You are on page 1of 8

Catal Lett (2010) 134:250–257

DOI 10.1007/s10562-009-0248-9

Diesel-like Hydrocarbons from Catalytic Deoxygenation


of Stearic Acid over Supported Pd Nanoparticles
on SBA-15 Catalysts
S. Lestari • P. Mäki-Arvela • K. Eränen •
J. Beltramini • G. Q. Max Lu • D.Yu. Murzin

Received: 16 October 2009 / Accepted: 3 December 2009 / Published online: 22 December 2009
Ó Springer Science+Business Media, LLC 2009

Abstract Palladium catalysts with different loading on The next generation of biofuels focuses on diesel-like
SBA-15 have been prepared using a direct synthesis hydrocarbons. Thus, it can be utilized directly using current
method and characterized with X-ray diffraction, micro- combustion engines without engine modification, as well as
graph images, and X-ray photoelectron spectroscopy. avoiding the ethical issue of using biomass source compet-
Mesoporous Pd-SBA-15 was an active and selective cata- ing with food sources, as is the case of ethanol derived from
lyst for deoxygenation of stearic acid in dodecane as a agricultural crops. One possible route is via catalytic deox-
solvent at 300 °C under 17 bar of 5 vol% H2 in argon as a ygenation of fatty acids and their derivatives into diesel-like
carrier gas in a semibatch reactor. hydrocarbons which have been recently developed [3–7].
The advantage of deoxygenation route is that no hydrogen is
Keywords Deoxygenation  SBA-15  Fatty acids required which makes the process more economically fea-
sible compared to an alternative hydrotreating process.
The catalytic deoxygenation of different fatty acids has
1 Introduction been intensively investigated, ranging from both saturated
and un-saturated C12 to C22 acids, as well as esters and
The amount of primary energy produced in the world was triglycerides [8, 9]. Catalyst screening experiments
500 EJ (5 9 1024 J) in 2005 with 86.5% from fossil fuels. revealed that Pd supported on either microporous or mes-
The energy consumption is projected to expand by 50% in oporous carbon catalyst is the most promising catalyst used
2030, therefore demand of energy is increasing continu- in deoxygenation reaction [6]. Typically, a complete con-
ously worldwide [1]. Since fossil fuels are non-renewable version of the representative model compound of fatty
and the source is limited, the development of new tech- acids, i.e. stearic acid was achieved at 300 °C under 17 bar
nologies for fuel production becomes a great challenge. of 5 vol% H2 in argon, producing n-heptadecane as the
Apart from fossil sources, energy can be also derived from main product. The off-line gas analysis showed that CO
biomass. Few technologies have been developed for the and CO2 are the main gaseous products, confirming that
conversion of biomass into fuels, such as pyrolysis, gasi- deoxygenation occurs via decarbonylation and decarbox-
fication of synthesis gas followed by Fischer–Tropsch ylation. Therefore, utilization of a semibatch reactor is
route, as well as transesterification of fatty acid and their preferable in order to purge out the gaseous products since
derivatives to fatty acid methyl ester (FAME) [2]. CO is well known for poisoning the metal catalyst surface.
Supported metal nanoparticles (SMNPs) have attracted
attention over the past decade due to their superior physi-
S. Lestari  J. Beltramini  G. Q. Max Lu
ARC Centre of Excellence for Functional Nanomaterials, cochemical properties compared to their bulk materials.
University of Queensland, Brisbane, QLD 4072, Australia The ability to control the particle size, shape, and metal
dispersion as well as the advantage of nanoparticle aggre-
S. Lestari  P. Mäki-Arvela  K. Eränen  D.Yu. Murzin (&)
gation inhibited by immobilization/separation on the het-
Process Chemistry Centre, Åbo Akademi University, 20500
Turku, Finland erogeneous support offer the potential for enhanced
e-mail: dmurzin@abo.fi selectivity and activity in catalytic applications [10, 11].

