis.42 The
heavier products decrease over time with increasing conversion,
but there is an equal probability of cracking for all bonds of
similar strength. Thus, heavy products do not just form light
products. Although kerogen is not wax, chemical bonds of equal
strength will have equal probability of homolytic bond
dissociation. Hence, the probability of forming nonvolatile
products from the kerogen decreases only with increasing level
of conversion, because as conversion progresses, the fraction of
very heavy material that can form nonvolatile products
decreases.
4.6. Implications for Practical Application. The geology Figure 13. Reaction time required after reaching α = 0.15 in order to
of the oil shale formation determines whether mining and reach the conversion values indicated. The calculated time is shown for
retorting or in situ subsurface production technology should be constant k (●), conversion-dependent k (■) based on kinetic derived
used. Retorting processes operate mainly on the principle of a by the Friedman method, and the experimentally measured conversion
moving bed for the oil shale with continuous stripping of the (▲).
vapor phase. The way in which it is implemented differs
between technologies. For example, the Petrosix process uses a isothermal reaction at 673 K based on kinetic data determined
moving bed similar to that found in moving bed gasifiers; the by the Friedman method using constant values for E and k0 at α
Tosco II process uses a rotary kiln similar to that found in = 0.40 (Table 6), the conversion-dependent E and k0 values (eq
cement production.5,8,10 In situ subsurface production by 11 and 12), and the isothermal kinetics from the control
definition operate on the principle of a semibatch process, experiment (Figure 3). A temperature of 673 K was selected for
keeping the oil shale stationary and continuously removing the the comparison, because the isothermal data in Figure 3 was
produced oil and gas. The technologies differ mostly in terms of collected at that temperature; industrial processes typically
the way in which heating is supplied and subsurface operate at a higher temperature.
permeability is created.10,41,43,44 There is an observable difference in calculated reaction time
Considering the differences in potential application, the required for kerogen conversion using kinetics using a single
implication of this work will be analyzed by looking at how the value for E and k0 and that taking the conversion dependence of
kinetics describes the rate of oil and gas production by pyrolysis E and k0 into account (Figure 13). Using a constant k has
as a function of time. Although the rate equation is first-order potentially serious implications for process design aimed at high
with respect to kerogen, in practice the pyrolysis rate is not a kerogen conversion. The activation energy for conversion of
simple first-order relationship with respect to time, due to the the more refractive species in the kerogen is high, and it is these
conversion dependence of the rate constant. species that remain mostly unconverted during the initial stages
The conversion dependence of the activation energy over the of pyrolysis. In the example given (Figure 13), the additional
conversion range 0.15 ≤ α ≤ 0.65 was regressed as a third- reaction time required for incremental conversion at 673 K
order polynomial based on the results (Table 6 and Figure 11) increases rapidly as the conversion exceeds about 50%. This is
of the Friedman method (eq 11). not captured by a description based on a single value for
I DOI: 10.1021/acs.energyfuels.7b00578
Energy Fuels XXXX, XXX, XXX−XXX
Energy & Fuels Article

activation energy and a conversion-independent kinetic Notes


constant. To put this into perspective, the conversion after 3 The authors declare no competing financial interest.


