You are on page 1of 6

The comparison of sulfide CoMo/-Al

2
O
3
and NiMo/-Al
2
O
3
catalysts in
methyl palmitate and methyl heptanoate hydrodeoxygenation

IRINA V. DELIY
1,2
, EVGENIA N. VLASOVA
1
, ALEXEY L. NUZHDIN
1
,
GALINA A. BUKHTIYAROVA
1

1
Boreskov Institute of Catalysis
Pr. Lavrentieva 5, Novosibirsk 630090, Russia
2
Novosibirsk State University
Pirogova Street 2, Novosibirsk 630090, Russia
RUSSIA
delij@catalysis.ru


Abstract: - The hydrodeoxygenation of methyl palmitate and methyl heptanoate as the model compounds of
bio-oil in the presence of sulfided CoMo/-Al
2
O
3
and NiMo/-Al
2
O
3
catalysts was studied at the temperature
300
o
C and hydrogen pressure 35 bars. In hydrodeoxygenation experiments the performances of sulfided CoMo
and NiMo catalysts were compared in terms of methyl palmitate and methyl heptanoate conversions, oxygen
removal, and the reaction products distribution. The CoMo/-Al
2
O
3
catalyst exhibited higher catalytic activity
in the methyl palmitate HDO than the NiMo/-Al
2
O
3
using n-tetradecane as a solvent. The main product
obtained over CoMo/-Al
2
O
3
catalyst was hexadecane whereas pentadecane is preferably formed over NiMo/-
Al
2
O
3
catalyst, pointing to different reaction route of HDO. The tentative reaction scheme of methyl palmitate
transformation was considered in comparison with ones obtained for methyl heptanoate hydrodeoxygenation.

Key-Words: - biofuels; hydrodeoxygenation; methyl palmitate; methyl heptanoate; CoMoS/-Al
2
O
3
;
NiMoS/-Al
2
O
3


1 Introduction
The decrease in available fossil fuel resources
induces an increasing interest for alternative and
renewable solutions like the use of biomass [1]. The
biomass itself may be divided into three basic
categories: starchy feedstock (including sugars),
triglyceride feedstock, and lignocellulosic feedstock
[2]. Oils are mainly comprised of triglycerides and
fatty acids. The oxygen content of bio-oil is high,
about 35%, that causes there deleterious properties
such as high viscosity, thermal and chemical
instability as well as lower heating values than fossil
sources [3]. The necessary upgrading of bio-oil, in
which the oxygen content is reduced, can be
achieved by hydrodeoxygenation (HDO), which
entails the elimination of oxygen and the formation
of normal paraffins with carbon chain lengths from
15 to 22 respectively.
HDO are considered important commercial routs
for producing distillate range bio-fuels from
triglyceride-based feed-stocks such as vegetable
oils, animal fats and non-edible greases, for example
a tall oils [4-8]. Promising strategies of upgrading
these bio-oils could be the direct hydroprocessing to
produce hydrocarbons or the co-hydrotreatment
with petroleum fractions, such as gas oil, to achieve
the technical and environmental fuel standards
[9-11]. The bio-oil co-hydrotreating process has
been approached by applying typical
hydrodesulfurization catalysts, such as sulfidied
CoMo and NiMo [12-15]. From the economic point
of view, co-processing of triglyceride-based feed-
stocks with petroleum fractions in existing refinery
units is a relatively low cost method to produce
renewable fuel.
In order to ensure successful operation, it is
imperative to understand and control the new types
of reactions that occur when higher levels of
oxygenates are processed. However, several
reaction pathways exist, and other reactions such as
saturation of double bonds and reactions involving
carbon monoxide and carbon dioxide complicate the
picture. Thus, a fundamental knowledge of the
detailed reaction chemistry is needed for the catalyst
development and the process design evaluation.
Methyl palmitate (MP) and methyl heptanoate (MH)
were chosen in order to model hydrotreating of
normal seed oils and animal fats, as this molecule
shares the main characteristics (an ester bonded
fatty acid) of the naturally occurring triglycerides.
The aim of the present work is a comparison of
the reactivity and the reaction pathways of methyl
palmitate and methyl heptanoate as the
representative model components of bio-oil in the
Recent Researches in Engineering and Automatic Control
ISBN: 978-1-61804-057-2 24
hydrodeoxygenation reactions depending of the
catalyst used (CoMo or NiMo) and the reaction
medium.


