You are on page 1of 7

ARTICLE

pubs.acs.org/IECR

Kinetic Models for Upgrading Athabasca Bitumen Using Unsupported


NiWMo Catalysts at Low Severity Conditions
Carmen E. Galarraga,* Carlos Scott, Herbert Loria, and Pedro Pereira-Almao
Department of Chemical and Petroleum Engineering, Schulich School of Engineering, University of Calgary, Calgary, AB T2N 1N4,
Canada
ABSTRACT: Typically, the catalytic upgrading of heavy fractions (VGO and VR) has been studied using mostly conventional
supported catalysts at temperatures and pressures higher than 400 C and 6 MPa, respectively. This work focuses on the upgrading
of heavy oils at much lower severity conditions using dispersed NiWMo catalysts for processing whole oil with no fractionation. A
kinetic study was developed to determine parameters from experimental data obtained at temperatures of 320380 C and reaction
times from 3 to 70 h at a total hydrogen pressure of 3.45 MPa and a stirring speed of 500 rpm in a batch reactor. The conversion,
estimated as the reduction of the residue 545 C+ fraction, was tted for a rst-order reaction with an apparent activation energy of
200 kJ mol1. Two kinetic models are proposed to predict the conversion of the residue fraction and its product distribution.
Comparison between experimental data and predictions using the proposed models exhibited good agreement with average
absolute errors lower than 5%.

temperatures and high pressures of hydrogen.7,8 Generally speaking, it has been accepted that the upgrading mechanism for heavy
feedstocks, such as heavy oils and bitumen, is similar to thermal
cracking,7,9 but having hydrogen transfer/hydrogenation superimposed, which helps to improve the quality of the upgraded
product while decreasing the coke production.10 It would be very
benecial to nd an alternative way of treating these feeds in situ,
in the reservoir, to generate an upgraded low-viscosity bitumen
that would eliminate the need for addition of diluent when transported. Also, the removal of contaminants and heteroatoms
would help to decrease the hydrogen requirements for treatment
during the renement process.
Regarding catalysts improvement, heavy oil upgrading technologies have advanced by using dispersed unsupported catalysts.7,9,1113 Dispersed catalysts can be produced from oil
soluble precursors11,14 as well as water-in-oil emulsions containing the metallic precursors in the water droplets.15,16 The most used
formulations for these developments continue to be those formulations typically used for conventional supported catalysts, such
as molybdenum, tungsten, cobalt, nickel, and mixtures thereof.11,1719
Le Perchec and co-workers have indicated that dispersed unsupported catalysts may deactivate slower than typical supported
hydroprocessing catalysts due to the higher reaction rates exhibited by unsupported catalysts, which not only can locally generate hydrogen spillover, thus increasing the upgrading of residual
feedstocks,20 but also activate the hydrogen required for stabilizing the free radicals normally produced in the thermal cracking of
heavy petroleum feeds.9 The interest in dispersed catalysts is
mainly that they could be included in the reaction medium to
navigate along with the targeted feed as small catalytic particles.15

1. INTRODUCTION
According to the Energy Information Administration (EIA),
the world marketed energy consumption will continue its tendency to growth; from 500 quadrillion British Thermal Units (BTU)
marketed in 2007, the forecast predicts a consumption of 700
quadrillion BTU by 2035.1 Moreover, the International Energy
Agency (IEA) has estimated that the oil demand will keep rising
from the current 85 million barrels per day (MBBL/d) to 106
MBBL/d by 2030.2,3 Traditionally, these demands have been
covered by exploitation of conventional light crude oils due to
their easiness of production; however, this practice has generated
a depletion in light oils worldwide. Therefore, the production of
energy from hydrocarbons is an important issue for Canada because of its large amount of proven oil reserves, which are estimated to be 178.8 billion barrels of crude oil. Nevertheless, most
of these reserves (about 174.5 billion barrels) are bitumen contained in oil sands, which are a complex mixture of sand, water,
clay, and bitumen.4,5
Bitumen is a type of hydrocarbon, not only very viscous (its
transportation inside pipelines usually requires diluent addition),
but also with a low American Petroleum Institute (API) gravity
value (low marketing value).6 Additionally, bitumen is rich in
asphaltenes (1030 wt %), aromatics, and other multiple contaminants, thus requiring intense treatment to produce high valuable cleaner fuels able to comply the recent environmental constraints. One important characteristic of heavy oils and bitumen
is their low hydrogen-to-carbon ratio,; thus their processing is
preferably performed via hydrogen addition instead of carbon
rejection processes. Conventional upgrading of this type of
feedstocks has been achieved in petroleum reneries by mildhydrocracking or hydroconversion from moderate to high temperatures under hydrogen pressure aiming to produce light highquality fuels by increasing their low hydrogen-to-carbon ratio and
diminishing the undesirable contaminants such as sulfur, nitrogen, aromatics, etc. Thus, such treatment usually requires high
r 2011 American Chemical Society

