You are on page 1of 9

ARTICLE

Renewable Diesel Production from the Hydrotreating of Rapeseed


Oil with Pt/Zeolite and NiMo/Al2O3 Catalysts
,† ‡ ‡
Rogelio Sotelo-Boyas,* Yanyong Liu, and Tomoaki Minowa

Instituto Politecnico Nacional, Mexico ESIQIE, Mexico D.F. 07738, Mexico

Biomass Technology Research Center-National Institute of Advanced Industrial Science and
Technology, Chugoku, 2-2-2 Hirosuehiro, Kure, Hiroshima 737-019, Japan

ABSTRACT: As an alternative way to produce diesel hydrocarbons, the hydrocracking of rapeseed oil was studied on three
different types of bifunctional catalysts: Pt/H-Y, Pt/H-ZSM-5, and sulfided NiMo/γ-Al2O3. Experiments were carried out in a
batch reactor over a temperature range of 300-400 LC and initial hydrogen pressures from 5 to 11 MPa. The reaction time was
limited to 3 h to prevent a high degree of cracking. The Pt-zeolite catalysts had a strong catalytic activity for both cracking and
hydrogenation reactions, and therefore a higher severity was required to reach a relatively high oil conversion into liquid
hydrocarbons. With dependence on the activity of the acid sites of the catalysts, the results show a trade-off between the yield of
green diesel and the degree of isomerization, which had a direct effect on the cold properties of the diesel. Among the three
catalysts, hydrocracking on Ni-Mo/γ-Al2O3 gave the highest yield of liquid hydrocarbons in the boiling range of the diesel
fraction, i.e., green diesel, containing mainly n-paraffins from C15 to C18, and therefore with poor cold flow properties. While for
both zeolitic catalysts, hydrotreating of rapeseed oil produced more iso- than n-paraffins in the boiling range of C5 to C22, which
included significant amounts of both green diesel and green gasoline. The gas chromatography (GC) analysis of the gaseous
phase revealed the presence of mainly CO2, CO, propane, and remaining hydrogen. It was observed that both pressure and
temperature play an important role in the transformation of triglycerides and fatty acids into hydrocarbons.

1. INTRODUCTION diesel is also environmentally competitive, as its use may produce


fewer greenhouse gases than petroleum diesel, biodiesel, and
Climate change has been shown to be directly influenced
1 fossil-derived syndiesel (without carbon sequestration). 5
by the increasing emission of greenhouse gases. As the
Trying to optimize the process variables that lead to a high-
combustion of transportation fuels contributes significantly
quality green diesel, a number of studies on the hydroprocessing
to the emission of CO2, it is indispensable to research for of vegetable oils have been carried out.6-18 Hydrotreating of
neutral carbon technol-ogies that lead to the production of model compounds derived from vegetable oil has also been
clean and renewable gasoline and diesel.
studied by Murzin et al.4,19 and Krause et al.20 Several oil com-
Hydroprocessing is a well-known technology in the panies have also developed commercial hydrotreating processes
petroleum refining industry, which can be carried out either by by considering the cofeeding of vegetable oil and vacuum gas oil
hydrocrack-ing technology or by the less severe hydrotreating (VGO).
21-27
The cofeeding with VGO has also been studied by some
technology. When they are applied to oxygenated researchers.
8,11,13,28
The main vegetable oils used in the above studies
hydrocarbons, the removal of oxygen can be carried out by have been sunflower,
8,18
rapeseed,
13,17
cottonseed,
11
2-4 10 22 21 24
decarboxylation, decarbonyla-tion, or hydrodeoxygenation. jatropha, soybean, castor, and palm oils. Animal fat
Hydroprocessing of vegetable oil leads to the production of 26
has also been considered as a viable feedstock. In general
hydrocarbons in the boiling range of diesel, and therefore it is these studies have shown that hydroprocessing vegetable
commonly called green diesel or renewable diesel. The basic idea oil contri-butes to the production of a renewable diesel
is to transform by effect of high pressure, high temperature, and a with a premium quality as a fuel.
bifunctional catalyst (e.g., sulfided NiMo/γ-Al2O3) the Among the most significant factors in the conversion of trigly-
triglycerides in the vegetable oil into high cetane hydrocarbons cerides into diesel hydrocarbons stands out that of the activity of
(mainly n-C17H36 and n-C18H38). The MoS2 phase of the catalyst the catalyst. Conventional hydrotreatment catalysts such as NiMo
and a high hydrogen pressure contribute to the saturation of the or CoMo supported on γ-Al2O3 have been mostly used for
side chains of the triglycerides. The acid function of the catalyst hydrotreating vegetable oils. Zeolites have also been con-sidered
contributes to the cracking of the C-O bound and to the in the processing of vegetable oils, however most of these studies
isomerization of the n-olefins formed, which after hydrogenation have been focused on catalytic cracking to produce
are transformed into isoparaffins.
Compared with biodiesel (fatty acids of methyl esters), green Special Issue: IMCCRE 2010
diesel in general has a higher oxidation stability, lower specific
gravity, higher cetane number, and when it is blended with petro- Received: April 5, 2010
leum diesel it has much better cold flow properties. In addition, Accepted: August 25, 2010
green diesel is totally compatible with petroleum diesel, thus Revised: August 18, 2010
changes or maintenance to the engine are not required. Green Published: September 23, 2010

r2010 American Chemical Society 2791 dx.doi.org/10.1021/ie100824d | Ind. Eng. Chem. Res. 2011, 50, 2791–2799
Industrial & Engineering Chemistry Research ARTICLE