2 Experimental
Catalysts and chemicals. Two commercial
hydrotreatment catalysts were used: a cobalt-
molybdenum and a nickel-molybdenum both
supported on -Al
2
O
3
.
The metal contents in the catalysts were
determined by X-ray fluorescence spectrometry
(XRF) using a VRA-30 analyzer with a Cr-anode
(Karl Zeiss Jenna, Germany). Catalysts composition
is presented in Table 1 for non-sulfided catalysts.

Table 1. Chemical composition of the catalysts
CoMo/Al
2
O
3
NiMo/Al
2
O
3

Co (wt%) 2.90
Ni (wt%) 2.65
Mo (wt%) 11.0 11.2

Methyl palmitatate (grade 97 wt%) and methyl
heptanoate (grade 99 wt%) were supplied by
Aldrich. n-Tetradecane (analytical grade) was
supplied by Vekton (Russia) and m-xylene
(analytical grade) were supplied by Reakhim
(Russia). Dimethyldisulfide (DMDS, 99 wt%) was
purchased from Acros Organics. A hydrotreated
diesel fuel (HTDF) was obtained after standard
hydrodesulfurization procedure with the following
characteristics: 8 ppm S, 7 ppm N, 21.1 wt% mono-,
1.89 wt% di- and 0.81 wt% polyaromatics,
respectively (IP-391).
Catalytic Reactions. The catalytic HDO of
methyl palmitate (6 wt%) and methyl heptanoate (3
wt%) were performed in a batch reactor at the
temperature 300
o
C and hydrogen pressure 35 bars
over CoMo/-Al
2
O
3
and NiMo/-Al
2
O
3
, using n-
tetradecane, m-xylene or hydrotreated diesel fuel as
a solvent. DMDS (0.6 wt%) was added in the
reaction mixture (V = 100 ml) to generate a constant
partial pressure of H
2
S and maintain the sulfided
state of the catalyst. Prior to use, the catalysts (m =
0.3 g) were activated by a standard presulfidation
procedure in the gas phase at 400
o
C under a mixture
of a 5 vol.% H
2
S/95 vol.%H
2
. The sampling was
carried out in the course of reaction in the definite
intervals.
Product analysis. The reaction products were
identified by GC/MS technique (Agilent
Technologies 7000 GC/MS Triple QQQ GC System
7890A) using HP-5MS quartz capillary column
(30m0.25mm) and quantified by gas
chromatography system (Agilent 6890N) equipped
with HP-1MS column (60m0.32mm0.25m) and
atomic emission detector (GC-AED). The total
oxygen content in the reaction mixture was
determined using CHNSO elemental analyzer Vario
EL Cube (Elementar, Germany). For
determination of H
2
O content produced in the
reaction mixture the coulometric Karl Fischer
titration (coulometer DL39 KF, Mettler Toledo
GmbH) was used.


3 Results and Discussion
Effect of the catalyst. The series of experiments
were performed using methyl palmitate as the
feedstock to compare CoMo and NiMo catalysts
activity. The results obtained are represented on
Fig.1 and Fig. 2. A complete conversion of oxygen-
containing compounds was achieved in 4 or 6 hours
according to total oxygen conversion data. The
CoMo/-Al
2
O
3
catalyst exhibited higher catalytic
activity in the methyl palmitate HDO than the
NiMo/-Al
2
O
3
using n-tetradecane as a solvent,
whereas no difference in activity was observed
when hydrotreated diesel fuels was used.
According to GC/MS data for both applied
catalytic systems it was observed that methyl
palmitate transformed into oxygen- and sulfur-
containing compounds, such as aldehyde, alcohol,
alkanethiols, carboxylic acid and ester-type
compounds, at 35 bar approximately 0.5 wt% of
cycloparaffins and no aromatic products were
found. The formation of alkanethiols proved the
literature data [16] that propose the mechanism of
alkanethiol formation from the primary alcohol by
the SN2 nucleothilic substitution mechanism.

