Você está na página 1de 10

Regulating product distribution in deoxygenation of methyl laurate

on silica-supported NiMo phosphides: Effect of Ni/Mo ratio


Jixiang Chen
a,
, Yan Yang
a
, Heng Shi
a
, Mingfeng Li
b
, Yang Chu
b
, Zhengyi Pan
a
, Xinbin Yu
a
a
Tianjin Key Laboratory of Applied Catalysis Science and Technology, Department of Catalysis Science and Engineering, School of Chemical Engineering and Technology,
Tianjin University, Tianjin 300072, China
b
Research Institute of Petroleum Processing, SINOPEC, 18 Xue Yuan Road, 100083 Beijing, China
h i g h l i g h t s
Ni/Mo ratio determines acidity, dispersion and phosphide phase.
There is an electron transfer from Ni to Mo.
Catalyst activity and product distribution can be tuned by altering Ni/Mo ratio.
Catalyst with Ni/Mo ratio of 1 shows particular performance.
a r t i c l e i n f o
Article history:
Received 20 November 2013
Received in revised form 16 March 2014
Accepted 23 March 2014
Available online 4 April 2014
Keywords:
Metal phosphide
Acidity
Dispersion
Hydrodeoxygenation
Decarbonylation
a b s t r a c t
SiO
2
-supported Ni
2
P, MoP and NiMo bimetallic phosphides with different Ni/Mo ratios were investi-
gated for the deoxygenation of methyl laurate to C11 and C12 hydrocarbons. They were characterized
by means of N
2
sorption, X-ray diffraction, transmission electron microscope, CO chemisorption, X-ray
photoelectron spectroscopy and NH
3
temperature-programmed desorption. In the NiMo bimetallic
phosphide, the NiMoP
2
phase was formed apart from Ni
2
P and MoP, and the incorporation of Mo into
Ni
2
P took place. These led to an interaction between Ni and Mo via the electron transfer from Ni to
Mo. In addition, the increase in the Ni/Mo ratio tended to reduce the phosphide dispersion and catalyst
acidity. In the deoxygenation, the turnover frequency of methyl laurate and the C11/C12 ratio tended to
increase as the Ni/Mo ratio increased (apart from Ni/Mo ratio of 1). This is related to not only the different
catalytic roles of Ni and Mo sites but also the interaction between Ni and Mo and the phosphide disper-
sion. In all, the C11/C12 ratio can be regulated by altering the Ni/Mo ratio. The catalyst acidity obviously
affected the distributions of the oxygenated intermediates.
2014 Elsevier Ltd. All rights reserved.
1. Introduction
Nowadays, fossil energy is fast being consumed, which simulta-
neously causes serious environmental problem. To develop clean
and sustainable energy has drawn great attention in globe. As a
kind of renewable resource, vegetable oils and animal fats (com-
plex mixture of triglycerides) have been commercially used to pro-
duce biodiesel (namely fatty acid methyl esters) via
transesterication. Recently, as an alternative to biodiesel, the die-
sel-like hydrocarbons derived from vegetable oils or animal fats via
catalytic hydrotreating have been rapidly developed because of
their similar fuel property to petroleum-based diesel, high
oxidation stability, high cetane number as well as the economical
efciency of production [1]. Also, the diesel-like hydrocarbons
can be further cracked to produce jet fuel. There are two possible
pathways for the deoxygenation of fatty ester, this is, hydrodeoxy-
genation (HDO) and decarboxylation/decarbonylation. The HDO
pathway leads to the hydrocarbon with same number of carbon
atoms to the corresponding fatty acid, and oxygen is removed as
water. The decarboxylation/decarbonylation pathway gives rise
to the hydrocarbon with one carbon less than the corresponding
fatty acid, and oxygen is removed as CO
2
/CO. In comparison with
HDO pathway, decarboxylation/decarbonylation pathway offers
the advantage of lower hydrogen consumption if the methanation
of CO
2
/CO does not occur; however, it also gives lower carbon
yield. If hydrogen is insufcient, the decarbonylation pathway is
preferential. Otherwise, the HDO pathway is expected from the
point of enhancing carbon yield. Additionally, the decarbonylation
hydrocarbon may be more suitable for producing jet fuel via its
http://dx.doi.org/10.1016/j.fuel.2014.03.049
0016-2361/ 2014 Elsevier Ltd. All rights reserved.

Corresponding author. Tel.: +86 22 27890865; fax: +86 22 87894301.