123
Diesel-Like Hydrocarbons from Catalytic Deoxygenation of Stearic Acid 251

Many immobilization methods and support materials have Zhang et al. and Selvaraj et al. [18, 19]. In a typical syn-
been reported in the literature. Han et al. described syn- thesis, 4.0 g of P123 was dispersed in 30 g of de-ionized
thesis of Pd/SBA-15 by one step synthesis method through water. After stirring the mixture for 4 h, 70 mL of 0.29 M
sol–gel reaction leading to a catalyst which exhibited high HCl were added and the solution was stirred for 2 h. Then,
catalytic activity and better recycling in the Suzuki and 9 g of TEOS and an appropriate amount of palladium
Heck coupling reactions [12]. chloride were added under stirring to the above homoge-
In this work, Pd nanoparticles on SBA-15 by direct neous solution. After further stirring at 40 °C for 24 h, the
synthesis have been applied as a catalyst for deoxygenation resulted gel (molar composition of TEOS:PdCl2:P123:
of stearic acid. The main parameter under study is the HCl:H2O & 1:0.0052–0.0748:0.016:0.47:129) was trans-
effects of metal loading. The supported Pd nanoparticles ferred to a Teflon-lined autoclave for hydrothermal pro-
that have large support surface area and porosity are cessing at 100 °C for 48 h. The obtained mixtures were
reported to be more suitable for catalytic conversion of washed with de-ionized water, filtered, and dried at 120 °C
larger fatty acid molecules [13–15]. overnight. The obtained powders were calcined at 550 °C
for 5 h. The nominal Pd metal loading was calculated
based on weight ratio of palladium to silica which gave
2 Experiment Section weight percentage in range 0.3–5 wt%. The final sample
then denoted as xPd-SBA-15 in which x represent Pd metal
2.1 Materials loading in wt%.

The following chemicals were used in preparation of SBA-


15 support and Pd/SBA-15 catalyst: Pluronic P123 triblock 2.4 Catalyst Characterization
copolymer (EO20PO70EO20), (Sigma–Aldrich) with aver-
age molecular weight, Mn 5,800 and tetraethyl orthosilicate The ICP-AES technique was used to determine the palla-
(TEOS), (Aldrich, purity [99%) as the organic template dium content in the final samples. Before analysis, the
and silica sources, respectively. Palladium chloride anhy- samples were digested in a mixture of HF, HCl and HNO3
drous, Fluka (purity [60% of Pd) was used as a metal using a CEM MDS2000 microwave digester at 68% power
precursor. In the deoxygenation experiments, a pure, 99% for 20 min. The digested samples were analyzed using a
stearic acid (Sigma–Aldrich) was used as a feedstock and Varian Vista Pro ICP-AES instrument running at 1,200 W
dodecane (Fluka, purity [98%) was applied as a solvent. forward power.
N,O-bis(trimethyl)-trifluoroacetamide, BSTFA was used as The crystal phase and pore structure of samples were
a silylation agent. All chemicals were used as received. determined by X-ray diffractometer (XRD, Bruker/AXS
D8 Advance) operating at 40 kV using Cu Ka irradiation
2.2 Synthesis of SBA-15 (k = 0.15406 nm). The XRD patterns were recorded in the
2h range of 0.5–6o for small angle profiles and 10–90o for
SBA-15 was synthesized according to the procedures wide-angle profiles. Surface area and pore size distribu-
reported previously [16, 17]. A homogeneous mixture tions as well as adsorption–desorption isotherms were
containing EO20PO70EO20 and TEOS in hydrochloric acid measured via N2 adsorption using Quadrasorb with all
solution was first stirred in a beaker glass at 40 °C for 2 h. samples out-gassed at 120 °C for 6 h before measurement.
Then the resultant gel was transferred to a Teflon-lined The images of samples were recorded by SEM F6300 and
autoclave. After the hydrothermal treatment at 100 °C for TEM JEOL 1010 operating at 100 kV.
24 h, the solid product was recovered by filtration, fol- X-ray photoelectron spectroscopy (XPS, Kratos Axis
lowed by washing with de-ionized water and drying at Ultra) was employed to determine the O 1s, Pd 3d, and Si
50 °C in vacuum. The removal of organic template was 2p binding energies of surface oxygen, palladium, and
done by ethanol-solvent extraction using an excess amount silicon species with Mono Al Ka (hm = 1486.6 eV) as the
of ethanol (1 g of SBA-15/150 mL ethanol) and the mix- excitation source. The survey and high resolution scans
ture was stirred vigorously for 5 h at 60 °C. After being done at analyzer pass energy of 160 and 20 eV, respec-
treated with ethanol for three times, finally the powder was tively. The C 1s peak at 284.6 eV was taken as a reference
dried overnight at 100 °C. for binding energy calibration.
The metal dispersion of supported palladium catalysts
2.3 Direct Synthesis of Pd-SBA-15 (In Situ Method) were investigated by CO-chemisorption. Detail procedure
and setup used in analysis are described elsewhere [23].
Palladium anchored SBA-15 with various palladium load- The molar ratio between Pd:CO assumed to be unity
ing were synthesized by in situ method adopted from [13].