gasoline29,30 rather than on hydrocracking to produce diesel. The Table 1. Fatty Acid Composition of Rapeseed Oil
moderate acidity of alumina has been preferred due to its reduced fatty acid a wt %
structure
cracking activity, which leads to a high yield of green diesel,
palmitic C16:0 3.65
containing mainly normal C17 and C18 paraffins and therefore
with a high cetane number. The cold flow properties of green palmitoleic C16:1 0.18
diesel are however poor and it would preferably be used blended, stearic C18:0 1.65
i.e., with petroleum diesel. To enhance the cold flow properties of oleic C18:1 63.72
green diesel, it is important to research for improved hydroi- linoleic C18:2 15.42
somerization activity of n-C17 and n-C18 paraffins, for instance by linolenic C18:3 14.28
using catalysts with more acidic sites than those of alumina. cis-11-eicosenoic C20:1 1.10
Traditional hydrocracking catalysts such as faujasite zeolites have
been shown to produce up to 30 wt % of isoparaffins in the aCx:y, where x is the number of carbon atoms; y is the number of
31 double bonds.
range from C7 to C17.
The hydrotreating of long chain n-paraffins on Pt/zeolite
Table 2. Physicochemical Properties of Rapeseed Oil
catalyst has been extensively studied by Froment et al. 32-35 who
have investigated the mechanisms of hydroisomerization and property
hydrocracking of long n-paraffins on different types of zeolites
and have modeled the single-event kinetics of the various elemental composition (wt %)
C 77.903
elementary steps by considering alkylcarbenium ions as key
reaction intermediates. The understanding of the chemistry of the H 11.689
hydrocracking process and of the catalyst structure lead to a better N 0.041
understanding of the isomerization/cracking selectivity of the S 0.000
catalyst, which finally may lead to selection of the best catalyst O 10.367
for the hydroprocessing of vegetable oils. density at 20 LC, g cm
-3 0.89
Thus, in this work, the effect of different active sites and viscosity at 25 LC, mPa s 63.0
sup-ports is investigated by comparing the hydroconversion of viscosity at 40 LC, mPa s 28.1
rape-seed oil on NiMo/γ-Al2O3 and on zeolites Pt/H-Y and acid value, mg of KOH/g of oil 0.55
Pt/H-ZSM-5. The scope of the present work is the study of the a 0.28
FFA content , wt %
optimal conditions at which both types of catalysts can
aOn the basis of oleic acid.
efficiently hydro-convert rapeseed oil, providing thus insights
that allow for a future optimal operation and industrial scaling. As the reactor was heated up to temperatures close to 300
Experiments were conducted on a batch reactor, and the yield LC, the elemental sulfur reacted with hydrogen to form
of green diesel was monitored.
H2S, which sulfided and activated the NiMo/Al2O3 catalyst
in the reactor for the hydrotreatment of vegetable oil.
2. METHODS 2.2.2. Pt/H-ZSM-5 Catalyst (2 wt % Pt). The precursor
NH4-H-ZSM-5 was obtained by ionic exchange between Na-
2.1. Materials. The rapeseed oil used in the experiments
corresponded to a commercial type sold in Japan. Table 1 shows the ZSM-5 and ammonium chloride (NH4Cl). This precursor was
total fatty acid composition of the oil. This was determined by dried for 24 h at 110 LC and then calcined at 550 LC in air for 3 h
derivation of the corresponding fatty acid methyl esters (FAME). A to form H-ZSM-5. The SiO 2/Al2O3 molar ratio in H-ZSM-5 was
36
modified method AOAC 969.33 by Lee et al. was used to obtain the 30. Pt/ H-ZSM-5 catalyst was then prepared using an incipient
wetness method by impregnating H-ZSM-5 particles in a 1 wt %
FAME. A GC-2014 Shimadzu was used. Table 2 shows some physical
properties of the oil. The CHNS composi-tion of the oil was H2PtCl6 aqueous solution. For a catalyst containing 2% of
determined by using an elemental analyzer (CE instruments EA1110); platinum, 1.5 g of H-ZSM-5 powder and 7.95 g of H 2PtCl6
the oxygen content was obtained by balance. The density was aqueous solution were used. About 20 mL of water were also
determined at 20 LC using a density/ specific gravity meter (Kyoto added. The mixture was heated at 30 LC with stirring for 1 h, and
Electronics DA-130N). The vis-cosity of the oil was determined at 20 then it was heated at 95 LC to evaporate all the water until a solid
and 40 LC using a vibro-viscometer (A&D Co. Lim. Japan, SV-10). sample was obtained. Subsequently, the solid sample was dried at
The acid value of rapeseed oil was obtained by titration with a KOH 110 LC for 24 h and then calcined at 400 LC in air for 3 h.
solution (0.1 N).
Finally, the catalyst sample was reduced at 350 LC for 2 h in a
tubular reactor with an internal diameter of 10 mm. and a
2.2. Catalysts. 2.2.1. NiMo/γ-Alumina. The catalyst precur- hydrogen flow rate of 50 mL (STP)/min.
sor consisting in a mixture of NiO and MoO supported in γ- 2.2.3. Pt/H-Y Zeolite Catalyst (2 wt % Pt). H-Y zeolite
alumina corresponded to a commercial type (CDS-R25NQ) and corresponded to a synthetic HS-320 catalyst and was purchased
was supplied by Catalyst and Chemicals Ind. Co. This catalyst is
from Waco Chemical Co. (I.D. 325-27765). The SiO 2/Al2O3
used in refining operations for hydrotreating of gas oil and
molar ratio in H-Y was 5.5. Pt/H-Y catalyst was prepared using
atmospheric residue. It contains a high desulfurization activity.
an impregnation method similar to Pt/H-ZSM-5. H-Y particles
The catalyst was crushed in particles of about 0.3 mm of diameter.
The activation of the catalyst was done in situ with elemental sulfur. were impregnated in an aqueous solution of H 2PtCl6, and the
water was then evaporated at 95 LC to obtain a solid sample. The
In a typical procedure, the elemental sulfur and the NiMo/Al 2O3
solid sample was then dried at 110 LC for 24 h, calcined in air at
catalyst (0.8 g of S/g of catalyst) were added to the vegetable oil
400 LC for 3 h, and finally reduced in hydrogen flow at 350 LC
inside the autoclave reactor. Then the reactor was tightly closed and
for 2 h. The loading of platinum in Pt/H-Y catalyst was 2 wt %.
purged with hydrogen at room temperature.