Received: June 5, 2011


Accepted: December 7, 2011
Revised:
December 2, 2011
Published: December 07, 2011
140

dx.doi.org/10.1021/ie201202b | Ind. Eng. Chem. Res. 2012, 51, 140146

Industrial & Engineering Chemistry Research

ARTICLE

The kinetic study of this type of reactions is very important


because the assessment of the reaction rate is a key activity for
selection of optimum operating parameters, moreover, in this
particular case, because not many kinetic studies have been presented at working conditions close to those used for in-reservoir
operation. These studies are required to establish sound models
for reaction simulation, allowing the prediction of both bitumen
upgrading and product quality. When dealing with heavy oils and
bitumen upgrading, one important step of research for reactor
design is the production of detailed kinetic models that allow for
precise estimation of product streams composition. However,
because heavy oils and bitumen are complex structures composed by large amounts of resins and asphaltenes, which consist
of large molecules of condensed poly aromatic rings, the kinetic
study of its reactivity is a complex task. One typical approach
to this problem involves the use of grouped lumps of pseudospecies,2124 connected through a reaction scheme by parameter
estimation and correlated with process data yield. The larger is
the number of lumps, the longer is the time for solving the intricate equations network from the kinetic analysis, because more
kinetic parameters need to be estimated and more experimental
data need to be gathered. Regarding the reaction order for these
type of reactions, there are some discrepancies; while some authors
consider this reaction to be second order,25,26 most researchers
agree on a power law of a rst-order reaction with respect to the
hydrocarbon.2731
Reported kinetic models for the catalytic and noncatalytic
upgrading of Athabasca bitumen used several levels of complexity
for which the lumps include a combination of the separated
fractions: heavy ends (coke + asphaltenes + resins), light oils
(saturates + aromatics), asphaltics (coke + asphaltenes), maltenes
(resins + aromatics + saturates), and gases.27,32,33 The increased
number of lumps rendered the best results; however, good approximations were obtained with intermediate numbers (24
lumps).8,34,35 Another way to the traditional lumping technique
is modeling based on distillation fractions, such as naphtha,
middle distillates, vacuum gas oil (VGO), and residue.28 Also,
dierent reactions schemes have been proposed in the literature,
in which it has been concluded that given the complexity of real
feedstocks, lumping techniques will continue as useful tools for
studying hydroprocessing reaction kinetics.21,36 With respect to
the use of ultradispersed catalysts for bitumen and heavy oil
upgrading, most of the published work presents kinetic models
at typical (high severity) hydroprocessing conditions, which is
temperatures from 350 C and above with minimum reacting
pressures of 7 MPa.7,10 One important goal of the present research is to evaluate the performance of these catalysts at operating conditions near those used in-reservoir operation, because
typical temperatures and pressures in petroleum reservoirs are
lower than those employed in conventional hydroprocessing.
Hence, in this Article, we present kinetic models to describe
the eects of temperature and reaction time on the extent of
conversion of the residue fraction of Athabasca bitumen when
hydroprocessed in the presence of unsupported dispersed NiWMo catalysts using a noncomplex stirred batch reactor at low
severity operating conditions.