Fig. 1. Kinetic curves: the total oxygen content in
the reaction mixture vs. time during methyl
palmitate HDO over NiMo/Al
2
O
3
and CoMo/Al
2
O
3
catalysts in n-tetradecane.
0 1 2 3 4 5 6
0.00
0.15
0.30
0.45
0.60
0.75


Time, h
O
x
y
g
e
n

c
o
n
t
e
n
t
,

w
t
.
%
NiMoS/Al
2
O
3
CoMoS/Al
2
O
3
Recent Researches in Engineering and Automatic Control
ISBN: 978-1-61804-057-2 25
0 1 2 3 4 5 6
0,00
0,15
0,30
0,45
0,60
0,75
Time, h
O
x
y
g
e
n

c
o
n
t
e
n
t
,

w
t
.
%


NiMoS/Al
2
O
3
CoMoS/Al
2
O
3
Fig. 2. Kinetic curves: the total oxygen content in
the reaction mixture vs. time during methyl
palmitate HDO over NiMo/Al
2
O
3
and CoMo/Al
2
O
3
catalysts in HTDF.

The concentration profiles of major product are
represented on Fig. 3 and Fig. 4. It was observed
that all liquid hydrocarbon products had 15 or 16
carbon atoms, and that the most abundant ones were
hexadecanoic acid, 1-hexadecanol, n-C15 and n-
C16 alkanes and the corresponding alkenes, but also
1-hexadecanal. This product distribution verifies the
existence of the two routes leading to n-C15 and n-
C16. The only products associated with the
decarboxylation route were C15 alkenes and
alkanes, and no oxygenate intermediates were
detected. However, the HDO route leading to C16
products appeared to proceed by a more
complicated mechanism, as several intermediates
with 16 carbon atoms were detected. The first step
of a simple reaction scheme would be a stepwise
hydrogenation of the connecting oxygen in the ester
forming an aldehyde, which is hydrogenated to the
alcohol and then to the alkane, or possibly water is
split from the alcohol, forming an alkene prior to the
alkane. After 6 hours methyl palmitate completely
transformed into saturated and unsaturated C15 and
C16 hydrocarbons. The main product obtained over
CoMo/-Al
2
O
3
catalyst was hexadecane whereas
pentadecane is preferably formed over NiMo/-
Al
2
O
3
catalyst, pointing out to the different reaction
route of methyl palmitate HDO.
Effect of the solvent. In the literature the authors
investigate the HDO of the fatty acids methyl esters
model compounds using different solvents: p-xylene
[17, 18], n-hexadecane [19, 20], dodecane [16]. At
industrial conditions the catalyst has to demonstrate
an excellent catalytic activity in the presence of light
gas oil. We investigate the effect of solvent nature
on the HDO catalyst activity.
0 1 2 3 4 5 6
0
20
40
60
80
100


P
r
o
d
u
c
t
s

c
o
n
t
e
n
t
,

w
t
.
%
Time, h
C
15
H
30
C
16
H
32
C
16
H
34
C
15
H
32
palmitic
acid
methyl
palmitate
Fig. 3. Effect of the reaction time on the products
distribution during methyl palmitate HDO over
NiMo/Al
2
O
3
catalyst in n-tetradecane.
0 1 2 3 4 5 6
0
20
40
60
80
100
P
r
o
d
u
c
t
s

c
o
n
t
e
n
t
,

w
t
.
%
Time, h


C
16
H
34
C
15
H
32
palmitic
acid
methyl
palmitate
C
16
H
32
C
15
H
30
hexadacanal
Fig. 4. Effect of the reaction time on the products
distribution during methyl palmitate HDO over
CoMo/Al
2
O
3
catalyst in n-tetradecane.