2792 dx.doi.org/10.1021/ie100824d |Ind. Eng. Chem. Res. 2011, 50, 2791–2799


Industrial & Engineering Chemistry Research ARTICLE

2.3. Experimental Procedure. Experiments were


conducted in an 80 mL-batch reactor with an internal diameter
of 20 mm and equipped with a mechanical stirrer. The
operative limits of the reactor were 30 MPa and 900 LC. For
all experiments, the stirrer speed was kept constant at 350
rpm, and the temperature control was (2 LC.
The feed consisted in rapeseed oil and catalyst in a ratio of
3 wt % cat/wt of oil. The reaction conditions for hydrotreating
experiments were a temperature range of 300-400 LC and initial
hydrogen pressures between 5 and 11 MPa. Before carrying out
the reaction, the reactor was purged several times with nitrogen
and then with hydrogen at room temperature. The reactor was
then heated up to the final temperature and kept at this condition
for 3 h (reaction time). Initial experiments from 1 to 6 h at 350 LC
were carried out to determine the optimal reaction time. A time of
3 h was found to be appropriate to avoid an excessive cracking of
the main components. Furthermore, after 3 h the yield of main Figure 1. Product distribution yields for paraffins after 3 h with
components did not increase significantly, indicating there was NiMo/ Al2O3 as the catalyst at 350 LC and an initial PH2 = 8 MPa.
enough time to reach equilibrium.
After cooling down the reactor, the gas and liquid products
were recovered. The liquid phase containing the catalyst was
filtered. When solid product was obtained, this was diluted
prior to its analysis with 1,2,3,4-tetra hydronaphthalene. Both
gas and liquid phases were analyzed by gas chromatography.
The pro-duced gases were analyzed by a GC 323 (GL
Sciences) equipped with a thermal conductivity detector and
two columns, one being a Pora-Q capillary column (30 m,
0.53 mm i.d.) for determina-tion of CO 2 and the other one a
packed column (MS-5A) for determination of H2, O2, N2, and
CO. Light hydrocarbons (C1-C4) were analyzed by a FID GC-
353 B (GL Sciences) equipped with a RT-QPLOT capillary
þ
column. C5 hydrocarbons were analyzed by an Agilent 6890
N FID-GC equipped with an UA-DX capillary column. Fatty
acids were analyzed by an Agilent 6890N FID-GC equipped
with a HP-624 capillary column. To verify the presence in the
liquid product of important com-pounds, such as remaining Figure 2. Yields of the C7-C24 iso-paraffins in the liquid fraction after
carboxylic acids, isoparaffins, cyclic paraffins, and aromatics, 3 h with NiMo/Al2O3 as the catalyst at 350 LC and the initial PH2
a detailed analysis by mass spectrometry was performed. = 8 MPa.

3. RESULTS AND DISCUSSION 3.2. Hydrotreating with NiMo/Al2O3 Catalyst. In this study, we
used a gas chromatography-mass spectrometry (GC-MS) analysis to
3.1. Rapeseed Oil Composition. The composition of the determine the composition of the different products, which allowed
vegetable to be hydrotreated has a direct effect on the con- us to present the results on the basis of carbon number intervals. All
sumption of hydrogen, which is needed to saturate the double yields in this paper are reported on a weight percent basis with
bonds in the side chains of triglycerides, as well as the unsatu- respect to the weight of the oil. Figure 1 shows the paraffinic product
rated carboxylic acids, and olefins that are formed by hydro- yields for hydrotreating rapeseed oil on NiMo/Al 2O3 from an initial
cracking. As shown in Table 1, the rapeseed oil used in this pressure of 8 MPa and a final temperature of 350 LC for 3 h. At these
work has a high content of unsaturated carboxylic acids, i.e., conditions, the rapeseed oil is converted into a liquid phase product
oleic, linoleic, and linolenic acids, which is an indication that that contains mainly n-paraffins and isoparaffins in between C 15 and
rapeseed oil contains mostly triolein, trilinolein, and trilinole- C24 hydrocar-bons. The total yield of liquid hydrocarbon was 86.3 wt
nin. As these are unsaturated triglycerides, the hydrotreating %. The main components were n-heptadecane with 32.7 wt % and n-
of rapeseed oil needs a higher consumption of hydrogen octadecane with 18.4 wt %. As the cetane number of n-C17 and n-C18
during the hydrotreating process than the one that would need is higher than 100, the cetane number of the green diesel will be very
the hydrotreating of less unsaturated oils, e.g., palm oil, which high and probably much higher than that of biodiesel or that of
is rich in palmitolein. petrodiesel. Figure 2 shows the distribution of isoparafins in the
In Table 2, it is also possible to notice that this commercial liquid phase. It was observed that the main isoparaffins were single-
rapeseed oil does not contain a significant amount of free fatty branched C17 and C18 paraffins, e.g., 1-methyl hexa-decane and 2-
acids (FFA), as it is given by the acid value. The low content of methyl heptadecane. Compared to the yield of n-paraffins, the global
FFA, characteristic of commercial oils, is desirable since FFA yield of isoparaffins is very low, which is thought to be related to the
could eventually be transformed into a more unsaturated FFA, moderate acidity of the alumina. The same reason can explain the
increasing the number of double bonds and therefore increasing nonsignificant yield of cracked
the consumption of hydrogen and the operative costs.
2793 dx.doi.org/10.1021/ie100824d |Ind. Eng. Chem. Res. 2011, 50, 2791–2799
Industrial & Engineering Chemistry Research ARTICLE