in the hydrocarbon. Conventional hydroprocessing catalysts are


bimetallic17 and mainly constituted by CoMo (maximum hydrogenolysis and hydrodesulfurization),37 whereas Ni and W are
included when higher hydrogenation activity is desired.18,38 The
catalyst for the present study was formulated as a mixture of
NiWMo to maximize hydrogenating activity. Thus, 1000 ppmw
of metallic species (respect to the bitumen) was added to produce atomic metallic ratios as follows: Ni/Me(atomic) = 0.3 with
Me = Ni + W + Mo, and Mo/W(atomic) = 3; additional details
regarding this catalytic formulation and its preparation are described elsewhere.12,13
2.2. Feedstock. Athabasca bitumen produced via steamassisted gravity drainage (SAGD) with an API gravity of 9.5,
viscosity of 7890 cP at 40 C, sulfur of 4.8 wt %, and a residue
content of 49 wt % (545 C+) was used as feedstock for the upgrading experiments.
2.3. Upgrading Experiments. In a typical experiment, about
30 g of catalytic emulsion was placed into a Parr batch autoclave
reactor (100 mL), where upgrading reactions are performed at
temperatures from 320 to 380 C, a total hydrogen pressure of
3.45 MPa, for various reaction times (370 h). All experiments
were carried out at a stirring speed of 500 rpm to diminish mass
transfer limitations. Zero time was taken when the temperature
inside the reactor reached the targeted value. Product samples are
constituted by hydrocracked products (gases and liquids), catalytic submicrometer particles, and coke (if produced). Gaseous
samples are collected in a vacuum-cleaned cylinder and analyzed
by gas chromatography (GC) to determine both hydrogen and
light hydrocarbons (C1C5) contents; thus, at the sample
conditions an average molecular weight can be estimated and
used to quantify the sample mass and to calculate the gas yield.
Liquid (feedstock or product) samples were analyzed using standard characterization techniques to determine viscosity and sulfur
content, whereas boiling point distribution curves were determined by high temperature simulated distillation (HTSD),39
which served to define four pseudocomponents distinguished by
boiling points: naphtha (IBP216 C), distillates (216343 C),
vacuum gasoil, VGO (343545 C), and residue (545 C +).
Detailed procedures for collection of both samples gases and
liquids are described elsewhere.12,13
Coke was dened as the dierence between the total solids
recovered (nonsoluble matter after contacting the liquid sample
with CHCl3 in a ratio 1:50 w/w) minus the nominal mass of catalyst component (0.15 wt %). Equation 1 was used to calculate
the conversion of the residue fraction, whereas complete mass
balances (average value 97.8 wt %), composition and quantication of gases, as well as coke determination were used to calculate
the entire product distribution. The liquid yield then is calculated
as 100 minus both gases and coke yield altogether.

2. EXPERIMENTAL SECTION

3.1. Product Composition. The liquid product yield distribution according to the aforementioned fractions together with
gases and coke for experiments carried out at temperatures
320380 C is listed in Table 1. As expected, the higher is the
temperature and the longer is the reaction time, the higher is the

conv 545C

mass 545 C
feed  mass 545 Cproduct

mass 545 C
feed

 100
1

3. RESULTS AND DISCUSSION

2.1. Catalyst Preparation. NiWMo catalysts in the form of


submicrometer dispersed species were prepared from aqueous
solutions of nickel, tungsten, and molybdenum that were emulsified
141

dx.doi.org/10.1021/ie201202b |Ind. Eng. Chem. Res. 2012, 51, 140146

Industrial & Engineering Chemistry Research

ARTICLE

Table 1. Product Yield Distribution (wt %) from HCK of


Athabasca Bitumen Catalyzed by Submicrometer NiWMo
Catalysts at a Total Pressure of 3.45 MPa, Temperatures
320380 C, and Reaction Times 370 h
product distribution, %
temperature, C time, h gases naphtha distillates VGO residue coke
feedstock

2.8

15.0

34.5

47.8

320

24

0.6

2.7

15.6

35.8

45.1

0.3

320

38

0.4

3.0

16.1

35.8

44.4

0.3

320

48

0.3

4.2

17.0

35.8

42.6

0.1

320

69

0.1

3.3

17.7

37.7

41.0

0.2

350

0.1

4.6

17.0

35.3

42.9

0.2

350
350

7
22

0.3
0.8

3.7
4.7

17.1
20.3

36.1
38.9

42.6
35.0

0.2
0.2

350

30

1.8

5.8

21.8

39.6

30.7

0.3

350

30

1.8

5.8

21.3

39.8

29.5

0.3

350

30

1.8

5.8

21.8

40.2

30.5

0.3

350

48

2.2

7.9

22.7

38.7

28.1

0.4

360

1.0

3.3

17.7

35.0

42.8

0.3

360

15

1.8

6.2

19.8

35.4

36.6

0.3

365
365

5
5.5

0.5
0.6

4.8
4.4

19.4
19.2

36.6
37.1

38.7
38.5

0.0
0.3

380

2.2

6.7

20.3

38.0

32.6

0.2

380

5.8

2.0

8.8

23.9

39.1

25.9

0.3

380

2.0

10.9

27.1

40.2

19.5

0.4

380

14

2.0

14.6

29.4

36.4

16.9

0.8

Figure 1. Rate constant plots for the hydrocracking of whole Athabasca


bitumen catalyzed by submicrometer NiWMo catalysts at a total pressure of 3.45 MPa and a stirring speed of 500 rpm.