123
252 S. Lestari et al.

2.5 Catalytic Deoxygenation Reaction

The catalytic deoxygenation of stearic acid was performed


in a semi batch reactor coupled with a condenser and
heating mantle with the liquid phase volume of 100 mL.
The flow of carrier gas and reaction pressure inlet and
outlet were controlled by a flow (Brookle 58505 S) and a
pressure controller (Brookle 5866), respectively. The stir-
ring speed was maintained suitably high, at 1,100 rpm to
prevent external mass transfer limitation. The catalysts
were reduced in situ by hydrogen for 1 h prior to the cat-
alytic test at constant temperature and pressure, 60 °C and
4.8 bar, respectively. In a typical reaction, the catalyst and
feed amount were used as follows: 0.5 g of catalyst;
0.05 M feed (1.4 g of stearic acid, 99% Sigma–Aldrich in
100 mL of dodecane). The reactor was then flushed thor-
oughly with an inert gas. Before the reaction started
25 mL/min of reaction atmosphere gas was passed through
the reactor until reaching the reaction pressure. Thereafter,
the reaction temperature was adjusted with heating ramp
10 °C/min to the desired reaction temperature. At this point
stirring commenced, which was considered as the begin-
ning of the reaction. After the reaction, the catalyst parti-
cles were filtered and washed with acetone for further
characterization. A schematic diagram of the semi batch
reactor used in these experiments is shown in Fig. 1.

2.6 Liquid Samples Analysis

Liquid phase samples were withdrawn from the reactor


vessel via a sampling valve during the experiment. Typi-
cally, the samples had to be dissolved in pyridine and si- Fig. 2 Small (a) and wide (b) angle XRD pattern of xPd-SBA-15
lylated with N,O-bis(trimethyl)-trifluoroacetamide, BSTFA synthesized by direct synthesis method with nominal palladium
loading = (a) 0.3 (b) 0.6 (c) 1.5 (d) 3 and (e) 5 wt%
in order to be analyzed by GC. Generally 100 wt% excess

Fig. 1 A schematic diagram of


semi batch reactor used for
deoxygenation experiment

123
Diesel-Like Hydrocarbons from Catalytic Deoxygenation of Stearic Acid 253

Table 1 Actual palladium content, metal dispersion and average crystallite size, and physical properties of Pd supported catalyst
Samplea Actual Pd Pd dispersionc Pd particle Pd particle SfBET (m2/g) Dgp (nm) Vgp (cm3/g)
contentb (wt%) sized (nm) sizee (nm)

SBA15 – 850 6.8 1.4


Direct synthesis
0.3 wt% Pd-SBA-15 0.26 68.0 2.0 3.5 921 5.6 1.3
0.6 wt% Pd-SBA-15 0.38 54.0 2.3 2.2 838 5.1 1.1
1.5 wt% Pd-SBA-15 1.05 30.7 4.1 3.5 934 5.3 1.2
3 wt% Pd-SBA-15 2.92 28.2 4.0 4.3 (3.9h) 870 5.6 (5h) 1.2
5 wt% Pd-SBA-15 3.74 15.0 6.4 4.3 834 5.6 1.2
a
Nominal palladium loading as denoted in sample xPd-SBA-15
b
Determined by ICP-AES
c
Measured by CO chemisorption
d
Measured by CO chemisorption
e
Calculated based on wide angle XRD using Scherrer equation
f
Calculated by BET method
g
Calculated from desorption branch by BJH method
h
Determined from TEM