Scheme 1. Main Reactions of Stearic Acid under Hydrogenation Conditions

gas shift reaction. Thus, with a decrease in the partial pressure


of steam (H2O) and CO, the equilibrium of decarbonylation
reaction would favor a higher production of n-C17H36.
The higher yield of n-C17H36 than that of n-C18H38 indicates
that at the above-mentioned conditions decarboxylation and
decarbonylation are favored over hydrodeoxygenation.
An important byproduct in the gas phase is propane, viz.
Figure 3. Its formation is believed to be mainly due to the
cracking of the carboxylic C17- and C18-side chains of the
triglycerides, with consequent transformation of the glycerol
backbone into propane. In a large scale operation, propane
pro-duction will be economically very beneficial. In addition,
carbon monoxide could be steam-reformed to generate the
27
hydrogen needed for producing green diesel, and by doing
so the process could be economically more viable. To reduce
Figure 3. Yields of the components in the gas phase after 3 h with the impact of CO2 as a green house gas, it could be
NiMo/Al2O3 as the catalyst at 350 LC and the initial PH2 = 8 MPa. sequestrated through en-hanced oil recovery, which can also
37
bring an economic profit for oil producers.
products, as it is noticed in Figure 1 by the yield of the 3.3. Effect of Initial Hydrogen Pressure. As the hydrogen
fraction corresponding to C5-C12 hydrocarbons (gasoline), consumption is one of the important limiting economical factors
which only accounts for 3.2 wt %. in hydrotreating, the effect of the initial hydrogen pressure in the
By GC-MS analysis, a very low content of cycloalkanes and reactor was studied. To evaluate this effect, several experiments
aromatics was also detected. The latter ones were more pre- were performed in an interval of 5-11 MPa. When using NiMo/
dominant in the products obtained with zeolitic catalysts. This Al2O3 as a catalyst, initial hydrogen pressures of 8-10 MPa were
may be due to the stronger acids sites in the zeolites that found to be more appropriate. At pressures below 8 MPa, the
promote isomerization, cracking, and cyclization. product was partially solid and a high content of saturated carbo-
As a basis for the subsequent discussion of results and to get a xylic acids were detected, i.e., palmitic and stearic acids
better understanding of the reaction mechanism, Scheme 1 shows primarily. With zeolitic catalysts, even higher pressures were
the transformation of stearic acid into hydrocarbons by the three required to obtain a liquid product. In this section, only the results
main pathways: n-C17 is the main product by the decarbonylation with NiMo/Al2O3 as the catalyst are shown. In Section 3.5, the
and decarboxylation pathways, while n-C18 is the main product results with zeolitic catalysts are described.
by the hydrodeoxygenation pathway (water is a byproduct). Figures 4-7 show the global yields for the different
The gas product distribution from the hydrotreating of rape- fractions obtained when using NiMo/Al 2O3, a reaction time of
seed oil is shown in Figure 3. The remaining hydrogen (∼40% 3 h at 350 LC, and different initial hydrogen pressures. As
þ
mol) is not included in this figure. The total yield of CO 2 observed in Figure 4, the yield of C 23 fraction (wax)
predominates over that of CO, which may indicate that decar- decreased, indicating that catalytic cracking of these
boxylation (loss of the carboxylic group as CO 2) is favored over compounds into lighter hydro-carbons is favored as pressure
decarbonylation (loss of the carboxylic group as CO, via hydro- increases.
genation) at the above conditions. This is because the former In the interval from 8 to 10 MPa, the global yield of green
reaction does not need hydrogen as the latter does, and since this diesel is slightly affected by the pressure, as it only varies
from 74.8 to 78.5 wt %, viz. Figure 5. In the same interval, the
study was carried out in a batch reactor, the hydrogen pressure is
continuously decreasing and also the rate of decarbonylation. yield of n-C17 was higher than that of n-C18, as it can be
observed in Figure 6. In the gas phase, there was also
It is important however to notice that the presence of CO 2
could also be due to the water gas shift reaction, in which CO and observed in all cases a higher amount of CO 2 than CO. These
water are transformed into CO2 plus hydrogen. As CO and water two observations confirm that decarboxylation is the main
are formed by decarbonylation (viz. Scheme 1), they are prob- reaction pathway at the condi-tions in the batch reactor.
ably transformed into CO2 and hydrogen by means of the water In Figure 6, it can also be observed that the effect on
increasing the initial hydrogen pressure is slightly more
2794 dx.doi.org/10.1021/ie100824d |Ind. Eng. Chem. Res. 2011, 50, 2791–2799
Industrial & Engineering Chemistry Research ARTICLE

Figure 7. Influence of initial PH2 on the global production of


Figure 4. Influence of initial PH2 on the global production of the carboxylic acids. T = 350 LC on NiMo/Al2O3.
main fractions: gas (C1-C4), gasoline (C5-C12), and wax (C23þ).
T = 350 LC on NiMo/Al2O3.

Figure 8. Influence of temperature on the production of the main


Figure 5. Influence of initial PH2 on the global production of the main
fractions: gas (C1-C4), gasoline (C5-C12), and wax (C23þ). Initial
fraction: green diesel (C13-C22). T = 350 LC on NiMo/Al2O3.
PH2 = 9 MPa; NiMo/Al2O3 as the catalyst.
Figure 7 shows the global yield of carboxylic acids at different
initial hydrogen pressures. At 350 LC and 10 MPa, the product still
contained fatty acids, mainly stearic and oleic acids. To totally
convert these fatty acids into hydrocarbons, it was necessary to
increase the temperature, as it is described in the next section.
3.4. Effect of Temperature. The effect of temperature on
the hydrocracking of rapeseed oil on NiMo/Al 2O3 was studied
by performing experiments from 300 to 400 LC and with an
initial hydrogen pressure of 9 MPa.
Global yields of the different fractions obtained at different
temperatures can be observed in Figures 8 and 9. The yields of
the main products are shown in Figure 10. At temperatures
below 350 LC, the product was partially solid. Only a small
Figure 6. Influence of initial PH2 on the production of major products: amount of liquid was obtained. Therefore only those results
n-heptadecane and n-octadecane. T = 350 LC on NiMo/Al2O3. obtained at temperatures at 350 LC and above are presented
here. At tem-peratures above 350 LC there is an increase in
pronounced on the yield of n-octadecane than on the yield of the degree of cracking of n-paraffins in the boiling range of
n-heptadecane. Thus, the relative rate of decarboxylation and the diesel fraction, as it can be observed by the decreasing
decarbonylation versus hydrodeoxygenation decreases as the yield trend of the major n-paraffins, i.e., n-C17 and n-C18, in
hydrogen pressure increases, as the latter reaction requires Figure 10. Consequently, the yield of the C5 to C12 fraction
more hydrogen, viz. Scheme 1. Therefore, with an increase in (gasoline) increases, as shown in Figure 8.
the pressure, hydrodeoxygenation could be pro-moted and a In Figure 9, the maximum yield (78 wt %) of the green diesel
higher yield of n-C18 would be reached. This reaction (C13-C22 fraction) is observed at 350 LC; at 375 LC the yield of
however produces water as a byproduct, which may deactivate this fraction is slightly reduced to about 76 wt %, but when the
the catalyst and reduce the yield of green diesel. Previously, temperature is increased above 380 LC (viz. Figure 10), there is a
38
Senol et al. on their study of hydrodeoxygena-tion of significant reduction in the yield of green diesel to about 52 wt %.
aliphatic esters on sulphided NiMo/γ-Al2O3 have observed a The small effect of increasing the reaction temperature from 350
decrease in the yields of hydrocarbons when the amount of to 375 LC on the yield of green diesel was observed to be due to
water was increased. the increase in the yield of C13 to C16 hydrocarbons with