E-mail address: jxchen@tju.edu.cn (J. Chen).
Fuel 129 (2014) 110
Contents lists available at ScienceDirect
Fuel
j our nal homepage: www. el sevi er . com/ l ocat e/ f uel
further cracking. Therefore, it is very important in practice to
regulate deoxygenated hydrocarbon composition and it can be
achieved by tuning catalyst property and reaction conditions.
The investigated hydrotreating catalysts mainly include suldes
(e.g., NiMo and CoMo suldes) and metals (such as noble ones
and Ni) [1,2]. The NiMo and CoMo suldes mainly give HDO
hydrocarbons, whereas metal catalysts mostly produce decarbony-
lation ones. However, for the sulde catalysts, sulphiding agent
(e.g. H
2
S or CS
2
) must be added to the feedstock to maintain the
catalyst stability. This leads to forming undesirable S-containing
products and increasing investment. The high cost limits the appli-
cation of noble metal. Metallic Ni catalysts also show good perfor-
mance, however, they are also very active for cracking and
methanation [14], leading to a decrease in the hydrocarbon yield
and an increase in the H
2
consumption. Recently, other active
phases have been explored for the hydrotreating of fatty ester
and fatty acid, such as transition metal carbide [3,4], nitride [5]
and phosphide [610]. They exhibit some advantages over suldes
or noble metals. For instance, they are used under the S-free con-
dition and have much lower costs than noble metals. Therefore,
it is very signicant to further develop these materials for the
deoxygenation of fatty esters.
Since Oyama et al. [11] found the excellent performance of MoP
for hydrodenitrogenation (HDN), the transition metal phosphides
have attracted considerable attention for hydrotreating process be-
cause of their high activities for HDN and hydrodesulfurization
(HDS) [12,13]. Last several years, a new use of sulfur-free metal
phosphides has also been focused on the HDO of bio-oils [1419]
as well as fatty ester and fatty acid [6,710]. The ligand and ensem-
ble effects of P can account for the special performance of metal
phosphide in the hydrodeoxygenation, such as lower activity of
Ni
2
P than that of metallic Ni for the hydrogenolysis of CC bond,
methanation and decarbonylation [79]. Ni, Co, Fe, Mo and W
phosphides exhibit very different deoxygenation behaviors [8]. In
the deoxygenation of methyl laurate, Ni, Co and Fe phosphides
mainly produce the decarbonylation hydrocarbons, while MoP
and WP primarily give the HDO hydrocarbons. Among these
phosphides, Ni
2
P and MoP are more active than the others, and they
give mainly decarbonylation and HDO hydrocarbons, respectively.
How the NiMo bimetallic phosphide performs in the deoxygenation
of fatty ester is worth investigating. This is illuminated from that the
bimetallic NiMo catalyst shows very different performance from the
monometallic Ni and Mo catalysts in other elds.
The NiMo bimetallic effect in suldes has been investigated on
the deoxygenation of fatty ester [2022]. It has been found that
sulded Ni and Mo yield almost decarboxylation/decarbonylation
and HDO hydrocarbon products, respectively, while the sulded
NiMo yields a mixture of decarboxylation and HDO hydrocarbons
[21,22]. Also, Ni has a promoting effect on the activity of MoS
2
[22].
NiMo bimetallic phosphides have not been studied for the deox-
ygenation of fatty ester, while they exhibit some particular perfor-
mance for HDS [23,24], HDN [25] and HDO of anisole [19]. Unlike
sulde and carbide, no synergetic effect was observed between the
phosphided Ni and the Mo atoms for HDS [23]. The HDS turnover
frequency (TOF) decreased in the order of Ni
2
P > Ni
x
MoP > MoP,
and it increased with Ni content for Ni
x
MoP [23,24]. Similar case
was also found for the HDO TOF of anisole [19]. However, increas-
ing Ni content reduced the HDN TOF but enhanced the selectivity
to hydrogenated products [25]. To our knowledge, there have been
no reports about the NiMo bimetallic effect for the phosphides on
the deoxygenation of fatty esters.
In this work, our aim is to explore and understand the NiMo
bimetallic effect in phosphide on the deoxygenation of methyl lau-
rate as a model compoundto hydrocarbons. For this, SiO
2
-supported
Ni
2
P, MoP and NiMo bimetallic phosphides with different Ni/Mo
ratios were prepared, characterized and evaluated. We analyzed
the inuence of the Ni/Mo ratio on the catalyst surface and bulk
structures as well as the correlation between structure and perfor-
mance. The result indicates that the deoxygenation pathway can
be tuned via altering the Ni/Mo ratio. This is benecial to the insight
into the structureactivity relationship of the metal phosphide
catalysts and the deoxygenation mechanism. Also, this provides an
informantion for designing catalyst compositions to produce
products with different hydrocarbon compositions to meet the
practical demand.
2. Experimental
2.1. Catalyst preparation
The SiO
2
-supported Ni
2
P, MoP and NiMo bimetallic phosphide
catalysts were prepared by the H
2
temperature-programmed
reduction (H
2
-TPR) method from the phosphate precursors. The
supported phosphate precursors were prepared by the successive
impregnation. The detail for the catalyst preparation is presented
in the supplementary information. Two series of phosphide cata-
lysts were prepared (see Table 1). One is to increase the Mo content
when the mass ratio between Ni and SiO
2
was xed as 12%, and the
another is to increase the Ni content when the mass ratio between
Mo and SiO
2
was xed as 19.6 wt.%. As a result, Ni
2
P/SiO
2
, MoP/SiO
2
and SiO
2
-supported NiMo bimetallic phosphides with different Ni/
Mo molar ratios were prepared. In addition, Ni
2
P/SiO
2
and MoP/
SiO
2
had the similar metal moles. For convenience, the NiMo
bimetallic phosphide catalysts are denoted as NiMoa/b, where a/b
indicates the Ni/Mo molar ratio.
2.2. Catalyst characterization
H
2
-TPR, CO chemisorption and NH
3
temperature-programmed
desorption (NH
3
-TPD) were used to characterizing the reducibility
of catalyst precursor, the surface density of metal site and the
catalyst acidity. The details for those are given in the supplementary
information.
N
2
adsorption was measured on a Quantachrom QuadraSorb SI
at 196 C. The BrunauerEmmettTeller (BET) equation was used
to calculate the surface area. X-ray diffraction (XRD) patterns were