The main difference in the solvent composition:
hydrotreated diesel fuel contains 21.1% mono-,
1.89% di- and 0.81% polyaromatics ( 8 ppm S, 7
ppm N). It was found that the activity of the
CoMo/-Al
2
O
3
and NiMo/-Al
2
O
3
catalysts in
methyl palmitate HDO in the presence of different
solvents decreases in order:
n-Tetradecane > HTDF m-Xylene.
Reducing the catalytic activity of the catalysts in
methyl palmitate HDO using as a solvent m-xylene
or HTDF, containing appreciable quantities of
aromatic compounds, testify the inhibition of some
reaction stage by the aromatic components. The
solvent possess the complicate effect on the reaction
products distribution, particularly during methyl
palmitate HDO over CoMo/-Al
2
O
3
and NiMo/-
Al
2
O
3
catalysts at the presence of m-xylene the
benzene ring alkylation products had been found in
Recent Researches in Engineering and Automatic Control
ISBN: 978-1-61804-057-2 26
the reaction mixture. This phenomenon requires a
more careful investigation in the future.

Fig. 5. Effect of the fatty acids methyl esters carbon
chain on the initial HDO rate (W
0
HDO
) of methyl
palmitate (MP) and methyl heptanoate (MH) over
CoMo/Al
2
O
3
catalyst in n-tetradecane.

Effect of the fatty acids methyl esters carbon
chain. We use methyl palmitate and methyl
heptanoate as model reaction compounds to
compare how the length of the carbon chain can
effect on the HDO rate. Methyl heptanoate has been
usually used as model compound to investigation of
fatty acid methyl esters HDO over conventional
sulfided catalysts [16, 21, 22], whereas the effect of
the fatty acid carbon chain length on the HDO rate
also as methyl palmitate HDO were studied
insufficiently. We observed that the initial HDO rate
of methyl heptanoate was about two times higher
than the initial HDO rate for methyl palmitate
(Fig.5). Apparently with the increasing the fatty acid
carbon chain the rate of reaction decreases because
of the presence of steric limitations.
Using as a solvent n-tetradecane we have studied
in detail the main intermediates of methyl palmitate
HDO over sulfided CoMo/-Al
2
O
3
and NiMo/-
Al
2
O
3
catalysts. Given our results and literature data
[21, 23] it was assumed a possible reaction
pathways of methyl palmitate transformation under
applied conditions that shown in Fig. 6. Esters can
be converted to carboxylic acids by means of
hydrolysis, which is the reverse of the acid-
catalysed esterification reaction by which a
carboxylic acid and an alcohol form an ester and
water. In addition, acid-base reactions are known to
take place at high temperature on sulfided
hydrotreating catalysts. The second route for
hexadecanoic acid formation may be the acid-
catalysed hydrolysis of methyl palmitate on the
sulfided NiMo and CoMo catalysts. The alumina
support was earlier found to catalyze the formation
of carboxylic acids and alcohols from aliphatic
methyl esters. It is therefore not excluded that the
alumina support, with its Lewis acid sites, may
participate in the hydrolysis of the ester.
MP MH
0
4
8
12
16
W
o
H
D
O
*
1
0
6
,

m
o
l
/
(
g
c
a
t

s
)
C
13
H
27
H
H
H
H
O
OCH
3
methyl palmitate
pentadecenes
C
13
H
27
H
H
H
-CH
4
-CO
2
decarboxylation
[1]
de-esterification [2]
C
15
H
31
OH
O
palmitic acid
-CO
2
-H
2
O [1]
C
14
H
29
H
H
O
H
hydrodeoxygenation
[3]
+H
2
-CH
3
OH
hexadecanal
CH
3
H
H
OH
H
H
hexadecanol
H
C
14
H
29
H
H
-H
2
O
hexadecenes
+H
2
n-C
15
H
32
product from
decarboxylation path
-CO
2
[1]
+H
2
-H
2
O
[3]
n-C
16
H
34
hydrodeoxygenation path
product from
+H
2
+H
2
-H
2
O
+H
2
Fig. 6. Possible reaction pathways of the HDO reactions of methyl palmitate over sulfided CoMo/-Al
2
O
3

and NiMo/-Al
2
O
3
catalyst.
Recent Researches in Engineering and Automatic Control
ISBN: 978-1-61804-057-2 27
Hexadecanol did not accumulate in the reaction
mixture indicating that it was formed as an
intermediate in the reactions from the most likely
route hydrogenation of the palmitinic acid to an
alcohol via an aldehyde. Thus, hexadecanol might
have been formed by the hydrogenation of
hexadecanal, which was detected in only trace
amounts presumably due to its high reactivity. In a
study of deoxygenation of stearic acid on various
supported metal catalysts, Snare and coworkers [24]
found that saturated and unsaturated hydrocarbons
were produced by decarboxylation and
decarbonylation reactions, respectively. CO
2
and
CO were formed as the side products. Similarly,
haxadecanoic acid might produce saturated and
unsaturated C16 hydrocarbons on the sulfided NiMo
and CoMo catalysts. Moreover, cleavage of C-C
bond in hexadecanal results in C15 hydrocarbons.