of BSTFA were added to the sample. After addition of the SBA-15 as reported by Zhao [17]. These peaks still could
silylation agent, the samples were kept in an oven at 60 °C be clearly observed at high loading of Pd although their
overnight. The internal standard eicosane, C20H42 was intensities are varied to some extent. The results indicated
added for quantitative calculations. The samples were that the well ordered pore structures of SBA-15 are still
analyzed with a gas chromatograph (GC, HP 5890) retained in xPd-SBA-15 samples suggesting that the
equipped with non-polar column (HP-1, with dimension of anchoring of Pd2? during synthesis of mesoporous frame-
60 m 9 0.32 mm 9 0.5 lm) and a flame ionization work has no significant effect on the formation of mesop-
detector. About 1 lL sample was injected into the GC with ores structures.
split ratio 50:1, and the carrier gas (helium) flow rate was Figure 2b shows a wide XRD pattern of xPd-SBA-15.
137 mL/min. The injector and detector temperature were The results demonstrated that at wide diffraction angles,
280 and 290 °C, respectively. The following temperature 10–70o all the samples show a relatively sharp peak at
program was used for analysis: 130, 169 °C (1 °C/min) 2h = 33–34o assignable to metallic Pd. The intensity of
hold for 15 min, 246 °C (5 °C/min) hold for 3 min and this peak increased with increasing of Pd content. This
300 °C (10 °C/min). The gas chromatographic method was peak is already observed for the sample with Pd loading as
calibrated using the following chemicals: n-heptadecane low as 0.3 wt%, however, the exact location of Pd crys-
(Acros, 99%) and stearic acid (Sigma–Aldrich, 99%). The tallites is unclear. Therefore, more sensitive analysis such
product identification was validated with a gas chromato- as XPS and morphology imaging were utilized to identify
graph-mass spectrometer (GC–MS). the position of Pd particles.
The physical properties of xPd-SBA-15 and SBA-15 are
summarized in Table 1. It is seen that SBA-15 sample
3 Results and Discussion possesses high surface area and large pore volume of
850 m2/g and 1.4 cm3/g, respectively, with an average pore
3.1 Phase Structure and Physical Properties diameter of 6.8 nm. This value is slightly higher from those
of Supported Pd Catalysts reported previously [13, 17]. For the xPd-SBA-15 samples,
with anchoring of Pd ions during synthesis, the surface area
The small and wide angle of XRD patterns of xPd-SBA-15 varies in the range 834–934 m2/g, whereas pore volume
synthesized by direct synthesis method are shown in Fig. 2. and average pore diameter seems to be quite similar,
It is observed, that almost all samples exhibit one intense around 1.2 cm3/g and 5.3 nm, respectively.
peak at 0.84–0.96o along with two peaks at 1.4–1.5o and The N2 adsorption–desorption isotherm of xPd-SBA-15
1.6–1.7o, which correspond to 100, 110, and 200 reflec- are shown in Fig. 3a, which exhibited type-IV isotherms
tions, respectively (Fig. 2a). These XRD patterns verify the with a regular type H1 hysteresis loop at relative pressure
formation of a typical hexagonal structure of mesopores (P/P0) of 0.6–0.8. These isotherm features are quite similar

123
254 S. Lestari et al.

these numbers with pore size of SBA-15, which are larger,


it can be tentatively concluded that palladium particles are
anchored inside the pore structure of SBA-15. However,
the reliability of metal particle size calculated by Scherer
equation is questionable as TEM shows bigger particle size
of Pd, ca. 7 nm.
The Si 2p, O 1s and Pd 3d XPS spectra of the xPd-SBA-
15 samples are shown in Fig. 4. For the xPd-SBA-15
samples, the symmetric signals of Si 2p remained almost
unchanged at the binding energy of 103.7 eV (Fig. 4a),
whereas the O 1s signals shifted from BE = 533.2 eV for
Pd-SBA-15 with Pd loading = 0.3 wt% to BE = 532.4 eV
for Pd-SBA-15 with Pd loading = 5 wt%. The reason for
the binding energy shift is most probably an interaction
between O and Pd. Moreover, the decline in O 1s peak is
also observed with an increase in Pd content (Fig. 4b).
According to the literature [20–22], the Pd 3d5/2 peak
binding energy at 335 eV is associated with Pd(0) whereas
the Pd 3d3/2 at 337.5–338 eV are characteristic for Pd(II).
For the xPd-SBA-15 samples, the latter two peaks of Pd are
slightly shifted for the sample with Pd loading 0.5–5 wt%.