obtained on a D8 Focus powder diffractometer using Cu Ka radia-
tion (k = 0.15406 nm). High resolution transmission electron
microscope (HRTEM) images were obtained on a Tecnai G2 F20
instrument.
X-ray photoelectron spectroscopy (XPS) was performed on a
PHI 5000VersaProbe instrument with Al Ka radiation (1486.6 eV).
Binding energies were determined with adventitious carbon (C1s
at 284.8 eV) as the reference. Reduced catalysts were transferred
to the bottles in a glovebox and stored in sealed bottles with Ar
atmosphere. When the measurement was carried out, the catalysts
were transferred to the analysis chamber using a glovebox with N
2
atmosphere. That is, the reduced catalysts were avoided to be
exposure to air before the XPS measurement.
2.3. Activity test
The catalyst activity for the deoxygenation of methyl laurate
was tested on a stainless-steel xed-bed reactor (inner diameter
of 12 mm). 0.35 g passivated catalyst blended with 2.8 g quartz
sand in same diameter was loaded in the reactor, and 2 g quartz
sand was placed on the catalyst bed to preheat the reactants.
Before the reaction, the passivated catalyst was re-reduced at
450 C for 1 h. After that, the temperature was adjusted to the de-
sired value, the H
2
pressure was set at 3.0 MPa, and methyl laurate
was fed into the reactor using a pump. The weight hourly space
2 J. Chen et al. / Fuel 129 (2014) 110
velocity (WHSV) of methyl laurate was 14 h
1
and the H
2
/methyl
laurate molar ratio was 50. The liquid products were quantitatively
analyzed on a SP-3420 gas chromatograph equipped with a ame
ionization detector (FID) and a HP-5 capillary column
(30 m 0.32 mm 0.5 lm). Tetrahydronaphthalene was used as
an internal standard. The analytic results were conrmed by a
gas chromatograph (GC) with the relative standard deviations
lower than 5.0%. The gaseous products (CO and CH
4
) were quanti-
ed on an on-line 102 GC equipped with a TCD and a TDX-101
packed column, and N
2
was used as an internal standard. The mass
balance ((the mass of liquid efuent from reactor per hour+the
mass of gaseous products per hour)/the mass of methyl laurate
entered reactor per hour) was better than 97%.
The conversion of methyl laurate (X) and the selectivity to
product i (S
i
) were dened as follows:
X n
0
n=n
0
100%; S
i
n
i
=n
0
n 100%
where n
0
and n denote the moles of methyl laurate in the feed and
the product, respectively; n
i
denotes the mole of methyl laurate
converted to product i.
In addition, the turnover frequency (TOF) of methyl laurate and
the site time yield (STY) of the C11 and C12 hydrocarbons were cal-
culated as follows:
TOF(s
1
) = Amount of converted methyl laurate per second
(lmol/s
1
)/(CO uptake (lmol/g) catalyst weight(g)).
STY(s
1
) = Amount of produced C11 and C12 hydrocarbons per
second (lmol/s
1
)/(CO uptake (lmol/g) catalyst weight(g)).
3. Results and discussion
3.1. Catalyst characterization
Fig. 1 shows the H
2
-TPR proles of the catalyst precursors. The
prole for MoP/SiO
2
precursor shows a small peak at 515 C and a
big peak at 715 C. The former is due to the reduction of Mo
6+
?
Mo
4+
, while the latter is ascribed to the reductions of Mo
4+
?Mo
and phosphate [26]. The Ni
2
P/SiO
2
precursor gives lower initial
reduction temperature than the MoP/SiO
2
one. The small peak ob-
served at 235 C is due to the reduction of Ni
3+
?Ni
2+
. The obvious
peaks at about 470, 605 and 736 C are ascribed to the reductions
of NiO and/or nickel silicate, nickel species in phosphates and PO
bond [8], respectively. Because the Ni
2
P/SiO
2
precursor was not
calcined after the introduction of phosphate, it showed a very dif-
ferent reduction behavior from the calcined one [8]. In the calcined
Ni
2
P/SiO
2
precursor, a strong interaction between nickel species
and phosphate leads to no reduction due to NiO and only the
reduction ascribed to nickel phosphate. The reductions of NiMo
phosphide catalyst precursors were complex. In contrast to the
reductions of Ni
2
P/SiO
2
and MoP/SiO
2
precursors, the peaks below
600 C are related to the reductions of NiO, nickel silicate and
Mo
6+
?Mo
4+
, while the ones above 600 C are related to the metal
(Ni and Mo) phosphates. On the whole, the Ni introduction
promoted the reduction. This is due to the easier reducibility of
Ni
2+
than that of Mo
6+
. The reduced metal Ni can dissociate H
2
to
form reactive atomic hydrogen.
Fig. 2 shows the XRD patterns of different catalysts. Ni
2
P
(2h = 40.8, 44.6, 47.4 and 54.5, PDF 03-0953) and MoP
(2h = 28.1, 32.3, 43.3 and 57.4, PDF24-0771) were detected for
Ni
2
P/SiO
2
and MoP/SiO
2
, respectively. For the Ni
2
P/SiO
2
catalysts
with the introduction of Mo, as the Ni/Mo ratio decreased from
3/1 to 3/3, the peaks due to Ni
2
P became weak and shifted to
low angles (obviously shown in Fig. 1S (2h = 3848) in supple-
mentary information), and the peaks due to MoP and MoNiP
2
(2h = 31.3 and 45.3, PDF 65-1985) appeared and became intense.
This indicates that some Mo atoms entered the lattice of Ni
2
P
through a homogeneous substitution of Ni with Mo because the
radius (0.145 nm) of Mo atom is larger than that (0.135 nm) of
Ni one. Also, the incorporation of Mo in the lattice of Ni
2
P promoted
the dispersion of Ni
2
P and/or reduced the crystallinity of Ni
2
P. This
is also reected by the following HRTEM images. For the MoP/SiO
2
catalysts with the introduction of Ni, as the Ni/Mo ratio increased
from 1/3 to 3/3, the peaks due to MoP and MoNiP
2
became intense,
while the peaks due to Ni
2
P only became slightly intense and also
located at lower 2h than those of Ni
2
P/SiO
2
. Clearly, the introduc-
tion of Ni promoted the formation of MoP phase with high crystal-
linity and large crystallite (also conrmed by HRTEM images). This
Table 1
Catalyst properties.
Catalyst Ni content
a
(wt.%) Mo content
a
(wt.%) Surface composition
b
BET surface area (m
2
/g) CO uptake (lmol/g) Acid amount (lmol/g)
c
Ni/Mo P/(Ni + Mo)
Ni
2
P/SiO
2
12.0 0.0 4.76 349 30 285
NiMo3/1 12.0 6.5 1.10 1.79 316 34 276
NiMo3/2 12.0 13.0 280 58 436
NiMo3/3 12.0 19.6 0.68 1.42 299 44 333
NiMo2/3 8.0 19.6 268 57 445
NiMo1/3 4.0 19.6 0.41 1.10 271 98 428
MoP/SiO
2
0.0 19.6 0.92 267 95 570
a
Relative to SiO
2
.
b
Obtained from XPS measurement.
c
Obtained from NH
3
-TPD measurement.
200 400 600 800 1000
H
2