h of conversion at 673 K can be calculated. The experimentally
observed conversion is 68.5%. The conversion-dependent ACKNOWLEDGMENTS
kinetics based on the Friedman method calculated a kerogen
conversion of 61.6%. Using conversion-independent kinetics, This study was funded through the program “Science Without
the calculated conversion is 89.5% after 3 h (and near complete Borders”, the University of Alberta, and the State University of
conversion is predicted after 8 h). Campinas. Mohan S. Pathak (University of Alberta) is thanked
The control experiments conducted at 673 K (Figure 3) for performing the control experiments.
indicated that reaction over a longer period of time is unlikely
to lead to near complete conversion. Employing the
simplification of a single activation energy for the practical
■ NOMENCLATURE
E = activation energy (kJ/mol)
design of an oil shale pyrolysis can lead to an over prediction of f(α) = reaction model
the kerogen conversion, especially for designs aimed at more g(α) = integrated reaction model
extensive conversion. k = reaction rate constant (units depend on reaction order)
k0 = pre-exponential factor (min−1 for first order)
5. CONCLUSIONS m = mass of oil shale (mg)
The mass loss associated with heating of oil shale could be mo = mass of oil shale before pyrolysis (mg)
divided into different temperature regions: evaporation of light mf = mass of oil shale after pyrolysis (mg)
fraction of kerogen (region I), kerogen pyrolysis (region IIa), n = reaction order
kerogen pyrolysis and catalytic cracking by mineral matter r = correlation coefficient for linear regression
(region IIb), and high-temperature decomposition of kerogen R = universal gas constant (kJ/mol·K)
and mineral matter (region III). Kinetic analysis was limited to t = time (min)
region IIa, where kerogen pyrolysis and cracked product T = temperature (K)
evaporation were the dominant processes. The main con- α = kerogen conversion as defined by eq 3 (mg/mg)
clusions from the work are provided below. β = heating rate (K/min)
(a) Devolatilization kinetics of kerogen pyrolysis was
calculated from the nonisothermal TGA measurements by the
model-free methods of Friedman, Flynn−Wall−Ozawa, and
■ REFERENCES
(1) George, R. D. Origin of oil shales. In Shale Oil; McKee, R. H.,
Kissinger−Akahira−Sunose. It was found that activation energy Ed.; ACS Monograph Ser. 25; Chemical Catalogue Company: New
depended on the extent of conversion and increased over the York, 1925.
range 215−255 kJ/mol for conversion in the range 0.15 ≤ α ≤ (2) Duncan, D. C. Geological setting of oil-shale deposits and world
0.55, where α = 1 for pyrolysis at 1173 K. prospects. In Oil shale; Yen, T. F., Chilingarian, G. V., Eds.; Elsevier:
(b) Kinetic analysis by the Coats−Redfern method using Amsterdam, 1976; pp 13−26.
first-order kinetics was insensitive to heating rate but resulted in (3) Cane, R. F. The origin and formation of oil shale. In Oil shale;
activation energies that were unrealistically low. Yen, T. F., Chilingarian, G. V., Eds.; Elsevier: Amsterdam, 1976; pp
27−60.
(c) The prediction of kinetics based on the model-free (4) Knaus, E.; Killen, J.; Biglarbigi, K.; Crawford, P. An overview of
methods was compared against the experimentally measured oil shale resources. ACS Symp. Ser. 2010, 1032, 3−20.
kinetics during isothermal pyrolysis at 673 K. The reaction rate (5) Weiss, H.-J.; Lisboa, A. C. L. Oil shale. In Ullmann’s Encyclopedia
calculated using kinetic parameters derived by the Friedman of Industrial Chemistry, Vol. 25; Wiley-VCH: Weinheim, 2012; pp
method was more accurate than those of other methods at 315−343.
lower conversion, typically 0.15 ≤ α ≤ 0.30, when the reaction (6) Boak, J. Shale-hosted hydrocarbons and hydraulic fracturing. In
rate was high. Kinetic parameters calculated by the Flynn− Future energy. Improved, sustainable and clean options for our planet, 2nd
Wall−Ozawa and Kissinger−Akahira−Sunose methods became ed; Letcher, T. M., Ed.; Elsevier: Amsterdam, 2014; pp 117−143.
more accurate at higher conversion, when reaction rate was (7) Perrini, E. M. Oil from shale and tar sands; Noyes Data
lower. Corporation: Park Ridge, NJ, 1975.
(8) Synthetic fuels data handbook; Hendrickson, T. A., Ed.; Cameron
(d) Isothermal kerogen pyrolysis approached an incomplete Engineers Inc.: Denver, CO, 1975.
conversion limit that could be increased only by increasing the (9) Office of Technology Assessment. An assessment of oil shale
temperature. This type of behavior was predicted by the technologies; McGraw-Hill: New York, 1980.
conversion dependence of activation energy. The scientific basis (10) Lee, S. Oil shale technology; CRC Press: Boca Raton, FL, 1991.
for the conversion-dependent change in activation energy could (11) Speight, J. G. Shale oil production processes; Elsevier: Amsterdam,
be explained. 2012.
(e) For process design purposes, the importance of modeling (12) Li, S.; Yue, C. Study of pyrolysis kinetics of oil shale. Fuel 2003,
oil shale pyrolysis kinetics with conversion-dependent 82, 337−342.
activation energy was highlighted. (13) Petersen, H. I.; Bojesen-Koefoed, J. A.; Mathiesen, A. Variations