4 Conclusion
The results obtained let us come to several
conclusions. The hydrodeoxygenation of methyl
palmitate and methyl heptanoate as the model
compounds of bio-oil in the presence of sulfided
CoMo/-Al
2
O
3
and NiMo/-Al
2
O
3
catalysts was
studied. The CoMo/-Al
2
O
3
catalyst exhibited
higher catalytic activity in the methyl palmitate
HDO than the NiMo/-Al
2
O
3
using n-tetradecane as
a solvent. The main product obtained over CoMo/-
Al
2
O
3
catalyst was hexadecane whereas pentadecane
is preferably formed over NiMo/-Al
2
O
3
catalyst,
pointing to different reaction route of HDO. The
proposed reaction scheme for methyl palmitate
transformations is corresponding with the literature
data [21] for methyl heptanoate HDO studied using
m-xylene as a solvent.
The activity of the CoMo/-Al
2
O
3
and NiMo/-
Al
2
O
3
catalysts in methyl palmitate HDO in the
presence of different solvents increases in order:
HTDF m-xylene < n-tetradecane.

Acknowledgments
The research work was supported by RFBR Grant
No. 11-03-00611. The authors would like to thank
Utkin V.A. for product identification by the GC/MS
technique.

References:
[1] D.M. Alonso, J.Q. Bond, J.A. Dumesic,
Catalytic conversion of biomass to biofuels,
Green Chemistry, Vol.12, 2010, pp. 1493-1513.
[2] G.W. Huber, A. Corma, Synergies between bio-
and oil refineries for the production of fuels
from biomass, Angewandte Chemie
International Edition, Vol.46, 2007, pp. 7184-
7201.
[3] G. Knothe, Biodiesel and renewable diesel: A
comparison, Progress in Energy and
Combustion Science, Vol.36, 2010, pp. 364-373.
[4] J. Jakkula, J. Akkula, V. Niemi, J. Nikkonen,
V.-M. Purola, J. Myllyoja, P. Aalto, J.
Lehtonen, V. Alopaeus, Process for producing a
hydrocarbon component of biological origin, EP
1,396,531 (A2), 2004.
[5] W. Vermeiren, F. Bouvart, N.Dubut, A process
for the production of bio-naphtha from complex
mixtures of natural occurring fats and oils, EP
2,290,045 (A1), 2011.
[6] T. Hamamatsu, H. Ono, Y. Iguchi, H. Iki, Y.
Kinoshita, Process for production hydrocarbon
oil, EP 2,333,030 (A1), 2011.
[7] A. Guzman, J.E. Torres, L.P. Prada1, Manuel L.
Nunez, Hydroprocessing of crude palm oil at
pilot plant scale, Catalysis Today, Vol.156,
2010, pp. 38-43.
[8] M. Toba, Y. Abe, H. Kuramochi, M. Osako, T.
Mochizuki, Y. Yoshimura, Hydrodeoxygenation
of waste vegetable oil over sulfide catalysts,
Catalysis Today, Vol.164, 2011, pp. 533-537.
[9] G.W. Huber, P. OConnor, A. Corma,
Processing biomass in conventional oil
refineries: Production of high quality diesel by
hydrotreating vegetable oils in heavy vacuum
oil mixtures, Applied Catalysis A: General,
Vol.329, 2007, pp. 120-129.
[10] S. Melis, Albemarle catalytic solutions for the
co-processing of vegetable oil in conventional
hydrotreaters, Catalysts Courier, No.73, 2008,
pp. 6-8.
[11] J. Walendziewski, M. Stolarski, R. uny, B.
Klimek, Hydroprocesssing of light gas oil
rape oil mixtures, Fuel Processing Technology,