c
o
n
s
u
m
p
t
i
o
n
/
a
.
u
.
Ni
2
P/SiO
2
NiMo3/1
NiMo3/2
NiMo3/3
NiMo2/3
NiMo1/3
MoP/SiO
2
Temperture/
o
C
Fig. 1. H
2
-TPR proles of different catalyst precursors.
J. Chen et al. / Fuel 129 (2014) 110 3
should be related to the role of Ni. During the H
2
-TPR process to
prepare the catalyst, the metallic Ni was more easily formed than
the metallic Mo, and it could adsorb and dissociate H
2
to formreac-
tive atomic hydrogen, which spilled over to Mo species and pro-
moted their reduction. As a result, MoP was easily formed and its
crystallite was apt to grow up. In all, the MoNiP
2
phase was the
most obvious at the Ni/Mo ratio of 1 (i.e., in the NiMo3/3 catalyst).
Fig. 3 shows the HRTEM images of different catalysts. The dark
globular particles belong to metal phosphides, which is demonstrated
by the crystallographic planes due to phosphides in the magnied
images and the EDX data (Fig. 2S). Ni
2
P/SiO
2
, NiMo3/1, NiMo3/2
and NiMo3/3, whose phosphide particles distributed between 3
and 20 nm(particle size distribution shown in Fig. 3S in supplemen-
tary information), had the average particle sizes of about 12.3, 13.2,
5.5 and 8.8 nm, respectively. NiMo2/3 with the particles ranged
from 3 to 15 nm had the average size of about 7.0 nm. NiMo1/3
and MoP/SiO
2
mainly contained the particles with 24 nm. Clearly,
Ni
2
P/SiO
2
showed much lower dispersion than MoP/SiO
2
, which is
mainly due to the easier reduction of the Ni
2
P/SiO
2
precursor. In a
whole, there was a trend that the introduction of Mo to Ni
2
P/SiO
2
enhanced the phosphide dispersion, while the introduction of Ni
to MoP/SiO
2
reduced the phosphide dispersion.
Fig. 4 shows the XPS spectra in the P 2p, Ni 2p
3/2
and Mo 3d
regions. In P 2p XPS spectra, two peaks due to P 2p
3/2
(dash lines)
are observed at about 129 and 134 eV, respectively. The former is
ascribed to the reduced P in phosphides, while the latter is due to
PO
4
3
species that were resulted from the surface passivation and
the unreduced PO
4
3
[27]. The reduced P had lower electron bind-
ing energy than the element P (130.2 eV). This is attributed to
the electron transfer from metal atom to P one, leading to a small
negative charge on the P site (P
d
) and simultaneously a small po-
sitive charge on the metal site (i.e., Ni
d+
and Mo
d+
). In the Ni 2p
3/2
spectra of Ni
2
P/SiO
2
, two peaks at 852.7 and 856.6 eV are assigned
to Ni
d+
in Ni
2
P phase and Ni
2+
ions [27], respectively. The peaks due
to Ni
d+
in NiMo3/1, NiMo3/3 and NiMo1/3 were located at about
853.1 eV that was higher than that in Ni
2
P/SiO
2
, indicating that
the positive charge (d) of Ni
d+
in NiMo bimetallic phosphide was
larger. In the Mo 3d spectra of MoP/SiO
2
, the Mo 3d
5/2
peaks at
228.0, 229.2 and 232.6 eV are attributed to Mo
d+
in MoP, Mo
4+
and Mo
6+
ions [28], respectively. The peaks due to Mo
d+
in
NiMo1/3, NiMo3/3 and NiMo3/1 were respectively at about
227.8, 227.6 and 227.5 eV, lower than that in MoP/SiO
2
. Therefore,
the positive charge (d) of Mo
d+
in NiMo bimetallic phosphide was
smaller than that in MoP. This can be ascribed to the electron
transfer from Ni to Mo, which is consistent that the Paulings elec-
tronegativity (1.91) of Ni is lower than that (2.16) of Mo. The elec-
tron transfer from Ni to Mo indicates an interaction between Ni
and Mo in phosphide, which is derived from the introduction of
Mo atoms into Ni
2
P lattice and the formation of MoNiP
2
phase. This
interaction will inuence the catalyst performance. As to the value
of positive charge on Ni or Mo site in the phosphide, some
researchers have investigated. Based on the result of infrared spec-
troscopic technique of CO adsorption [23,29], the value of positive
charge (d) on Ni
d+
in Ni
2
P is determined as 0 < d < 1, while that on
Mo
d+
in MoP is considered to 0 < d < 2. Additionally, Bussell et al.
[30] assigned the binding energy of 228.4 eV to a Mo
d+
(0 < d 6 4)
species. According to the above reports, the values of positive
charge on Ni
d+
and Mo
d+
in the present catalysts are considered
as 0 < d < 1 and 0 < d < 4, respectively. The Ni
d+
site had higher elec-
tron density than the Mo
d+
site in the phosphide. This is consistent
with the DFT calculation [31].
Table 1 presents the surface composition measured by XPS. As
the Ni/Mo ratio decreased, the surface P/M(Ni + Mo) ratio de-
creased. Compared surface Ni/Mo ratio with the bulk one, the Mo
species were surface enrichment on NiMo3/1 and NiMo3/3, while
the Ni species were surface enrichment on NiMo1/3.
Table 1 also shows the surface areas of different catalysts. MoP/
SiO
2
gave lower surface area than Ni
2
P/SiO
2
. This may be due to
that the MoP particles were much smaller than the Ni
2
P ones.
The small particles might enter and subsequently block the sup-
port pores, reducing the surface area. In the order of Ni
2
P/SiO
2
,
NiMo3/1 and NiMo3/2, the decreased surface area is related to
the increased loading of metal phosphide. In the sequence of
MoP/SiO
2
, NiMo1/3 and NiMo2/3, the loading of metal phosphide
increased, whereas the surface area did not change obviously. This
can be explained by that the increased phosphide particles reduced
the blockage of pores to some extent. The larger surface area of
NiMo3/3 than those of MoP/SiO
2
, NiMo1/3 and NiMo2/3 may be
ascribed to its larger particles.
The CO uptake represents the surface density of metal site (see
Table 1). In the sequence of Ni
2
P/SiO
2
, NiMo3/1 and NiMo3/2, the
CO uptake increased. This is related to the increased metal content
and the decreased surface P/M(M = Ni + Mo) ratio and particle size.
The lower CO uptake of NiMo3/3 than that of NiMo3/2 may be
related to the larger phosphide particle size. In the order of MoP/SiO
2
,
NiMo1/3, NiMo2/3 and NiMo3/3, the total metal content increased,
whereas the CO uptake tended to decrease. This is due to the
increased surface P/M(M = Ni + Mo) ratio and metal phosphide
crystallites. In all, the catalyst with higher dispersion gave larger
CO uptake.
The catalyst acidity was measured by NH
3
-TPD (the proles
shown in Fig. 4S in supplementary information). Each prole
shows a main peak at about 200 C with a shoulder at higher tem-
perature. The acidity of metal phosphide catalyst is derived from
the phosphide itself and surplus phosphorus species. Both Lewis
and Brnsted acidities have been found on the phosphide catalysts
[8,32]. As indicated by the previous work [8], the acid sites on Ni
2
P/
SiO
2
include POH group and Ni
d+
site, while those on MoP/SiO
2
in-
clude POH, MoOH, Mo
d+
as well as Mo
n+
(n = 4 and 6). Therefore,
the acid sites on the NiMo bimetallic phosphide catalysts con-
tained POH, MoOH, Ni
d+
, Mo
d+
as well as Mo
n+
. The POH and
MoOH groups belong to Brnsted acidity, while Ni
d+
, Mo
d+
as well
as Mo
n+
are Lewis acid sites. The acid amounts obtained from
NH
3
-TPD are provided in Table 1. The larger acid amount of MoP/
SiO
2
than that of Ni
2
P/SiO
2
is mainly ascribed to the higher MoP dis-
persion. As the Mo content introduced into Ni
2
P/SiO
2
increased, the
acid amount tended to increased (apart from NiMo3/3). However,
as the Ni content introduced into MoP/SiO
2
increased, the acid
Fig. 2. XRD patterns of different catalysts.
4 J. Chen et al. / Fuel 129 (2014) 110
Fig. 3. TEM images of different catalysts.
J. Chen et al. / Fuel 129 (2014) 110 5
amount tended to decrease. On the whole, the smaller phosphide
particles gave rise to more acid sites.
3.2. Catalyst reactivity
On the phosphide catalysts, the deoxygenated products from
methyl laurate include C11 and C12 hydrocarbons (n-alkanes, iso-
alkanes and alkenes), which are generated from the decarbonyla-
tion and HDO pathways, respectively. The detected oxygenated
intermediates were lauric acid, dodecanal, dodecanol and lauryl
laurate. During the HDO pathway, several reactions (such as
hydrogenolysis, hydrolysis, hydrogenation and dehydration)
consecutively occurred, that is, methyl laurate ?laurate
acid ?lauraldehyde ?lauryl alcohol ?C12 hydrocarbons. Apart
from methyl laurate, the oxygenated intermediates can also be
converted to C11 hydrocarbons via decarbonylation [8]. The pro-
posed reaction pathway scheme is presented in Scheme 1. As indi-
cated previously [810], the active sites on the surface of
phosphide particles include Ni
d+
, Mo
d+
and POH group. Ni
d+
and
Mo
d+
mainly contribute to decarbonylation, hydrogenolysis and
hydrogenation, while the POH group primarily catalyzed the
hydrolysis, dehydration as well as esterication.
Similar to the previous result [8], there was a great deal of
alkenes produced on MoP/SiO
2
(alkene/n-alkane molar ratio
>0.37 as indicated in Table 1S in supplementary information),
which is due to the low hydrogenation ability of MoP. However,
Ni-containing catalysts (Ni
2
P/SiO
2
and NiMo bimetallic phos-
phide) gave much less amounts of alkenes (alkene/n-alkane molar
ratio <0.06 as shown in Table 1S in supplementary information).
This should be ascribed to the high hydrogenation ability of Ni
[8]. Indeed, the alkene/n-alkane molar ratio decreased as the
Ni/Mo ratio increased.
3.2.1. Activity
Herein, the catalyst activity is presented as both total conver-
sion and turnover frequency (TOF) of methyl laurate (Fig. 5). As
was expected, the total conversion and TOF on each catalyst in-
creased as the reaction temperature. Generally, the NiMo
bimetallic phosphide catalysts had slightly higher conversion than
Ni
2
P/SiO
2
and MoP/SiO
2
. Among them, NiMo3/3 gave almost the
NiMo1/3
NiMo3/3
NiMo3/1
Ni
2
P
MoP
P 2p
R
e
l
a
t
i
v
e