in composition, petroleum potential and kinetics of ordovician −
miocene type I and type I-II source rocks (oil shales): implications for
AUTHOR INFORMATION hydrocarbon generation characteristics. J. Pet. Geol. 2010, 33, 19−42.
Corresponding Author (14) Burnham, A. K. Chemistry and kinetics of oil shale retorting.
*Phone: 780-248-1903. Fax: 780-492-2881. E-mail: deklerk@ ACS Symp. Ser. 2010, 1032, 115−134.
ualberta.ca. (15) Hillier, J. L.; Fletcher, T. H. Pyrolysis kinetics of a Green River
oil shale using a pressurized TGA. Energy Fuels 2011, 25, 232−239.
ORCID (16) Tiwari, P.; Deo, M. Compositional and kinetic analysis of oil
Arno de Klerk: 0000-0002-8146-9024 shale pyrolysis using TGA−MS. Fuel 2012, 94, 333−341.

J DOI: 10.1021/acs.energyfuels.7b00578
Energy Fuels XXXX, XXX, XXX−XXX
Energy & Fuels Article

(17) Janković, B. The kinetic modeling of the non-isothermal (38) Cleyle, P. J.; Caley, W. F.; Stewart, I.; Whiteway, S. G.
pyrolysis of Brazilian oil shale: Application of the Weibull probability Decomposition of pyrite and trapping of sulphur in a coal matrix
mixture model. J. Pet. Sci. Eng. 2013, 111, 25−36. during pyrolysis of coal. Fuel 1984, 63, 1579−1582.
(18) Bai, F.; Guo, W.; Lü, X.; Liu, Y.; Guo, M.; Li, Q.; Sun, Y. Kinetic (39) Sousa, J. J. F.; Vugman, N. V.; Neto, C. C. Free radical
study on the pyrolysis behavior of Huadian oil shale via non-isothermal transformations in the Irati oil shale due to diabase intrusion. Org.
thermogravimetric data. Fuel 2015, 146, 111−118. Geochem. 1997, 26, 183−189.
(19) Burnham, A. K. A simple kinetic model of oil generation, (40) Torrente, M. C.; Galán, M. A. Kinetics of the thermal
vaporization, coking, and cracking. Energy Fuels 2015, 29, 7156−7167. decomposition of oil shale from Puertollano (Spain). Fuel 2001, 80,
(20) Brown, M. E.; Maciejewski, M.; Vyazovkin, S.; Nomen, R.; 327−334.
Sempere, J.; Burnham, A.; Opfermann, J.; Strey, R.; Anderson, H. L.; (41) Symington, W. A.; Kaminsky, R. D.; Meurer, W. P.; Otten, G.
Kemmler, A.; Keuleers, R.; Janssens, J.; Desseyn, H. O.; Li, C.-R.; A.; Thomas, M. M.; Yeakel, J. D. ExxonMobil’s Electrofrac process for
Tang, T. B.; Roduit, B.; Malek, J.; Mitsuhashi, T. Computational in situ oil shale conversion. ACS Symp. Ser. 2010, 1032, 185−216.
(42) De Klerk, A. Thermal cracking of Fischer−Tropsch waxes. Ind.
aspects of kinetic analysis Part A: The ICTAC kinetics project-data,
Eng. Chem. Res. 2007, 46, 5516−5521.
methods and results. Thermochim. Acta 2000, 355, 125−143. (43) Burnham, A. K.; Day, R. L.; Hardy, M. P.; Wallman, P. H.
(21) Vyazovkin, S.; Burnham, A. K.; Criado, J. M.; Pérez-Maqueda, L. AMSO’s novel approach to in-situ oil shale recovery. ACS Symp. Ser.
A.; Popescu, C.; Sbirrazzuoli, N. ICTAC Kinetics Committee 2010, 1032, 149−160.
recommendations for performing kinetic computations on thermal (44) Ryan, R. C.; Fowler, T. D.; Beer, G. L.; Nair, V. Shell’s in situ
analysis data. Thermochim. Acta 2011, 520, 1−19. conversion process − from laboratory to field pilots. ACS Symp. Ser.