Vol.90, 2009, pp. 686-691.
[12] P. imacek, D. Kubicka, G. ebor, M. Pospiil,
Fuel properties of hydroprocessed rapeseed oil,
Fuel, Vol.89, 2010, pp. 611-615.
[13] S. Bezergianni, A. Kalogianni, I.A. Vasalos,
Hydrocracking of vacuum gas oil-vegetable oil
mixtures for biofuels production, Bioresource
Technology, Vol.100, 2009, 3036-3042.
[14] R. Nava, B. Pawelec, P. Castano, M.C.
Alvarez-Galva, C.V. Loricera, J.L.G. Fierro,
Upgrading of bio-liquids on different
mesoporous silica-supported CoMo catalysts,
Applied Catalysis B: Environmental, Vol.92,
2009, pp. 154-167.
[15] Ch. Templis, A. Vonortas, I. Sebos, N.
Papayannakos, Vegetable oil effect on gasoil
Recent Researches in Engineering and Automatic Control
ISBN: 978-1-61804-057-2 28
HDS in their catalytic co-hydroprocessing,
Applied Catalysis B: Environmental, Vol.104,
2011, pp. 324-329.
[16] E.-M. Ryymin, M.L. Honkela, T.-R. Viljava,
A.O.I. Krause, Insight to sulfur species in the
hydrodeoxygenation of aliphatic esters over
sulfided NiMo/-Al
2
O
3
catalyst, Applied
Catalysis A: General, Vol.358, 2009, pp. 42-
48.
[17] M. Ferrari, B. Delmon, P. Grange, Influence of
the impregnation order of molybdenum and
cobalt in carbon supported catalysts for
hydrodeoxygenation reactions, Carbon, Vol.40,
2002, pp. 497-511.
[18] G. de la Puente, A. Gil, J.J. Pis, P. Grange,
Effects of support surface chemistry in
hydrodeoxygenation reactions over
CoMo/activated carbon sulfided catalysts,
Langmuir, Vol.15, 1999, pp. 5800-5806.
[19] E. Laurent, B. Delmon, Study of the
hydrodeoxygenation of carbonyl, carboxyl.ic
and guaiacyl groups over sulfided CoMo/-
A1
2
0
3
and NiMo/-A1
2
0
3
catalysts. I. Catalytic
reaction schemes, Applied Catalysis A,
Vol.109, 1994, pp. 77-96.
[20] A. Centeno, E. Laurent, B. Delmon, Influence
of the support of CoMo sulfide catalysts and of





























the addition of potassium and platinum on the
catalytic performances for the
hydrodeoxygenation of carbonyl, carboxyl, and
guaiacol-type molecules, Journal of Catalysis,
Vol. 154, 1995, 288-298.
[21] O.I. Senol , E.-M. Ryymin, T.-R. Viljava,
A.O.I. Krause, Reactions of methyl heptanoate
hydrodeoxygenation on sulphided catalysts,
Journal of Molecular Catalysis A: Chemical,
Vol. 268, 2007, pp. 1-8.
[22] O.I. Senol, T.-R. Viljava, A.O.I. Krause, Effect
of sulphiding agents on the
hydrodeoxygenation of aliphatic esters on
sulphided catalysts, Applied Catalysis A:
General, Vol.326, 2007, pp. 236-244.
[23] B. Donnis, R.G. Egeberg, P. Blom, K.G.
Knudsen, Hydroprocessing of Bio-Oils and
Oxygenates to Hydrocarbons. Understanding
the Reaction Routes, Topics in Catalysis,
Vol.52, 2009, pp. 229-240.
[24] M. Snare, I. Kubickova, P. Maki-Arvela, D.
Chichova, K. Eranen, D.Yu. Murzin, Catalytic
deoxygenation of unsaturated renewable
feedstocks for production of diesel fuel
hydrocarbons, Fuel, Vol.87, 2008, pp. 933-945.
Recent Researches in Engineering and Automatic Control
ISBN: 978-1-61804-057-2 29

You might also like