Table 2. Kinetic Constants for a First-Order HCK Reaction


of Whole Athabasca Bitumen Using Submicrometer NiWMo
Catalysts
temperature, C kinetic constant, h1 regression coecient, r2 AAE, %

conversion of the residue fraction into lighter products. At the


lowest temperature (320 C), the production of gases and coke is
almost negligible. At the highest temperature (380 C), the
scattered pattern observed for gases yield is probably due to
experimental error, because one would expected that the gases
yield increases with temperature.40 A similar behavior has been
reported for the conversion of Maya heavy oil.28 The content of
residue 545 C+ decreased with both temperature and reaction
time even at the lowest temperature of 320 C. At moderate to
high temperatures (350380 C), the majority of the residue
fraction produces VGO, distillates, naphtha, and gases. The
production of coke increases with severity, as expected;
however, because of the mild conditions used in this work,
selected to be near those of the in-reservoir operation, this
undesired product was kept at a minimum. In a previous work,
it was demonstrated that including these catalytic species
at 380 C and 8 h the coke amount was decreased from 8 to
0.2 wt %.12 Reproducibility of these experiments was evaluated by performing three different runs at 350 C, 3.45 MPa
for 30 h. These results are also included in Table 1. A good
reproducibility was found with an average absolute error less
than 3 wt % for all of the fractions.
3.2. Kinetic Models. As already mentioned, the cracking
reaction of bitumen and its fractions is considered to be first
order with respect to hydrocarbons28 and zero order with respect
to hydrogen.25 Also, it is reported that the cracking of hydrocarbons is a bimolecular reaction taking place on the catalytic
surface,25 in which both reactants should adsorbed on adjacent
surface sites. Thus, given the complexity of the reacting system,
the mathematical model for this work adopted a power law
simple model.

320

0.0022

0.98

0.53

350

0.0127

0.95

1.56

360

0.0178

0.99

0.08

365

0.0418

0.99

0.44

380

0.1108

0.99

2.32

average AAE for all data, %

1.21

The rst approach in this kinetic evaluation, so-called model A,


considers the conversion of the residue fraction into lighter products as an irreversible rst-order reaction, as follows:
residue 545 C bitumen H2 f lighter products

with the following rate equation:


dCR
 kCR
dt

where CR is the concentration of fraction 545 C+, t is the reaction time (h), and k is the rate constant (h1).
The linear form of eq 3 (using values from Table 1) is plotted
in Figure 1. The linearity of these plots allows one to conclude
that a rst-order reaction describes quite well the conversion of
the residue fraction of Athabasca bitumen when hydroprocessing it
in the presence of the catalysts prepared from catalytic emulsions.
The calculated rate constants are included in Table 2 along
with the data correlation indexes, r2, and the average absolute
errors (AAE), which were obtained from eq 4 by comparing the
experimental value, CR exp (listed in Table 1), with the calculated
concentration of residue, CR cal (estimated using the kinetic
constant, slope of the tting line, at each temperature) at the ith
reaction time for n total number of concentrations. As expected
for a cracking reaction, the kinetic parameter, k, increased with
temperature. The AAE for the calculated kinetic constants resulted as less than 3%. These results agree with previous literature
data obtained for supported catalysts and operating conditions
typical of commercial reactors.25,28
142

dx.doi.org/10.1021/ie201202b |Ind. Eng. Chem. Res. 2012, 51, 140146

Industrial & Engineering Chemistry Research

ARTICLE

Figure 2. Arrhenius plot for model A.

Figure 3. Fractional conversion of the residue fraction 545 C+ as a


function of reaction time at temperatures 320380 C. Symbols,
experimental data; lines, calculated data from model A.

Additionally, in the present case, it is important to mention


that at low severity conditions, that is, reaction temperature e
365 C, the AAE was smaller than 1%. The highest error (>2%)
was observed for the highest temperature (380 C); this increased error might be due to the high level of conversion
achieved at long reaction times for this temperature, which in
consequence might produce a higher amount of coke and gases.
Also, it has been reported that at higher severity conditions a
second-order reaction seems to t better than the rst-order typically assumed for this reaction.26
n

AAE%

i1

jCR i exp  CR i calj


100
CR i exp

!
4

The Arrhenius plot (ln k vs 1/T) to determine the apparent


activation energy (Ea) for the data included in Table 2 is displayed in Figure 2. As a general result, the average activation
energy for this range of temperatures was found to be 204 kJ mol1
with a correlation index, r2 = 0.98, which is in close agreement
with values already reported in the literature for this type of
reacting system.7,9 It has been mentioned that in these processes,
the catalysts do not have catalytic activity toward the cracking
reactions.9 Thus, it can be assumed that the apparent activation
energy observed in these experiments is related to a noncatalytic
hydrocracking reaction, that is, thermal cracking with excess of
hydrogen.
It is worth mentioning that as the temperature decreases the
thermal cracking of the CC bond reaction may be inhibited,
but the hydrogenating reactions will get promoted because
they require lower energy as compared to the thermal cracking ones.41 In this case, the dispersed catalysts are eectively
catalyzing the hydrogenating reactions, and therefore helping
to synthesize a product with a better quality exhibiting not only
an improved viscosity and API gravity but also a lesser amount
of solids, as previously reported.12 These improved product
characteristics not only would help to decrease the amount
of solvent required for bitumen transportation (from remote
areas to rening facilities) but also would promote product
stability.
The apparent kinetic parameters from model A were then
used to estimate the theoretical conversion at temperatures
from 320 to 380 C as a function of reaction time and compared to the conversion from the experimental data (Table 1).
The results are displayed in Figure 3. As it can be observed,