i
n
t
e
n
s
i
t
y
/
a
.
u
.
Binding energy/eV
138 136 134 132 130 128 126 862 860 858 856 854 852 850
NiMo1/3
NiMo3/3
NiMo3/1
Ni
2
P
Ni 2p
3/2
R
e
l
a
t
i
v
e

i
n
t
e
n
s
i
t
y
/
a
.
u
.
Binding energy/eV
238 236 234 232 230 228 226 224
NiMo3/1
NiMo3/3
NiMo1/3
MoP/SiO
2
Mo 3d
R
e
l
a
t
i
v
e

i
n
t
e
n
s
i
t
y
/
a
.
u
.
Binding energy/eV
Fig. 4. XPS spectra of P 2p, Ni 2p
3/2
and Mo 3d in phosphide catalysts. For P 2p, solid lines and dotted lines correspond to P 2p
1/2
and P 2p
3/2
, respectively. For Mo 3d, solid
lines and dotted lines correspond to Mo 3d
3/2
and Mo 3d
5/2
, respectively.
6 J. Chen et al. / Fuel 129 (2014) 110
highest one. TOF displays a very different variation tendency from
the total conversion. Ni
2
P/SiO
2
and MoP/SiO
2
gave the highest and
the lowest TOF, respectively. TOF was calculated on the base of the
surface density of metal site (i.e., CO uptake). However, the catalyst
acidity also contributes to the conversion of methyl laurate. As
indicated in Scheme 1, methyl laurate can be converted via decarb-
onylation, hydrogenolysis and hydrolysis. The Brnsted acid site is
very low active for decarbonylation and hydrogenolysis [8]. The
decarbonylation and hydrogenolysis primarily occurred on metal
(Mo
d+
and Ni
d+
) sites, while the hydrolysis took place on acid site.
MoP/SiO
2
should be more active for hydrolysis than Ni
2
P/SiO
2
because of its more acid sites. Undoubtedly, per Ni site was
intrinsically more active than Mo
d+
site for the conversion of
methyl laurate. This can partially be explained by the higher
electron density of Ni
d+
site than that of Mo
d+
site in phosphide,
because the high electron density favors the cleavages of both
CC and CO bonds [8,33]. Additionally, because the Mo
d+
site
had larger positive charge (i.e., lower electron density) than the
Ni
d+
one, it was more electrophilic and so more easily combined
with oxygen in methyl laurate, the oxygenated intermediates as
well as the product water [8,20,34]. Moreover, the MoO bond
strength (502 kJ/mol) is much larger than the NiO one
(366 30 kJ/mol) [35], which is also indicated by that MoO
3
is
more difcultly reduced than NiO (see H
2
-TPR proles of
MoO
3
/SiO
2
and NiO/SiO
2
in Fig. 5S in supplementary information).
Therefore, it can be inferred that the Mo
d+
site has stronger inter-
action with oxygen than the Ni
d+
one. Contrary to Sabatier princi-
ple, the interaction between Mo and O may be too strong to
favor the desorption of O-containing products, subsequently giving
rise to low TOF. However, just as this strong interaction, the HDO
pathway rather than decarbonylation was preferentially catalyzed
on Mo site (indicated in the following).
For the NiMo bimetallic phosphide, TOF was contributed by
both Ni and Mo sites. On one hand, the introduction of Mo to
Ni
2
P/SiO
2
reduced TOF. With decreasing Ni/Mo ratio from 3/1 to
3/2, TOF decreased. However, NiMo3/3 had higher TOF than
NiMo3/2. On the other hand, the introduction of Ni to MoP/SiO
2
enhanced TOF. With increasing the Ni/Mo ratio from 1/3 to 3/3, TOF
increased. As a whole, TOF decreased with decreasing the Ni/Mo
ratio (apart from NiMo3/2). This is related to the Ni/Mo ratio, phos-
phide particle size and the electronic interaction between Ni and
Mo. The decrease in the Ni/Mo ratio led to low TOF because the
Ni site had higher TOF than the Mo one. In addition, the decrease
in the crystallite size may be another reason. The previous work
indicates that the larger Ni
2
P crystallite gives higher TOF [9].
Although NiMo3/2 had higher Ni/Mo ratio than NiMo3/3, it gave
lower TOF. This may be related to its smaller phosphide particle
size. Again, the electron transfer from Ni to Mo may also reduce
TOF on Ni site. Although the electron density of Mo site increased,
this might not compensate the effect of the decreased electron
density of Ni site.
It is worth noticing that the methyl laurate is not completely
converted to form the deoxygenated hydrocarbons (i.e., C11 and
C12 hydrocarbons) because the oxygenated intermediates were
produced. Here, we also provided the site time yield (STY) of the
C11 and C12 hydrocarbons (see Fig. 6S in supplementary informa-
tion). STY is dened as the number of molecules of both C11 and
C12 hydrocarbons produced per metal site and per unit time
[36]. It is reasonable that STY was always smaller than TOF at each
temperature. But then, as the reaction temperature increased, the
difference between STY and TOF decreased, indicating that the high
Scheme 1. Proposed reaction pathway for deoxygenation of methyl laurate.
Ni2P NiMo3/1 NiMo3/2 NiMo3/3 NiMo2/3 NiMo1/3 MoP
0
20
40
60
80
100
MoP/SiO
2
Ni
2
P/SiO
2
320
o
C 300
o
C
C
o
n
v
e
r
s
i
o
n
/
%
Catalyst
340
o
C
Ni2P NiMo3/1 NiMo3/2 NiMo3/3 NiMo2/3 NiMo1/3 MoP
0.0
0.1
0.2
0.3
0.4
0.5
MoP/SiO
2
Ni
2
P/SiO
2
T
O
F
/
s
-
1
Catalyst
300
o
C
320
o
C
Fig. 5. Conversion and TOF of methyl laurate on different catalysts.
J. Chen et al. / Fuel 129 (2014) 110 7
temperature favors the deoxygenation reaction. Also, similar to
TOF, STY decreased with decreasing Ni/Mo ratio (except NiMo3/2).
3.2.2. Product selectivity
It is known that decarbonylation is endothermic and HDO is
exothermic [37]. The increase of reaction temperature favors the
decarbonylation rather than HDO from the thermodynamic view,
while it accelerates both decarbonylation and HDO from the ki-
netic view. The deoxygenation pathway is affected not only by