2795 dx.doi.org/10.1021/ie100824d |Ind. Eng. Chem. Res. 2011, 50, 2791–2799


Industrial & Engineering Chemistry Research ARTICLE

Figure 9. Influence of temperature on the production of green diesel


(C13-C22). Initial PH2 = 9 MPa; NiMo/Al2O3 as the catalyst. Figure 11. Influence of temperature on the global yield of fatty acids.
Initial PH2 = 9 MPa; NiMo/Al2O3 as the catalyst.

reported by Huber et al.8 and Simacek et al.,17 which have


performed hydrotreating of similar vegetable oils on a contin-
uous reactor, in which the pressure is kept constant and there is
enough hydrogen to promote the hydrodeoxygenation reaction.
Thus, these other authors have found that n-C18 yield is higher
than that of n-C17.
In Figure 11, the global yield of fatty acids in the product
obtained at different temperatures is also observed. These results
indicate that an enough high temperature (∼375 LC) is required
to hydrocrack the carboxylic acids. At 350 LC, the product analy-
sis revealed the presence of carboxylic acids, mainly stearic acid,
which is solid at room temperature. The total transformations of
Figure 10. Influence of temperature on the production of major carboxylic acids lead to a purely hydrocarbon-based product. This
products. Initial PH2 = 9 MPa; NiMo/Al2O3 as the catalyst. product without carboxylic acids is liquid at room tempera-ture.
Thus, the hydrocracking of carboxylic acids can be con-sidered
temperature, probably by means of cracking of C 17 to C22 the determining step in the conversion of vegetable oils into
hydrocarbons and the consequent production of lighter gasoline and diesel hydrocarbons.
hydro-carbons. The yield of i-C17 was also observed to It is important to mention that in a previous study by Simacek
slightly increase in this temperature range. et al.17 using a continuous reactor and a similar catalyst to the one
The global yield of C 5 to C22 hydrocarbon products was near 81 wt used in this work have shown that a temperature close to 310 LC
% at 375 LC (viz. Figures 8 and 9), which indicates that almost all C- was enough to hydrocrack carboxylic acids from rapeseed oil into
CO bonds were cracked at 375 LC. When the temperature was liquid products. Those authors have used a high-pressure flow
increased from 375 to 400 LC, the yields of green diesel and wax system in which the hydrogen pressure was kept at 7 MPa with a
significantly decreased and the amount of light hydrocarbons greatly surplus flow rate of hydrogen to the raw material (1000:1). In the
increased. This indicates that temperature plays an important role in present work, we have used a batch reactor in which hydrogen is
the cracking of the intermediate carbe-nium ions when the continuously decreasing and therefore a high temperature is
temperature is higher than 375 LC. needed to hydrocrack the carboxylic acids. With the use of the
This work considers that the transformations of hydrocarbons autoclave reactor, a large amount of hydrogen has been saved.
follow a bifunctional mechanism. After n-C17H36 and n-C18H38 Some other differences with the previous work are the activation
were formed by the hydrotreatment of vegetable oil on NiMo method and the catalyst amounts used in the experiments.
sites, the corresponding carbenium intermediates were formed on 3.5. Hydrotreating with Pt/Zeolite Catalysts. The high
the acid sites of Al2O3. The carbenium intermediates under-went content of C15 to C18 n-paraffins assures that the liquid produced
cracking to form light hydrocarbons and isomerization to form
with NiMo/Al2O3 as catalyst is a high cetane number diesel fuel
isoparaffins. The formed isoparaffins were also hydro-cracked in
the reactor. These caused that the products from the containing mainly n-C17 and n-C18. However the pour point of
hydrotreatment of vegetable oil contained different paraffins these n-paraffins is relatively high. Although it is not likely that
although n-C17H36 and n-C18H38 were the main components green diesel will be used anytime soon unblended, it could be
when the temperature was lower than 375 LC. desirable to improve its cold properties. The best option is to
transform these n-paraffins into their corresponding isoparaffins,
From Figure 10, it is possible to observe the influence of which have a very low pour point. As the isomerization reaction
temperature on the yields of n-C17 and n-C18. The n-C17/n-C18 is promoted by the acids sites of the catalyst, the use of strong
yield ratios at 350, 375, and 400 LC were 2.2, 1.6, and 1.5, acid catalyst, such as zeolites, can be appropriate.
respectively. This indicates that by using a batch reactor in which Two types of zeolitic catalysts were used. The first is a plati-
the hydrogen pressure decreases with time, the relative rate of num catalyst containing H-ZSM-5 zeolite, which belongs to the
decarbonylation plus decarboxylation (n-C17 production) versus family of so-called 10-ring zeolites (having micropores with
hydrodeoxygenation (n-C18 production) decreases as the reac-tion windows circumscribed by 10 oxygen atoms). 35 The second cata-
temperature increases. This result is contrary to the ones lyst was platinum containing H-Y zeolite, which is acknowledged