Figure 4. Proposed lumped kinetic model for the hydrocracking


reaction of Athabasca bitumen using ultradispersed catalysts at conditions near in-reservoir operation.

the model represents very well the experimental residue conversion in the interval of operational conditions here explored. This model, however, is not sucient to further
predict the composition and quality of the obtained products.
Therefore, a more complex approach is required, and thus the
lumped technique (based on Sanchez et al.28 ) was adopted
to propose a second model, model B, which is exhibited in
Figure 4.
The base for model B includes 10 kinetic parameters (k1, ..., k10).
Considering the same assumptions as for model A, eq 3 is
adapted for each pseudo component j, with j = residue, VGO,
distillates, naphtha, and gases, and thus the following equations
can be written:
R
residue : r R  k1 k2 k3 k4 C
143

dx.doi.org/10.1021/ie201202b |Ind. Eng. Chem. Res. 2012, 51, 140146

Industrial & Engineering Chemistry Research

ARTICLE

Table 3. Kinetic Parameters for Hydrocracking of Athabasca Bitumen Using Submicronic NiWMo Catalysts As Described by
Model B
temperature
1

kinetic constant (h )

320 C

350 C

k1

1.18  103
4

380 C

r2

activation energy Ea (kcal/mol)

6.81  103

4.48  102

3

0.983

172.1

k2

1.51  10

4.49  10

2.68  102

0.978

276.7

k3
k4

1.79  104
3.62  105

2.49  103
5.09  104

2.87  102
1.06  102

1.000
0.995

271.7
303.1

k5

9.78  104

2.10  103

1.78  102

0.916

157.0

6

4

0.965

342.9

0.928

242.0

k6

3.81  10

3.13  10

2.26  103

k7

3.62  108

5.19  104

k8

3.19  106

5.86  104

2.86  103

4

k9

1.54  10

k10

1.24  104

2.02  103

 VGO
 R  k5 k6 k7 C
VGO : r VGO k1 C

D
 R k5 C
 VGO  k8 k9 C
distillates : r D k2 C

 R k6 C
 VGO k8 C
 D  k10 C
N
naphtha : r N k3 C

 R k7 C
 VGO k9 C
 D k10 C
N
gases : r G k4 C

j
C

Cj
1  xj
Coj

were controlled by the inclusion of the hydrogenating phases via


the catalytic emulsion. (b) The rst four kinetic rate constants
(k1k4) represent the global (apparent) kinetic rate constant
(k from model A); therefore, the sum of these rate constants was
set to be equal to the global k already found from model A at each
temperature (k1 + k2 + k3 + k4 = k). (c) The kinetic rate constants
must follow the Arrhenius temperature dependence; then the
values of the kinetic rate constants at a specic temperature were
set to be higher than those from lower temperatures (kn at T2 > kn
at T1, where T2 > T1). Results of these calculations are shown in
Table 3, which also includes the activation energies determined from
Arrhenius law as well as the corresponding correlation indexes.
It should be mentioned that it was not possible to calculate the
activation energies for reactions 7, 9, and 10 because the estimation gave scattered values for these parameters, some of them
resulting in zero value. A similar trend was observed before.28
Because reactions 7, 9, and 10 are for gases production, the reason for this uncertainty might be either some experimental error
in the gases quantication or that those reactions do not proceed
in the same extent as the temperature changes. Additional experiments will be required to clarify this uncertainty. All of the regression indexes were found to be higher than 0.9, thus indicating
good condence for this set of data. The activation energies range
from 157 to 342 kJ mol1. The least energy-demanding reactions
were found to be VGO to distillates and residue to VGO, which
indicates that these reactions are feasible at the lowest temperatures here evaluated. It also conrms that naphtha and middle
distillates are essentially nonreactive at the present conditions.8
Figure 5 shows the parity plots for calculated and experimental
compositions for products collected at 380, 350, and 320 C. As
observed, the data t very well with a correlation index higher than
0.999. At these temperatures, the scattered pattern of the residuals
(calculated as the dierence between the experimental and the calculated composition) plotted in Figure 6 against the experimental
composition permits one to conclude that any error in these calculations may be of experimental nature and not of the model.
The so-developed model B along with empirical correlations
(conversion-quality) already proposed12 for the liquids products
whose characterization produced the set of data presented in
Table 1 can be used as a valuable tool to predict not only the
extent of residue conversion but also the product distribution and
the quality of the converted product at any operating condition
within the range here evaluated: 320380 C and 370 h of
reacting time.