reaction condition but also by catalyst property. Fig. 6 shows the
selectivities to C11 and C12 hydrocarbons and the C11/C12 molar
ratios on different catalysts. As the temperature increased from
300 to 340 C, the selectivity to C11 hydrocarbons on all catalysts
increased, whereas the selectivity to C12 hydrocarbons increased
on NiMo1/3 and MoP/SiO
2
. The temperature had very small effect
on the selectivity to C12 hydrocarbons on Ni
2
P/SiO
2
. For the cata-
lysts with the Ni/Mo ratios between 3/1 and 2/3, the selectivity
to C12 hydrocarbons rstly increased and then decreased. The
above results indicate that the decarbonylation to produce C11
hydrocarbons and the HDO to yield C12 hydrocarbons were more
favorable on Ni and Mo sites, respectively. This is more obviously
reected by the C11/C12 ratio.
The C11/C12 ratio represents the selectivity between the
decarbonylation and HDO pathways. As the temperature increased,
the C11/C12 ratio increased on all catalysts, especially for the cat-
alysts with high Ni/Mo ratios. The temperature had a very small
inuence on the C11/C12 ratio on MoP/SiO
2
. At all temperatures,
Ni
2
P/SiO
2
gave the largest C11/C12 ratio (from 2.9 at 300 C to
6.9 at 340 C), whereas MoP/SiO
2
gave the lowest C11/C12 ratio
(always lower than 0.1). Clearly, Ni
2
P/SiO
2
and MoP/SiO
2
preferen-
tially catalyzed decarbonylation and HDO pathways, respectively.
As discussed previously, this is mainly ascribed to their different
electronic properties and interaction with oxygen [8]. The Mo site
had lower electron density and stronger interaction with oxygen
than the Ni site. The metal site with lower electron density is more
electrophilic and so more easily interact with oxygen in the reac-
tant [8,20,34], preferentially giving rise to the HDO product (here-
in, C12 hydrocarbon).
For the NiMo bimetallic phosphide catalysts, the C11/C12 ratio
tended to decrease as the Ni/Mo ratio decreased (except NiMo3/3).
The introduction of Mo to Ni
2
P/SiO
2
reduced the C11/C12 ratio,
whereas the introduction of Ni to MoP/SiO
2
enhanced the C11/
C12 ratio. This is mainly attributed to the different deoxygenation
mechanism on Ni and Mo sites as mentioned above. In addition, as
indicated by previous work [9], larger Ni
2
P particles give higher
C11/C12. Interestingly, the C11/C12 ratio on the NiMo bimetallic
phosphide was lower than the weighted value of those on Ni
2
P/
SiO
2
and MoP/SiO
2
. As to NiMo3/1 that had similar average Ni
2
P
particle size to Ni
2
P/SiO
2
, for instance, the main phase was Ni
2
P
incorporated with Mo and the surface Ni/Mo ratio was about 1.1.
If the Ni and Mo sites on NiMo3/1 catalyzed the deoxygenation
as the same performance to those on Ni
2
P/SiO
2
and MoP/SiO
2
,
respectively, the C11/C12 ratio on NiMo3/1 would be larger than
1.0 at 300 and 320 C. In fact, the C11/C12 ratio was lower than
1.0. In other word, the introduction of Mo to Ni
2
P inhibited the
decarbonylation. This may be attributed to the electronic interac-
tion between Ni and Mo. Because of the electron transfer from Ni
to Mo, the increase in positive charge can enhance the electrophlic-
ity of Ni, which favored the HDO pathway on Ni site. Meanwhile,
the decrease in positive charge may reduce the electrophlicity of
Mo, promoting the decarbonylation pathway on Mo site. We spec-
ulate that the interaction between Ni and Mo may more obviously
inuence the deoxygenation pathway on Ni. Surprisingly, NiMo3/3
gave larger C11/C12 ratio than NiMo3/1 and NiMo3/2 although it
had lower Ni/Mo ratio. We temporarily attribute this special phe-
nomenon to the obvious MoNiP
2
phase in NiMo3/3.
Based on the above analysis, the C11/C12 ratio is affected by the
Ni/Mo ratio, the phosphide particle size, the interaction between Ni
and Mo as well as the role of MoNiP
2
phase. Among them, the Ni/
Mo ratio is the determining factor. In other word, the deoxygen-
ation pathway (i.e.,C11/C12 ratio) can be regulated by changing
the Ni/Mo ratio in the bimetallic phosphide catalyst. This provides
the approach to tuning the product composition. If the deoxygen-
ated hydrocarbons are used for producing jet fuel via further
cracking, the preferential reaction is decarbonylation. However, if
the diesel-like hydrocarbon is expected, the HDO pathway is
preferable. Notably, the introduction of Ni to MoP greatly reduced
the amount of alkenes that is not expected in the nal product as
fuel, whereas it did not alter the deoxygenation pathway essen-
tially. For example, NiMo1/3 mainly catalyzed the HDO pathway,
and the C11/C12 molar ratio was lower than 0.4 even at 340 C.
During the deoxygenation of methyl laurate, the detected oxy-
genated intermediates include lauric acid, dodecanal, dodecanol
and lauryl laurate. The selectivities to different oxygenated inter-
mediates are shown in Fig. 7. We found that the selectivity de-
creased in the sequence of dodecanol, lauryl lautate, lauric acid
and dodecanal. Dodecanal was very trace (the selectivity less than
0.4%, not shown here) because it is very reactive. The selectivity to
lauric acid was lower than 2% and was also not shown here. Our
previous work [8] shows that lauric acid was main oxygenated
intermediate on Ni
2
P/SiO
2
and MoP/SiO
2
under 2.0 MPa. Here,
the increased H
2
pressure (3.0 MPa) may account for that
catalyst
Ni2P NiMo3/1 NiMo3/2 NiMo3/3 NiMo2/3 NiMo1/3 MoP
0
20
40
60
80
MoP/SiO
2
Ni
2
P/SiO
2
S
e
l
e
c
t
i
v
i
t
y