2796 dx.doi.org/10.1021/ie100824d |Ind. Eng. Chem. Res. 2011, 50, 2791–2799


Industrial & Engineering Chemistry Research ARTICLE

5.5 for NiMo/Al2O3; while at 380 LC and at initial pressure of


11 MPa, the ratios n-P/i-P in the product for Pt/H-ZSM-5 and
Pt/HY were 0.527 and 0.744, respectively. The acid sites of the
zeolite clearly promoted the isomerization of n-paraffins. Because
of its stronger acid sites and probably due to the shape selectivity
that are known to have the 10-ring zeolites, Pt/H-ZSM-5 was the
one which gave the highest yield of isoparaffins. This is in line
with the results found by Park et al. 40 and Souverijnis et al.35 who
have studied the hydroisomerization of long chain paraffins in
different Pt/zeolite catalysts. The shape selectivity of the H-ZSM-
5 zeolite may have prevented the cracking of multi-branched
paraffins by limiting the introduction of these mole-cules into the
inner micropores of the zeolite. Previous studies 35 have shown
that the H-Y catalyst promotes the formation of multibranched
Figure 12. Global yields obtained with the operation of the two isoparaffins that have the advantage of lowering the pour point
zeolitic catalysts used in the experiments. Initial PH2 = 11 MPa; but the disadvantage of being susceptible to cracking, resulting in
final T = 380 LC for 3 h. a yield loss of diesel hydrocarbons.
Compared with the yield of diesel obtained with
NiMo/Al2O3 (from 70-80%), the yield of green diesel with
Pt/zeolite catalysts is very low (from 20 to 40%). Although
hydrotreating of rapeseed oil with both zeolitic catalysts
produce a green diesel with a relatively high concentration of
isoparaffins and hence with better cold flow properties than
that produced with NiMo/ Al2O3, they require a higher
severity, i.e., higher temperature, which increases the cracking
of long paraffinic chains and pro-duces middle molecules in
the range of the gasoline fraction. Thus, both thermal and
catalytic cracking reduce the yield of diesel hydrocarbons.
Among both zeolitic catalysts, the Pt/H-Y catalyst seems to be
more appropriate to produce a high yield of green diesel, while
Pt/H-ZSM-5 is more appropriate for producing green gasoline,
though with this catalyst there are also more light ends and CO 2
produced, viz. Figure 13. On the other side, in some countries,
Figure 13. Yields of the gas phase components obtained with the gasoline would be more profitable than diesel and thus its
operation of the two zeolitic catalysts used in the experiments. production could be of great interest for refiners as well.
Initial PH2 = 11 MPa; final T = 380 LC for 3 h. In Figure 13, it is also possible to notice that the total amount
of CO plus CO2 is very different for both catalysts. A possible
to have a larger size of intersections between micropore channels reason could be that Pt/H-Y favors reduction, while Pt/H-ZSM-
and a greater hydrogen transfer capacity than H-ZSM-5. 39 5 favors decarbonylation and decarboxylation from the
Figure 12 compares the yields of the major fractions obtained viewpoint of the reaction mechanism. As compared with the
with both catalysts when operating with an initial hydrogen
Pt/H-Y catalyst, Pt/H-ZSM-5 catalyst possesses stronger acid
sites but the acid amount in H-ZSM-5 is less than that in H-Y
pressure of 11 MPa and 380 LC during 3 h. At these conditions,
both catalysts gave the highest yield of diesel. The hydrotreating because the SiO2/Al2O3 molar ratio was 5.5 in H-Y, while in
of rapeseed oil on Pt/H-ZSM-5 produced a higher yield of H-ZSM-5 was 30.The acid strength and acid amount in
gasoline (C5-C12 fraction) than diesel (C13-C22 fraction). The zeolites probably influenced the hydrocracking of carboxylic
contrary happened with Pt/H-Y. The higher production of cracked acids and hence the yield of CO þ CO2 in the products.
products with the 10-ring zeolite suggests that the acid sites of It is important to mention that for most of experiments with
Pt/H-ZSM-5 are stronger than those of Pt/H-Y. Similar results both zeolite catalysts, there was also observed the presence of
have been found by Souverijns et al. 35 In Figure 13, the higher about a 10-20% yield of carboxylic acids in the products, mainly
yield of propane produced by hydrocracking with Pt/ H-ZSM-5 stearic, oleic, and palmitic acids. To increase their conversion into
reveals also its strong cracking activity by means of the β-scission hydrocarbons, it was necessary to increase the severity. However,
mechanism. The presence of methane and ethane in the gas with an increase in temperature, e.g., higher than 380 LC, a high
product with Pt/H-ZSM-5 could be due to hydrogeno-lysis of degree of cracking was observed. After 3 h of reaction, the pres-
long paraffins on the platinum sites. sure in the reactor was slightly decreased for Pt/H-Y and slightly
As with the NiMo/Al2O3 catalyst, there is a higher content of increased for Pt/H-ZSM-5, with respect to the initial pressure.
This is probably due to the high content of light hydrocarbons
CO2 than CO in the gas phase produced by the hydrocracking of
that are formed because of cracking. In addition, thermal and
rapeseed oil with both Pt/zeolites, which could be an indication
that decarboxylation is the main reaction pathway at the condi- catalytic cracking produce olefins which are then saturated, and
tions in which the experiments were carried out. Another char- therefore, as the degree of cracking increases, the consumption of
acteristic found for the three catalysts used in this study was the hydrogen increases as well. Thus, as the reaction proceeds, the
ratio of n-paraffins to isoparaffins in the interval of C 5 to C22 hydrogen pressure eventually is not enough to convert the fatty
paraffins. At 375 LC and initial pressure of 9 MPa, the ratio was acids into hydrocarbons.