10

 j is the dimensionless concenwhere rj is the rate of reaction; C


tration relating the initial and actual concentration of component
j at any reaction time t; xj is the fractional conversion of
component j; and k1,...,10 is the kinetic constant for each reaction,
as depicted in Figure 4.
Kinetic rate constants, k1k10, were calculated by means of a
semilinearization method. The rst step was to represent reaction rate equations by integral equations dened in time intervals
corresponding to dierent residence times; these integrals were
solved numerically for each time interval employing the trapezoidal rule. The integral equations produced an overdetermined
system of equations where the variables were represented by the
kinetic rate constants. The integral equations system was implemented in Microsoft EXCEL, and initial guesses for the kinetic
rate constants were selected randomly. The kinetic rate constants
were calculated employing the tool SOLVER from Microsoft
EXCEL; the objective of this tool was to obtain the group of
kinetic rate constants that provided the lower performance index
(PI). The lower is the PI, the closest are the experimental (Cj,iexp)
and model (Cj,imod) products concentrations. The performance
index is dened as:
PI

i j 1 jCj, i exp  Cj, i mod j2

11

where Cj,i represents the concentration of the product j at the ith


evaluated residence time and m is the total number of evaluated
residence times. The best t from the tool SOLVER from EXCEL
was subjected to the following constraints: (a) Only irreversible
reactions are taken into account; thus all kinetic rate constants
were set to be positive (kn > 0). From previous results,12,13 it was
conrmed that condensation reactions (which may produce coke)
144

dx.doi.org/10.1021/ie201202b |Ind. Eng. Chem. Res. 2012, 51, 140146

Industrial & Engineering Chemistry Research

ARTICLE

4. CONCLUSIONS
A kinetic study to describe the extent of upgrading and the
product distribution obtained during the hydrocracking reaction
of Athabasca bitumen in the presence of ultradispersed submicrometer catalysts at conditions similar to those used in-reservoir
operation was successfully performed. A model based on the
boiling points lumps and including ve pseudo components,
residue, vacuum gas oil, distillates, naphtha, and gases, serves to
predict the product composition with less than 5% error.
At moderate to low severity conditions (temperatures lower
than 380 C), the average apparent activation energy was found
to be 200 kJ mol1.
AUTHOR INFORMATION
Corresponding Author
Figure 5. Parity plots for experimental and calculated values of each
component produced by the hydrocracking reaction of Athabasca
bitumen using ultradispersed catalysts at conditions near in-reservoir
operation (T = 320380 C).

*Tel.: (403) 210-9590. Fax: (403) 210-3973. E-mail: cegalarr@


ucalgary.ca.

ACKNOWLEDGMENT
We would like to thank the Alberta Ingenuity Centre for In
Situ Energy (AICISE) funded by the Alberta Ingenuity Fund
and the industrial sponsors Shell International, Conoco-Phillips,
Nexen Inc., Total Canada, and Repsol-YPF for nancial support.
Valuable comments from three anonymous referees are acknowledged. C.E.G. appreciates the economical support granted by
The Schulich School of Engineering at the University of Calgary,
Canada.
REFERENCES
(1) U.S. Energy Information Administration. International Energy
Outlook 2010. World Energy Demand and Economic Outlook, Report #
DOE/EIA 0484 (2010); retrieved August 25, 2010; http://www.eia.doe.
gov/oiaf/ieo/world.html.
(2) U.S. Energy Information Administration. World Proved Reserves
of Oil and Natural Gas 2009; retrieved February 15, 2010; http://www.
eia.doe.gov/emeu/international/reserves.html.
(3) Shah, A.; Fishwick, R.; Wood, J.; Leeke, G.; Rigby, S.; Greaves,
M. A review of novel techniques for heavy oil and bitumen extraction and
upgrading. Energy Environ. Sci. 2010, 3, 700.
(4) Patel, S. Canadian oil sands: Opportunities, technologies and
challenges. Hydrocarbon Process. 2007, 6573.
(5) Stelmach, E. Speech at the II World Heavy Oil Congress Edmonton,
Alberta 2008; available via the Internet at http://www.premier.alberta.
ca/speeches/speeches-2008-mar-10-World_Oil.cfm (accessed March
2009).
(6) Strausz, O. P.; Lown, E. M. The Chemistry of Alberta Oil Sands,
Bitumens, and Heavy Oils; Alberta Energy Research Institute: Calgary,
2003.
(7) Del Bianco, A.; Panariti, N.; Anelli, M.; Beltrame, P. L.; Carniti, P.
Thermal cracking of petroleum residues 1. Kinetic analysis of the
reaction. Fuel 1993, 72, 75.
(8) Ayasse, A. R.; Nagaishi, H.; Chan, E. W.; Gray, M. R. Lumped
kinetics of hydrocracking of bitumen. Fuel 1997, 76, 1025.
(9) Del Bianco, A.; Panariti, N.; Di Carlo, S.; Beltrame, P. L.; Carniti, P.
New developments in deep hydroconversion of heavy oil residues with
dispersed catalysts. Part 2: Kinetic aspects of the reaction. Energy Fuels
1994, 8, 593.
(10) Mohanty, S.; Kunzru, D.; Saraf, D. N. Hydrocracking a review.
Fuel 1990, 69, 1467.
(11) Panariti, N.; Del Bianco, A.; del Piero, G.; Marchionna, M.;
Carniti, P. Petroleum residue upgrading with dispersed catalysts Part 2.
Eect of operating conditions. Appl. Catal., A 2000, 204, 215.