t
o

C
1
2
/
%
0
20
40
60
80
100
S
e
l
e
c
t
i
v
i
t
y

t
o

C
1
1
/
%
300
o
C
320
o
C
340
o
C
Ni2P NiMo3/1 NiMo3/2 NiMo3/3 NiMo2/3 NiMo1/3 MoP
0
2
4
6
8
MoP/SiO
2
Ni
2
P/SiO
2
C
1
1
/
C
1
2

m
o
l
a
r

r
a
t
i
o
Catalyst
300
o
C
320
o
C
340
o
C
Fig. 6. Selectivities to C11 and C12 hydrocarbons and C11/C12 molar ratios.
8 J. Chen et al. / Fuel 129 (2014) 110
dodecanol became the main oxygenated intermediate. This is
reasonable because the increase in H
2
pressure promoted the
hydrogenation of lauric acid. The small amount of lauric acid is also
due to its subsequent reaction with dodecanol to form lauryl laurate.
As was expected, the selectivity to the total oxygenated inter-
mediates decreased with increasing temperature, and it was lower
than 0.25% at 340 C. It is reasonable that MoP/SiO
2
had higher
selectivity to dodecanol than Ni
2
P/SiO
2
because it predominately
catalyzed the HDO pathway. Interestingly, the NiMo bimetallic
phosphides generally gave higher selectivity to dodecanol than
MoP/SiO
2
and Ni
2
P/SiO
2
. When Mo was introduced in Ni
2
P/SiO
2
,
the increased selectivity to dodecanol is due to that Mo promoted
the HDO pathway. However, as the Mo content increased, the
selectivity to dodecanol decreased. This may be related to the
increased acid amount. When Ni was introduced in MoP/SiO
2
,
the selectivity to dodecanol tended to increase. This may be due
to the decrease in the acid amount.
Dodecanol can be converted to dodecene via dehydration, to
lauryl laurate via esterication with lauric acid, to n-dodecane
via hydrogenolysis of CO bond, or to C11 hydrocarbons via dehy-
drogenation followed by decarbonylation [8]. The dissociation of
CO bond is very difcult due to higher energy (about 360 kJ/
mol) [18]. Therefore, the dehydration as well as the esterication
is easier for the conversion of dodecanol, which was promoted
with the increase in the acid amount.
Lauryl laurate was derived from the esterication between
lauric acid and dodecanol on the acid site. Thus, its selectivity in-
creased as the catalyst acidity and the concentrations of lauric acid
and dodecanol increased. On the different catalysts, the variation
tendency for the selectivity to lauryl laurate was opposite to that
for the selectivity to dodecanol. On one hand, the formation of
lauryl laurate consumed dodecanol. On the other hand, the high
acidity favored both dehydration and esterication of dodecanol.
Here, we also compared the performances of different catalysts
at the third and the thirteenth hours (see Fig. 7S in supplementary
information). It was found that NiMo3/2 and NiMo3/3 gave better
stabilities than other catalysts. In other words, to enhance the
stability of mono-metallic (Ni or Mo) phosphide catalyst, the
bimetallic phosphide catalyst with suitable Ni/Mo ratio is an
option. Along with the reaction, the C11/C12 ratio only slightly
increased for Ni
2
P/SiO
2
, NiMo3/1, NiMo3/2 and NiMo3/3, whereas
it did not change for NiMo2/3, NiMo1/3 and MoP/SiO
2
obviously.
We speculate that the surface active sites did not change obviously
during 313 h. The revolution of surface active sites may mainly be
due to the product water. Water can interact with the surface
metal site and the P site, reducing the density of surface metal site
and creating new POH group [38]. However, this process was very
fast and mainly occurred at the initial reaction phase [38,39].
4. Conclusion
In the present work, we investigated the sulfur-free NiMo
bimetallic phosphide catalysts for the deoxygenation of methyl
laurate as a model reactant to produce hydrocarbons. The develop-
ment of sulfur-free catalyst avoids the formation of undesirable S-
containing products and reduces the investment cost. XRD,
HRTEM, CO chemisorption and NH
3
-TPD results indicate that the
dispersion and acidity of phosphide catalyst tended to decrease
with increasing Ni/Mo ratio. XPS gave the evidence that the elec-
tron transfer occurred not only from metal (Ni and Mo) to P but
also from Ni to Mo. The electronic interaction between Ni and
Mo is mainly derived from the incorporation of Mo into Ni
2
P as
well as the formation of MoNiP
2
phase. The deoxygenation of
methyl laurate was carried out at 300340 C, 3 MPa, H
2
/methyl
laurate of 50 and WHSV of 14 h
1
. Apart from NiMo3/3, the TOF
of methyl laurate and the C11/C12 ratio tended to increase as the
Ni/Mo ratio increased. This is related to the Ni/Mo ratio, the phos-
phide particle size and the electron density of metal site. The in-
creases in the Ni/Mo ratio and phosphide particle size favored
the C11/C12 ratio, whereas the electron transfer from Ni to Mo
gave rise to the decrease in C11/C12 ratio. Under the present con-
dition, the main oxygenated intermediate was dodecanol, followed
by lauryl laurate and lauric acid. The catalyst acidity obviously af-
fected the distributions of the oxygenated intermediates. The in-
crease in acidity favored the conversion of dodecanol to C12
hydrocarbon via dehydration and the formation of lauryl laurate.
NiMo3/3 showed special performance in comparison with
NiMo3/2 and NiMo2/3, which is probably related to its more
remarkable MoNiP
2
phase and larger phosphide particles. The rea-
son is still needed to investigate. As a whole, the deoxygenation
pathway (i.e., C11/C12 ratio) can be tuned by altering the Ni/Mo ra-
tio in the bimetallic phosphide. This can satisfy the practical de-
mand in hydrocarbon composition to produce different fuels
(such as jet fuel and diesel).
Acknowledgements
The authors gratefully acknowledge the supports from the
National Natural Science Foundation of China (No. 21176177),
the Natural Science Foundation of Tianjin (No. 12JCYBJC13200),
State Key Laboratory of Catalytic Materials and Reaction Engineering
(RIPP, SINOPEC) and the Program of Introducing Talents to the
University Disciplines (B06006).
Ni2P NiMo3/1 NiMo3/2 NiMo3/3 NiMo2/3 NiMo1/3 MoP
0
5
10
15
20
MoP/SiO
2
Ni
2
P/SiO
2
S
e
l
e
c
t
i
v
i
t
y

t
o

d
o
d
e
c
a
n
o
l
/
%
Catalyst
300
o
C
320
o
C
340
o
C
Ni2P NiMo3/1 NiMo3/2 NiMo3/3 NiMo2/3 NiMo1/3 MoP
0
2
4
6
8
MoP/SiO
2
Ni
2
P/SiO
2
S
e
l
e
c
t
i
v
i
t
y