2797 dx.doi.org/10.1021/ie100824d |Ind. Eng. Chem. Res. 2011, 50, 2791–2799


Industrial & Engineering Chemistry Research ARTICLE

In summary, the results obtained with this study indicate that ’REFERENCES
both pressure and temperature play an important role in the (1) IPCC. Fourth Assessment Report: Climate Change 2007; 2007.
transformation of triglycerides and mainly carboxylic acids. A (2) Sivasamy, A.; Cheah, K. Y.; Fornasiero, P.; Kemausuor, F.;
high enough hydrogen pressure is needed to hydrogenate the Zinoviev, S.; Miertus, S. Catalytic Applications in the Production of
unsaturated chains of triglycerides and carboxylic acids as well as Biodiesel from Vegetable Oils. ChemSusChem 2009, 2, 278–300.
those of the olefins formed by cracking. Another important role (3) Furimsky, E. Catalytic Hydrodeoxygenation. Review.
played by hydrogen is keeping the catalytic activity while impe- Appl. Catal., A: General 2000, 199, 147–190.
ding coke formation. Increasing temperature also favors the (4) Lestari, S.; Mki-Arvela, P.; Bernas, H.; Simakova, O.;
cracking of the double bond CdO and that of the bond R— CO in Sjholm, R.; Beltramini, J.; Lu, G. Q. M.; Myllyoja, J.; Simakova,
I.; Murzin, D. Y. Catalytic Deoxygenation of Stearic Acid in a
the carboxylic acids, thus favoring the production of C 18 and C17 Continuous Reactor over a Mesoporous Carbon-Supported Pd
n-paraffins, respectively. Catalyst. Energy Fuels 2009, 23, 3842–3845.
(5) Kalnes, T. N.; Koers, K. P.; Marker, T.; Shonnard, D. R. A
4. CONCLUSIONS Technoeconomic and Environmental Life Cycle Comparison of
Green Diesel to Biodiesel and Syndiesel. Environ. Prog.
The hydrotreating of vegetable oils on bifunctional catalysts Sustainable Energy 2009, 28, 111–120.
makes possible the production of liquid hydrocarbons known as (6) Gusmao, J.; Brodzki, D.; Djega-Mariadassou, G.; Frety, R.
“green diesel” containing mostly n-heptadecane and n-octade- Utilization of Vegetable Oils as an Alternative Source for Diesel-Type
cane, which are mainly formed by decarboxylation/decarbonyla- Fuel: Hydrocracking on Reduced Ni/SiO2 and Sulphided NiMo/Al2O3.
tion and hydrodeoxygenation reactions, respectively. Catal. Today 1989, 533–544.
Because of their strong acid sites, both Pt/zeolites-supported (7) Da Rocha Filho, G. N.; Brodzki, D.; Djega-Mariadassou, G.
catalysts have a strong cracking activity. The less acidity of the Formation of Alkanes, Alkylcycloalkanes and Alkylbenzenes During
Pt-HY catalyst led to a larger production of green diesel than Catalytic Hydrocracking of Vegetable Oils. Fuel 1993, 72, 543–549.
green gasoline. The yield of diesel with these catalysts is lower (8) Huber, G. W.; O’Connor, P.; Corma, A. Processing
Biomass in Conventional Oil Refineries: Production of High
than that obtained when using NiMo/Al2O3 as a catalyst. Quality Diesel by Hydrotreating Vegetable Oils in Heavy Vaccum
Even though the NiMo/Al2O3 catalyst effectively Oil Mixtures. Appl. Catal., A: General 2007, 120–129.
promotes the hydrocracking of triglycerides and carboxylic (9) Stumborg, M.; Wong, A.; Hogan, E. Hydroprocessed Vegetable Oils
acids at lower temperature and pressure, its moderate for Diesel Fuel Improvement. Bioresour. Technol. 1996, 56, 13–18.
acidity does not largely contribute to the production of (10) Sotelo-Boyas, R.; Liu, Y.; Murata, K.; Minowa, T.; Sakanishi, K.
isoparaffins, which are also im-portant due to their lower Hydrotreatment of Jatropha Oil to Produce Green Diesel over Trifunc-
pour points than those of their cor-responding n-paraffins. tional Ni-Mo/SiO2-Al2O3 Catalyst. Chem. Lett. 2009, 38, 552–553.
The process to produce diesel with zeolitic catalysts requires a (11) Sebos, I.; Matsoukas, A.; Apostolopoulos, V.; Papayannakos, N.
Catalytic Hydroprocessing of Cottonseed Oil in Petroleum Diesel
higher severity than NiMo/Al2O3 and therefore a higher operative
Mixtures for Production of Renewable Diesel. Fuel 2009, 88, 145–149.
cost. The strong acid sites of zeolites favor the production of (12) Bezergianni, S.; Voutetakis, S.; Kalogianni, A. Catalytic
isoparaffins, which are desirable for the diesel to have a low pour Hydro-cracking of Fresh and Used Cooking Oil. Ind. Eng.
point. It is therefore necessary to moderate the acidity of the zeolite Chem. Res. 2009, 48, 8402–8406.
catalysts to increase the isomerization activity while keeping a (13) Donnis, B.; Gottschalck, R.; Blom, E. P.; Knudsen, K. G.
moderate cracking. Thus, a large amount of isoparaffins could be Hydroprocessing of Bio-Oils and Oxygenates to Hydrocarbons.
obtained. This work has shown the conditions at which two zeolitic Under-standing the Reaction Routes. Top. Catal. 2009, 52, 229–240.
catalysts can produce green diesel as well as green gasoline. Further (14) Kubicka, D.; Simacek, P.; Zilkova, N. Transformation of
research is needed to determine the appropriate acidity of zeolite Vegetable Oils into Hydrocarbons over Mesoporous-Alumina-
catalysts to improve the hydroisomerization selectivity while keep- Supported Como Catalysts. Top. Catal. 2009, 52, 161–168.
ing an effective hydrocracking of triglycerides and fatty acids. (15) Shimada, H.; Sato, T.; Yoshimura, Y.; Hinata, A.;
The transformation of carboxylic acids is the most Yoshitomi, S.; Castillo-Mares, A.; Nishijima, A. Application of
Zeolite-Based Catalyst to Hydrocracking of Coal-Derived
important step for the production of diesel hydrocarbons from
Liquids. Fuel Process. Technol. 1990, 25, 153–165.
rapeseed oil. The quality of the diesel product depends mainly (16) Zhao, Y.; Wu, J.; Wang, X.; Zhang, X.; Meng, X.
on the pressure, temperature, and on the balance and strength Advance in Hydroprocessing Technology of Manufacturing
of metal-acid sites of the catalyst used. Diesel with High Cetane Number from Vegetable Oil. Chem.
Because of its large concentration of n-C17 and n-C18 Ind. Eng. Prog. In Chinese 2007, 26, 1391–1394.
hydro-carbons, the green diesel as obtained with NiMo/Al 2O3 (17) Simacek, P.; Kubicka, D.; Sebor, G.; Pospisil, M.
probably has a high cetane number even larger than that of the Hydropro-cessed Rapeseed Oil as a Source of Hydrocarbon-
typical biodiesel composed by FAME. Green diesel can be Based Biodiesel. Fuel 2009, 88, 456–460.
(18) Hancsok, J.; Krar, M.; Magyar, S.; Boda, L.; Hollo, A.; Kallo, D.
used directly as transportation fuel, although at the current
Investigation of the Production of High Cetane Number Bio Gas Oil from
time it could be used rather as an excellent additive for
increasing the cetane number of petroleum diesel. Pre-Hydrogenated Vegetable Oils over Pt/Hzsm-22/ Al2O3.
Microporous Mesoporous Mater. 2007, 101, 148–152.
(19) Snare, M.; Kubickova, I.; Maki-Arvela, P.; Eranen, K.; Warna, J.;
’AUTHOR INFORMATION Murzin, D. Y. Production of Diesel Fuel from Renewable Feeds: Kinetics of
Corresponding Author Ethyl Stearate Decarboxylation. Chem. Eng. J. 2007, 134, 29–34.
(20) Senol, O. I.; Viljava, T. R.; Krause, A. O. I.
*E-mail: rsotelob@gmail.com. Hydrodeoxygenation of Methyl Esters on Sulphided NiMo/γ-Al2O3
Funding Sources and Co Mo/γ-Al2O3 Catalysts. Catal. Today 2005, 100, 331–335.
The Japanese International Cooperation Agency (JICA) is ac- (21) Gomes, J. R. Vegetable Oil Hydroconversion Process.
U.S. Patent Application 20060186020 A1, 2006.
knowledged for the financial support to carry out the present study.
2798 dx.doi.org/10.1021/ie100824d |Ind. Eng. Chem. Res. 2011, 50, 2791–2799
Industrial & Engineering Chemistry Research ARTICLE