3.2 Morphology and Pore Structure

Electron microscopic techniques were used for investiga-


tion of surface morphologies and pore structure of
supported Pd samples. Figure 5 shows the SEM images of
xPd-SBA-15, as well as parent SBA-15. It is observed that
SBA-15 possesses a typical long-chain architecture
Fig. 3 a N2 adsorption–desorption isotherm of xPd-SBA-15. Nota- (Fig. 5a); the pore structure is regular and highly ordered
tion is the same as in Fig. 2; b Pore size distribution of xPd-SBA-15
with a pore size of ca. 5 nm and wall thickness of 6–7 nm.
There was a good correlation of the pore sizes determined
with increasing Pd loading suggesting that anchoring of Pd from TEM images and using BJH method (see Sect.3.1).
ion by direct synthesis does not affect the mesoporous For xPd-SBA-15 prepared by the direct synthesis method,
channels of SBA-15. the banana-like morphology was still retained in the sample
The pore size distributions of xPd-SBA-15 calculated with Pd loading 3 wt% (Fig. 5c).
from the desorption branch by Barrett-Joiner-Halenda
method gave a narrow curve with a maximum at 7.0 nm, 3.3 Catalytic Deoxygenation of Stearic Acid
although the intensity were slightly lower with the
increasing of Pd loading on the surface as shown in As reported previously, supported palladium catalysts were
Fig. 3b. found to be the most active and selective catalyst in deox-
Metal dispersion of supported palladium catalysts ygenation of fatty acid and derivatives to produces diesel-
gradually decreased by increasing metal loading, as could like hydrocarbons. 6In this work, we aim to examine the
be expected (Table 1). Average metal crystallite sizes possibility of palladium supported on SBA-15 catalyst for
determined from CO chemisorption, TEM and XRD agree deoxygenation reaction of stearic acid. As mentioned above
very well for the 3 wt% Pd-SBA-15 (Table 1), while some utilization of mesoporous support SBA-15 could give a
discrepancy could be observed in the cases of 0.3 wt% Pd beneficial effect for the large molecules as the fatty acids.
and 5 wt% Pd supported on SBA-15. For the latter catalyst The catalytic deoxygenation of stearic acid over sup-
the average Pd crystallite size was overestimated by 75% ported palladium catalyst resulted in n-heptadecane as the
due to the fact that the metal crystallites smaller than 3 nm main product. Trace amounts of isomer n-C17, i.e. 3-hep-
cannot be determined by XRD. tadecene and/or 8-heptadecene were also formed in the
The average Pd particles size calculated by Scherrer liquid phase. A reactant and product concentration profiles
equation gave the range 2.2–4.3 nm. When comparing are shown in Fig. 6.

123
Diesel-Like Hydrocarbons from Catalytic Deoxygenation of Stearic Acid 255

Fig. 4 XPS spectra of xPd-SBA-15 (a, b, c) with Pd loading: (a) 0.6, (b) 1.5, (c) 3, and (d) 5 wt%

Fig. 5 SEM and TEM images,


respectively of: (a, b) parent
SBA15; (c, d) 3 wt%Pd-SBA-
15

The effect of palladium metal loading was investigated metal taking into account metal dispersion were calculated.
by conducting experiments over catalysts with different The highest initial TOF was 0.72 s-1 over 3 wt% Pd,
metal loading. The initial TOFs, defined as the amount of whereas Pd 0.3, 0.6 and 5 wt% catalysts exhibited initial
converted stearic acid divided by time and mole of surface TOF values of 0.16, 0.27 and 0.14 s-1, respectively.

123
256 S. Lestari et al.

A 100

80

Conversion (%)
60

40

20

0 1 2 3 4 5 6 7 8
Time *Mass of Pd (min*g)
Fig. 6 a Kinetics in the deoxygenation of stearic acid over: 3 wt%
Pd-SBA-15. The reaction conditions as follows: Cstearic acid = B
0.049 M (1.4 g in 100 mL of dodecane); masscatalyst = 0.5 g; 100
T = 300 °C; P = 17 bar in 5 vol% H2 in argon

Selectivity to heptadecane (%)