t
o

l
a
u
r
y
l

l
a
u
r
a
t
e
/
%
Catalyst
300
o
C
320
o
C
340
o
C
Fig. 7. Selectivities to dodecanol and lauryl laurate.
J. Chen et al. / Fuel 129 (2014) 110 9
Appendix A. Supplementary material
Supplementary data associated with this article can be found, in
the online version, at http://dx.doi.org/10.1016/j.fuel.2014.03.049.
References
[1] da Silva VT, Sousa LA. Catalytic upgrading of fats and vegetable oils for the
production of fuels. In: Triantafyllidis K, Lappas A, Stcker M, editors. The role
of catalysis for the sustainable production of bio-fuels and bio-
chemicals. Amsterdam: Elsevier B.V; 2013. p. 6792.
[2] Kubickova I, Kubicka D. Utilization of triglycerides and related feedstocks for
production of clean hydrocarbon fuels and petrochemicals: a review. Waste
Biomass Valor 2010;1:293308.
[3] Han J, Duan J, Chen P, Lou H, Zheng X, Hong H. Carbon-supported molybdenum
carbide catalysts for the conversion of vegetable oils. ChemSusChem
2012;5:72733.
[4] Wang H, Yan S, Salley SO, Simon NKY. Support effects on hydrotreating of
soybean oil over NiMo carbide catalyst. Fuel 2013;111:817.
[5] Monnier J, Sulimma H, Dalai A, Caravaggio G. Hydrodeoxygenation of oleic acid
and canola oil over alumina-supported metal nitrides. Appl Catal A
2010;382:17680.
[6] Yang Y, Ochoa-Hernndez C, Pizarro P, de la Pea OShea VA, Coronado JM,
Serrano DP. Synthesis of nickel phosphide nanorods as catalyst for the
hydrotreating of methyl oleate. Top Catal 2012;55:9918.
[7] Yang Y, Ochoa-Hernndez C, de la Pea OShea VA, Coronado JM, Serrano DP.
Ni
2
P/SBA-15 as a hydrodeoxygenation catalyst with enhanced selectivity for
the conversion of methyl oleate into n-octadecane. ACS Catal 2012;2:5928.
[8] Chen J, Shi H, Li L, Li K. Deoxygenation of methyl laurate as a model compound
to hydrocarbons on transition metal phosphide catalysts. Appl Catal B
2014;144:87084.
[9] Yang Y, Chen J, Shi H. Deoxygenation of methyl laurate as a model compound
to hydrocarbons on Ni
2
P/SiO
2
, Ni
2
P/MCM-41, and Ni
2
P/SBA-15 catalysts with
different dispersions. Energy Fuels 2013;27:34009.
[10] Shi H, Chen J, Yang Y, Tian S. Catalytic deoxygenation of methyl laurate as a
model compound to hydrocarbons on nickel phosphide catalysts: remarkable
support effect. Fuel Process Technol 2014;118:16170.
[11] Li W, Dhandapani B, Oyama ST. Molybdenum phosphide: a novel catalyst for
hydrodenitrogenation. Chem Lett 1998;27:2078.
[12] Oyama ST, Gott T, Zhao H, Lee YK. Transition metal phosphide hydroprocessing
catalysts: a review. Catal Today 2009;143:94107.
[13] Prins R, Bussell ME. Metal phosphides: preparation, characterization and
catalytic reactivity. Catal Lett 2012;142:141336.
[14] Wu S, Lai P, Lin Y, Wan H, Lee H, Chang Y. Atmospheric hydrodeoxygenation of
guaiacol over alumina-, zirconia-, and silica-supported nickel phosphide
catalysts. ACS Sustainable Chem Eng 2013;1:34958.
[15] Cecilia JA, Infantes-Molina A, Rodrguez-Castelln E, Jimnez-Lpez A, Oyama
ST. Oxygen-removal of dibenzofuran as a model compound in biomass derived
bio-oil on nickel phosphide catalysts: role of phosphorus. Appl Catal B
2013;136137:1409.
[16] Whiffen VML, Smith KJ. A comparative study of 4-methylphenol
hydrodeoxygenation over high surface area MoP and Ni
2
P. Top Catal
2012;55:98190.
[17] Zhao HY, Li D, Bui P, Oyama ST. Hydrodeoxygenation of guaiacol as model
compound for pyrolysis oil on transition metal phosphide hydroprocessing
catalysts. Appl Catal A 2011;391:30510.
[18] Bui P, Cecilia JA, Oyama ST, Takagaki A, Infantes-Molina A, Zhao H, et al.
Studies of the synthesis of transition metal phosphides and their activity in
the hydrodeoxygenation of a biofuel model compound. J Catal 2012;294:
18498.
[19] Li K, Wang R, Chen J. Hydrodeoxygenation of anisole over silica-supported
Ni
2
P, MoP, and NiMoP catalysts. Energy Fuels 2011;25:85463.
[20] Dupont C, Lemeur R, Daudin A, Raybaud P. Hydrodeoxygenation pathways
catalyzed by MoS
2
and NiMoS active phases: a DFT study. J Catal 2011;279:
27686.
[21] Kubicka D, Kaluza L. Deoxygenation of vegetable oils over sulded Ni, Mo and
NiMo catalysts. Appl Catal A 2010;372:199208.
[22] Ruinart de Brimont M, Dupont C, Daudin A, Geantet C, Raybaud P.
Deoxygenation mechanisms on Ni-promoted MoS
2
bulk catalysts: a
combined experimental and theoretical study. J Catal 2012;286:15364.
[23] Sun F, Wu W, Wu Z, Guo J, Wei Z, Yang Y, et al. Dibenzothiophene
hydrodesulfurization activity and surface sites of silica-supported MoP, Ni
2
P,
and NiMoP catalysts. J Catal 2004;228:298310.
[24] Wang R, Smith KJ. Hydrodesulfurization of 4,6-dimethyldibenzothiophene
over high surface area metal phosphides. Appl Catal A 2009;361:1825.
[25] Abu II, Smith KJ. Hydrodenitrogenation of carbazole over a series of bulk
Ni
x
MoP catalysts. Catal Today 2007;125:24855.
[26] Zuzaniuk V, Prins R. Synthesis and characterization of silica-supported
transition-metal phosphides as HDN catalysts. J Catal 2003;219:8596.
[27] Sawhill SJ, Layman KA, Van Wyk DR, Engelhard MH, Wang C, Bussell ME.
Thiophene hydrodesulfurization over nickel phosphide catalysts: effect of the
precursor composition and support. J Catal 2005;231:30013.
[28] Katrib A, Logie V, Saurel N, Wehrer P, Hilaire L, Maire G. Surface electronic
structure and isomerization reactions of alkanes on some transition metal
oxides. Surf Sci 1997;377379:7548.
[29] Layman KA, Bussell ME. Infrared spectroscopic investigation of CO adsorption
on silica-supported nickel phosphide catalysts. J Phys Chem B
2004;108:1093041.
[30] Phillips DC, Sawhill SJ, Self R, Bussell ME. Synthesis, characterization, and
hydrodesulfurization properties of silica-supported molybdenum phosphide
catalysts. J Catal 2002;207:26673.
[31] Rodriguez JA, Kim JY, Hanson JC, Sawhill SJ, Bussell ME. Physical and chemical
properties of MoP, Ni
2
P, and MoNiP hydrodesulfurization catalysts: time-
resolved X-ray diffraction, density functional, and hydrodesulfurization
activity studies. J Phys Chem B 2003;107:627685.
[32] Lee YK, Oyama ST. Bifunctional nature of a SiO
2
-supported Ni
2
P catalyst for
hydrotreating: EXAFS and FTIR studies. J Catal 2006;239:37689.
[33] Whiffen VML, Smith KJ. Hydrodeoxygenation of 4-methylphenol over
unsupported MoP, MoS
2
, and MoO
x
catalysts. Energy Fuels 2010;24:
472837.
[34] Chen L, Zhu Y, Zheng H, Zhang C, Li Y. Aqueous-phase hydrodeoxygenation of
propanoic acid over the Ru/ZrO
2
and RuMo/ZrO
2
catalysts. Appl Catal A
2012;411412:95104.
[35] David RL. CRC hand book of chemistry and physics. CD-ROM version
2010. Boca Raton, FL: CRC Press/Taylor and Francis; 2010.
[36] Boudart M. Turnover rates in heterogeneous catalysis. Chem Rev
1995;95:6616.
[37] Lestari S, Mki-Arvela P, Beltramini J, Max Lu GQ, Murzin DY. Transforming
triglycerides and fatty acids into biofuels. ChemSusChem 2009;2:110919.
[38] Guo T, Chen J, Li K. Promotion effect of steam treatment on activity of Ni
2
P/
SiO
2
for hydrodechlorination of chlorobenzene. Chin J Catal 2012;33:10805.
[39] Liu P, Rodriguez JA, Takahashi Y, Nakamura K. Watergas-shift reaction on a
Ni
2
P (001) catalyst: formation of oxy-phosphides and highly active reaction
sites. J Catal 2009;262:294303.
10 J. Chen et al. / Fuel 129 (2014) 110

Você também pode gostar