(22) PETROBRAS H-Bio Process. http://www2.petrobras.com.br/


tecnologia/ing/hbio.asp (August 2010).
(23) Duarte, F. A.; Soares, W.; Ferreira, A. E.; Ximenes, L. M.;
Gomes, J. R. Process to Obtain N-Paraffins from Vegetable Oil. U.S.
Patent Application 20070260102 A1, November 8, 2007.
(24) Ono, H.; Iki, H.; Koyama, A.; Iguchi, Y. In Production of
Bhd (Bio Hydrofined Diesel) with Improved Cold Flow
Properties; 19th Annual Saudi-Japan Symposium: Catalysts in
Petroleum Refining & Petrochem-icals, Dhahran, Saudi Arabia, 2009.
(25) Neste Oil. NexBTL Renewable Synthetic Diesel, http://www.
climatechange.ca.gov/events/2006-06-27þ28_symposium/presenta-
tions/CalHodge_handout_NESTE_OIL.PDF (December 2009).
(26) Yao, J.; Sughrue, E. L.; Cross, J. B.; Kimble, J. B.; Hsing, H.;
Johnson, M. M.; Ghonasi, D. B. Process for Converting Triglycerides to
Hydrocarbons. U.S. Patent Application 0175795 A1, August 2, 2007.
(27) Kalnes, T. N.; Marker, T.; Shonnard, D. R. Green Diesel: A
Second Generation Biofuel. Int. J. Chem. React. Eng. 2007, 5.
(28) Bezergianni, S.; Kalogianni, A.; Vasalos, I. A. Hydrocracking of
Vacuum Gas Oil-Vegetable Oil Mixtures for Biofuels Production.
Bioresour. Technol. 2009, 100, 3036–3042.
(29) Buzetzki, E.; Svanova, K.; Cvengros, J. In Zeolite Catalysts
in Cracking of Natural Triacylglycerols, 44th International
Petroleum Con-ference, Bratislava, Slovak Republic, 2009.
(30) Milne, A. T.; Evans, J. R.; Nagle, N. Catalytic Conversion of
Microalgae and Vegetable Oils to Premium Gasoline, with Shape-
Selective Zeolites. Biomass 1990, 21, 219–232.
(31) Martens, J. A.; Jacobs, P. A. Evidence for Hypocage
Catalysis in Fau/ Emt Zeolites Derived from Isomerisation and
Hydrocracking of Heptadecane. J. Mol. Catal. 1993, 78, L47–L52.
(32) Froment, G. F. Kinetics of the Hydroisomerization and Hydro-
cracking of Paraffins on a Platinum Containing Bifunctional Y-Zeolite.
Catal. Today 1987, 1, 455–487.
(33) Martens, G.; Froment, G. F. Kinetic Modeling of Paraffins
Hydrocracking Based Upon Elementary Steps and the Single Event
Concept. Stud. Surf. Sci. Catal. 1999, 122, 333–340.
(34) Martens, J. A.; Uytterhoeven, L.; Jacobs, P. A.; Froment,
G. F. Isomerisation of Long-Chain N-Alkanes on Pt/H-ZSM-22
and Pt/H-Y Zeolite Catalysts and on Their Intimate Mixtures.
Stud. Surf. Sci. Catal. 1993, 75, 2829–2832.
(35) Souverijns, W.; Martens, J. A.; Froment, G. F.; Jacobs, P.
A. Hydrocracking of Isoheptadecanes on Pt/H-ZSM-22: An
Example of Pore Mouth Catalysis. J. Catal. 1998, 174, 177–184.
(36) Lee, D. S.; Noh, B. S.; Bae, S. Y.; Kim, K. Characterization of
Fatty Acids Composition in Vegetable Oils by Gas Chromatography and
Chemometrics. Anal. Chim. Acta 1998, 163–175.
(37) Leach, A.; Mason, C. F.; van’t Veld, K., Co-Optimization of
Enhanced Oil Recovery and Carbon Sequestration; SSEE,
University of Oxford: Oxford, U.K., 2010; Working Paper 15035.
(38) Senol, O. I.; T. R.; Viljava, T. R.; Krause, A. O. I.
Hydrodeoxy-genation of Aliphatic Esters on Sulphided NiMo/γ-
Al2O3 and CoMo/ γ-Al2O3 Catalyst: The Effect of Water. Catal.
Today 2005, 106, 186– 189.
(39) Grzechowiak, J. R.; Masalska, A. Comparison of
Activities of Cracking Catalysts Containing Zsm-5 and Y-Zeolite.
React. Kinet. Catal. Lett. 1985, 29.
(40) Park, K. C.; Ihm, S. K. Comparison of Pt/Zeolite
Catalysts for n-Hexadecane Hydroisomerization. Appl. Catal., A:
General 2000, 203, 201–209.

2799 dx.doi.org/10.1021/ie100824d |Ind. Eng. Chem. Res. 2011, 50, 2791–2799

You might also like