(22) da S. Ramos, A.; Rodrigues, L. F.; de Araujo, G. E.; Pozocco, C. 2010, 1032, 161−183.
T. M.; Ketzer, J. M. M.; Heemann, R.; Lourega, R. V. Geochemical (45) Al-Ayed, O. S.; Matouq, M.; Anbar, Z.; Khaleel, A. M.; Abu-
characterization of Irati and Palermo formations (Paraná Basin− Nameh, E. Oil shale pyrolysis kinetics and variable activation energy
Southern Brazil) for shale oil/gas exploration. Energy Technol. 2015, 3, principle. Appl. Energy 2010, 87, 1269−1272.
481−487.
(23) ASTM D7582. Standard test methods for proximate analysis of
coal and coke by macro thermogravimetric analysis; ASTM: West
Conshohocken, PA, 2012.
(24) Carter, S. D.; Robl, T. L.; Taulbee, D. N.; Rubel, A. M. Testing
of an lrati oil shale in a multi-stage fluidized bed retorting process. Fuel
1991, 70, 1347−1351.
(25) Colthup, N. B.; Daly, L. H.; Wiberley, S. E. Introduction to
infrared and Raman spectroscopy, 3rd ed; Academic Press: Boston,
1990.
(26) Dolphin, D.; Wick, A. Tabulation of infrared spectral data; Wiley-
Interscience: New York, 1977.
(27) Solomon, P. R.; Hamblen, D. G.; Carangelo, R. M.; Serio, M. A.;
Deshpande, G. V. General model of coal devolatilization. Energy Fuels
1988, 2, 405−422.
(28) Solomon, P. R.; Serio, M. A.; Suuberg, E. M. Coal pyrolysis:
Experiments, kinetic rates and mechanisms. Prog. Energy Combust. Sci.
1992, 18, 133−220.
(29) Wang, Z.; Deng, S.; Gu, Q.; Zhang, Y.; Cui, X.; Wang, H.
Pyrolysis kinetic study of Huadian oil shale, spent oil shale and their
mixtures by thermogravimetric analysis. Fuel Process. Technol. 2013,
110, 103−108.
(30) Al-Harahsheh, M.; Al-Ayed, O.; Robinson, J.; Kingman, S.; Al-
Harahsheh, A.; Tarawneh, K.; Saeid, A.; Barranco, R. Effect of
demineralization and heating rate on the pyrolysis kinetics of
Jordanian oil shales. Fuel Process. Technol. 2011, 92, 1805−1811.
(31) Sütcü, H.; Pişkin, S. Pyrolysis kinetics of oil shale from Ulukişla,
Turkey. Oil Shale 2009, 26, 491−499.
(32) Zachariah, A.; De Klerk, A. Thermal conversion regimes for
oilsands bitumen. Energy Fuels 2016, 30, 239−248.
(33) Franco, N.; Kalkreuth, W.; Peralba, M. C. R. Geochemical
characterization of solid residues, bitumen and expelled oil based on
steam pyrolysis experiments from Irati oil shale, Brazil: A preliminary
study. Fuel 2010, 89, 1863−1871.
(34) Ambles, A.; Baudet, N.; Jacquesy, J.-C. Structural study of the
kerogen from Brazilian Irati oil shale by selective degradations.
Tetrahedron Lett. 1993, 34, 1783−1786.
(35) Rimkevich, V. S.; Pushkin, A. A.; Girenko, I. V. Synthesis and
properties of amorphous SiO2 nanoparticles. Inorg. Mater. 2012, 48,
355−360.
(36) Lawson, K. E. Infrared absorption of inorganic substances;
Reinhold: New York, 1961.
(37) Lennie, A. R.; Vaughan, D. J. Kinetics of the marcasite-pyrite
transformation: An infrared spectroscopic study. Am. Mineral. 1992,
77, 1166−1171.

K DOI: 10.1021/acs.energyfuels.7b00578
Energy Fuels XXXX, XXX, XXX−XXX

You might also like