90

The lowest TOF over palladium catalyst with the highest 80


loading of 5 wt% could be associated with the low metal
dispersion. Analogously to these results, an optimum metal 70
dispersion for deoxygenation of stearic and palmitic acid at
300 °C under 17.5 bar H2/Ar (5 vol% H2) was established 60

in [13]. It was stated in [13] that the small Pd particles


50
exhibited strong interaction with the support and thus their
structure was not suitable for deoxygenation of fatty acids,
40
while 6 nm Pd particles were too large to be very active in
fatty acid deoxygenation [13]. Comparison of the TOFs for 0 20 40 60 80 100
the 3 wt% Pd/SBA-15 with the one achieved over a Conversion (%)
microporous Pd/C under the same reaction conditions using
Fig. 7 a Conversion of stearic acid as a function of normalized time,
5 vol% H2 in argon as a carrier gas in a semibatch reactor b selectivity to heptadecane as a function of conversion in the
3
the results were as follows: In the current case the stearic deoxygenation of stearic acid over: 3 wt% Pd-SBA-15, b) the effect of
acid to Pd mass ratio was 933 resulting in a TOF of metal loading on the deoxygenation product over supported Pd
0.72 s-1 and 96% conversion during 300 min reaction time catalysts. The reaction conditions as follows: Cstearic acid = 0.049 M
(1.4 g in 100 mL of dodecane); masscatalyst = 0.5 g; T = 300 °C;
whereas for 5 wt% Pd/C the mass ratio of stearic acid to Pd P = 17 bar in 5 vol% H2 in argon. Symbols: (open square) 0.3 wt%,
was 900 giving TOF of 0.126 s-1 and 62% conversion of (filled triangle) 0.6 wt%, (open circle) 3 wt% and (filled square) 5
stearic acid after 360 min3 thus showing that the Pd/SBA- wt% Pd
15 catalyst was more efficient than Pd/C in deoxygenation
of stearic acid. which was already confirmed in the deoxygenation of
The catalytic activities after prolonged reaction times stearic acid over microporous Pd/C [6]. In [6] the BET
are depicted as a function of normalized time in Fig. 7a. surface area of the spent catalyst decreased indicating
Stearic acid deoxygenation was very fast over 0.3 and 0.6 coking of the catalyst. The formation of unsaturated C17
wt% Pd-SBA-15 catalysts, whereas lower deoxygenation compounds was more prominent in ethyl stearate deoxy-
rates were achieved over 3 and 5 wt% Pd-SBA-15 cata- genation under hydrogen scarce conditions [5]. It should be
lysts, respectively. The conversion levels after 5 h reaction pointed out here that ethyl stearate deoxygenation proceeds
time decreased in the following order: 0.6 wt% Pd [ 3 via formation of stearic acid. In the current work the
wt% Pd  0.3 wt% Pd  5 wt% Pd (Table 2). In all cases deoxygenation atmosphere applied was 5 vol% H2 in
the catalysts were active after prolonged reaction times, as argon, which thus promoted the formation of unsaturated
can be seen from Fig. 7a. Catalyst deactivation occurred C17 hydrocarbons.
during the deoxygenation of stearic acid mainly due to the The selectivity to the main deoxygenation product,
formation of unsaturated compounds, such as heptadecene, n-heptadecane as a function of conversion is shown in

123
Diesel-Like Hydrocarbons from Catalytic Deoxygenation of Stearic Acid 257

Table 2 Effect of various Pd loading on selectivity and conversion of stearic acid. Reaction conditions as follow: T = 300 °C; P = 17 bar; gas
atmosphere = 5% H2/Ar; cSA,0 = 0.049 M (1.49 g in 100 mL of dodecane); masscatalyst = 0.5 g
Catalyst ConversionSAa % Selectivity (n-C17)
b
% Selectivity (X = 60%)
c
%

Direct synthesis method


0.3 wt% Pd-SBA-15 64 98 98
0.6 wt% Pd-SBA-15 96 98 96
3.0 wt% Pd-SBA-15 94 94 80
5.0 wt% Pd-SBA-15 41 95 84
a
Conversion of stearic acid after 5 h reaction
b
Selectivity to main product of n-heptadecane after 5 h reaction
c
Selectivity to main product of n-heptadecane at conversion of 60%