Figure 6. Residual values from comparison in Figure 5.

Additional contributions from this research group will


address the scaling up of the present kinetic study but approaching a plug ow model (bench scale evaluation) in
which mass transfer limitations may be considered. Also,
research at temperatures lower than those explored in this
work is required to further evidence the hydrogenating catalytic eect of the dispersed NiWMo particles here employed. Estimating the extent of dispersion (or aggregation)
of these unsupported catalytic species during reaction is a
paramount activity mainly due to (a) the lack of transparency
of the oil medium (bitumen); (b) the change of properties
(viscosity and density) of the oil where the particles are dispersed; and (c) the sample recovery methodology after reaction, which requires additional manipulation techniques, such as
ltration, centrifugation, etc., and may alter the initial aggregation
state of these unsupported particles. Nevertheless, future work
attempting to study the eect of dispersion will be addressed
by contrasting the use of the catalytic emulsions alone versus
the use of a mixture containing sand particles along with the
catalytic emulsion as to simulate the catalytic behavior inside a
porous medium.
145

dx.doi.org/10.1021/ie201202b |Ind. Eng. Chem. Res. 2012, 51, 140146

Industrial & Engineering Chemistry Research

ARTICLE

(36) Ancheyta, J.; Sanchez, S.; Rodriguez, M. A. Kinetic modeling of


hydrocracking of heavy oil fractions: A review. Catal. Today 2005,
109, 76.
(37) Scheer, B.; van Oers, E. M.; Arnoldy, P.; de Beer, V. H. J.;
Moulijn, J. A. Suldability and HDS Activity of Co-Mo/Al2O3 Catalysts.
Appl. Catal. 1985, 25, 303.
(38) Sattereld, C. N.; Modell, M.; Mayer, J. F. Interactions between
catalytic hydrodesulfurization of thiophene and hydrodenitrogenation
of pyridine. AIChE J. 1975, 21, 1100.
(39) Carbognani, L.; Lubkowitz, J.; Gonzalez, M. F.; Pereira-Almao,
P. High temperature simulated distillation of Athabasca vacuum residue
fractions. Bimodal distributions and evidence for secondary on-column
cracking of heavy hydrocarbons. Energy Fuels 2007, 21, 2831.
(40) Koseoglu, R. O.; Phillips, C. R. Hydrocracking of Athabasca
bitumen: Kinetics of formation of gases. Fuel 1988, 67, 552.
(41) Demirel, B.; Wiser, W. H. Thermodynamic probability of the
conversion of multiring aromatics to isoparans and cycloparans. Fuel
Process. Technol. 1998, 55, 83.

(12) Galarraga, C. E.; Pereira-Almao, P. Hydrocracking of Athabasca


bitumen using submicronic multimetallic catalysts at near in-reservoir
conditions. Energy Fuels 2010, 24, 2383.
(13) Galarraga, C. E. Upgrading Athabasca bitumen using submicronic NiWMo Catalysts at conditions near to in-reservoir operation.
Ph.D. Dissertation, University of Calgary, Calgary, Canada, 2011; 359
pages.
(14) Bearden, R.; Aldridge, C. L.; Mayer, F.; Taylor, J.; Lewis, W.
Hydroconversion process using a sulde molybdenum catalyst concentrate. U.S. Patent No. 4,740,295, 1988.
(15) Contreras, C. Development of a new technology for preparing
nanometric Ni, Mo and NiMo catalytic particles from transient emulsions.
M.Sc. Thesis, University of Calgary, Calgary, Canada, 2009; 118 pages.
(16) Pereira, P. Fine tuning conventional hydrocarbon characterization to highlight catalytic upgrading pathways. Paper presented at the
AICISE Variability of the Oils Sands Resource Workshop; Ban Alberta:
Canada, 2007.
(17) Gates, B.; Katzer, J.; Schuit, G. C. A. Chemistry of Catalytic
Processes; McGraw-Hill, Inc.: New York, 1979.
(18) Heinrich, G.; Kasztelan, S. Hydrotreating. In Petroleum Rening.
3 Conversion Processes; Leprince, P., Ed.; Institut Francais du Petrole
Publications: Paris, France, 2001; 670 pages.
(19) Navarro, R. M.; Castano, P.; Alvarez-Galvan, M. C.; Pawelec, B.
Hydrodesulfurization of dibenzothiophene and a SRGO on sulde
Ni(Co)Mo/Al2O3 catalysts. Eect of Ru and Pd promotion. Catal.
Today 2009, 143, 108.
(20) Le Perchec, P.; Fixari, B.; Elmouchnino, J.; Peureux, S.; Vrinat,
M.; Morel, F. New developments in deep hydroconversion of heavy oil
residues with dispersed catalysts. Part 1: Thermocatalytic analysis of the
transformation with various catalysts precursors. Prepr. - Am. Chem. Soc.,
Div. Pet. Chem. 1993, 38, 401.
(21) Gray, M. R. Lumped kinetics of structural groups: hydrotreating of heavy distillate. Ind. Eng. Chem. Res. 1990, 29, 505.
(22) Martens, G. G.; Marin, G. B. Kinetics for hydrocracking based
on structural classes: Model development and application. AIChE J.
2001, 47, 1607.
(23) Susnow, R. G.; Dean, A. M.; Green, W. H.; Peczak, P.; Broadbelt,
L. J. Rate-based construction of kinetic models for complex systems.
J. Phys. Chem. A 1997, 101, 3731.
(24) Singh, J.; Kumar, M. M.; Saxena, A. K.; Kumar, S. Reaction
pathways and product yields in mild thermal cracking of vacuum
residues: A multi-lump kinetic model. Chem. Eng. J. 2005, 108, 239.
(25) Fumoto, E.; Matsumara, A.; Sato, S.; Takanohashi, T. Oxidative
cracking of residual oil with iron oxide catalyst in a steam atmosphere.
Prepr. - Am. Chem. Soc., Div. Pet. Chem. 2010, 50, 50.
(26) Sanchez, S.; Ancheyta, J. Eect of pressure on the kinetics of
moderate hydrocracking of maya crude oil. Energy Fuels 2007, 21, 653.
(27) Koseoglu, R. O.; Phillips, C. R. Kinetics and product yield
distributions in the CoO-MoO3/Al2O3 catalysed hydrocracking of
Athabasca bitumen. Fuel 1988, 67, 1411.
(28) Sanchez, S.; Rodriguez, M. A.; Ancheyta, J. Kinetic model for
moderate hydrocracking of heavy oils. Ind. Eng. Chem. Res. 2005, 44, 9409.
(29) Qader, S. A.; Hill, G. R. Catalytic hydrocracking mechanism of
hydrocracking of low temperature coal tar. Ind. Eng. Chem. Process Des.
Dev. 1969, 8, 456.
(30) Qader, S. A.; Hill, G. R. Hydrocracking of gas oil. Ind. Eng.
Chem. Process Des. Dev. 1969, 8, 98.
(31) Qader, S. A.; Hill, G. R. Hydrocracking of petroleum and coal.
Ind. Eng. Chem. Process Des. Dev. 1969, 8, 462.
(32) Koseoglu, R. O.; Phillips, C. R. Kinetics of non-catalytic
hydrocracking of Athabasca bitumen. Fuel 1987, 66, 741.
(33) Koseoglu, R. O.; Phillips, C. R. Eect of reaction variables on
the catalytic hydrocracking of Athabasca bitumen. Fuel 1988, 67, 1201.
(34) Meng, X.; Xu, C.; Li, J. G. Catalytic pyrolysis of heavy oils:
8-lump kinetic model. Appl. Catal., A 2006, 301, 32.
(35) Mosby, J. F.; Buttke, R. D.; Cox, J. A.; Nikolaides, C. Process
characterization of expanded-bed reactors in series. Chem. Eng. Sci. 1986,
41, 989.
146

dx.doi.org/10.1021/ie201202b |Ind. Eng. Chem. Res. 2012, 51, 140146

You might also like