Fig. 7b, demonstrating that the formation of n-heptadecane 2. Gerpen JV (2005) Fuel Process Technol 86:1097
was over 95% after prolonged reaction times. Initially the 3. Kubicková I, Snåre M, Eränen K, Mäki-Arvela P, Murzin DY
(2005) Catal Today 106:197
formation rate of n-heptadecene was high, but since there 4. Lestari S, Simakova I, Tokarev A, Mäki-Arvela P, Eränen K,
was hydrogen present in the reaction atmosphere, most of Murzin DY (2008) Catal Lett 122:247
the formed n-heptadecene was hydrogenated to saturated 5. Mäki-Arvela P, Kubickova I, Snåre M, Eränen K, Murzin DY
n-heptadecane during the course of the reaction. (2007) Energy Fuels 21:30
6. Snåre M, Kubickova I, Mäki-Arvela P, Eränen K, Murzin DY
(2006) Ind Eng Chem Res 45:5708
7. Snåre M, Kubicková I, Mäki-Arvela P, Eränen K, Wärnå J,
4 Conclusions Murzin DY (2007) Chem Eng J 134:29
8. Murzin DY, Kubickova I, Snåre M, Mäki-Arvela P, Myllyoja J
(2009) Method for manufacture of hydrocarbons. US Patent
The palladium nanoparticles anchored in SBA-15 with 7491858, 1990
various metal loading have been synthesized by direct in 9. Snåre M, Kubicková I, Mäki-Arvela P, Chichova D, Eränen K,
situ method. The well ordered structures of SBA-15 were Murzin DY (2008) Fuels 87:933
still retained giving surface area in the range of 834– 10. White RJ, Luque R, Budarin VL, Clark JH, Macquarrie DJ (2008)
Chem Soc Rev 38:481
921 m2/g and average pore size ca. 5.6 nm. 11. Astruc D, Lu F, Aranzaes JR (2005) Angew Chem Int Ed 44:7852
Catalytic deoxygenation of stearic acid was successfully 12. Han P, Wang X, Qiu X, Ji X, Gao L (2007) J Mol Catal A Chem
performed over these materials. The most active catalyst 272:136
was 3 wt% Pd-SBA-15 exhibiting high TOF of 0.72 s-1 at 13. Simakova I, Simakova O, Mäki-Arvela P, Simakov A, Estrada M,
Murzin DY (2009) Appl Catal A Gen 355:100
300 °C, under 17 bar of 5 vol% H2 in argon, using 0.5 g of 14. Simakova I, Simakova O, Romanenko AV, Murzin DY (2008)
catalyst and 0.049 M of reactant (in dodecane). The main Ind Eng Chem Res 47:7219
deoxygenation product was n-heptadecane with over 90% 15. Mbaraka IK, Shanks BH (2005) J Catal 229:365
selectivity above 95% conversion levels. Furthermore, 16. Zhao D, Chmelka BF, Stucky GD (1998) Science 279:548
17. Zhao D, Huo Q, Feng J, Chmelka BF, Stucky CD (1998) J Am
trace amounts of unsaturated C17 isomers, such as 3- and/ Chem Soc 120:6024
or 8-heptadecene were formed. 18. Selvaraj M, Kawi S (2007) Chem Mater 19:509
19. Zhang L, Zhao Y, Dai H, He H, Au CT (2008) Catal Today
Acknowledgments Siswati Lestari gratefully acknowledges the 131:42
research facilities and funding provided by ARC Centre of Excellence 20. Chen J, Zhang Q, Wang Y, Wan H (2008) Adv Synth Catal
for Functional Nanomaterials, University of Queensland, Australia. 350:453
The authors are grateful to Mr. Sandy Budihartono for performing 21. Rossi LM, Silva FP, Vono LLR, Kiyohara PK, Duarte EL, Itri R,
micrographs analysis of TEM and SEM. The XPS analysis performed Landers R, Machado G (2007) Green Chem 9:379
by Dr. Barry Wood from Centre of Microscopy and Microanalysis, 22. Wagner CD, Naumkin AV, Kraut-Vass A, Allison JW, Powell
University of Queensland is gratefully acknowledged. This work is part CJ, Rumble JR (2007) NIST X-ray photoelectron spectroscopy
of activities at the Åbo Akademi Process Chemistry Centre of Excel- database. National Institute of Standard and Technology (NIST)
lence Programmes (2000–2011) financed by the Academy of Finland. Technology Services, United States of America
23. Tanksale A, Beltramini J, Dumesic JA, Lu GQJ (2008) J Catal
258:366
References

1. International Energy Outlook (2008) Energy information admis-


tration. URL: http://www.eia.doe.gov/oiaf/ieo/

123

You might also like