Sie sind auf Seite 1von 10

Applied Catalysis A: General 435436 (2012) 5160

Contents lists available at SciVerse ScienceDirect


Applied Catalysis A: General
j our nal homepage: www. el sevi er . com/ l ocat e/ apcat a
Comparison of alumina- and SBA-15-supported molybdenum nitride catalysts
for hydrodeoxygenation of guaiacol
I. Tyrone Ghampson
a,b
, Catherine Seplveda
c
, Rafael Garcia
c
, J.L. Garca Fierro
d
,
Nestor Escalona
c,
, William J. DeSisto
a,e,
a
Department of Chemical and Biological Engineering, University of Maine, Orono, ME 04469, United States
b
Unidad de Desarrollo Tecnolgico, Universidad de Concepcin, Casilla 4051, Concepcin, Chile
c
Universidad de Concepcin, Facultad de Ciencias Quimicas, Casilla 160c, Concepcin, Chile
d
Instituto de Catalisis y Petroquimica, CSIC, Cantoblanco, 28049 Madrid, Spain
e
Forest Bioproducts Research Institute, University of Maine, Orono, ME 04469, United States
a r t i c l e i n f o
Article history:
Received 10 April 2012
Received in revised form23 May 2012
Accepted 25 May 2012
Available online 4 June 2012
Keywords:
Hydrodeoxygenation
Alumina
SBA-15
Nitride
Oxynitride
Catalyst
a b s t r a c t
The hydrodeoxygenation of guaiacol (2-methoxyphenol) has been studied in a batch reactor over
alumina- and SBA-15 silica-supported molybdenum nitride catalysts at 300

C and 5MPa of hydrogen


pressure. The catalysts were prepared by nitriding supported Mo oxide precursors with ammonia gas
or nitrogenhydrogen mixtures via temperature-programmed reaction. The alumina-supported cata-
lysts had a higher activity relative to the SBA-15 silica-supported catalysts which was essentially due
to catechol production, an effect of the alumina support. The SBA-15 silica-supported catalysts trans-
formed guaiacol directly to phenol by demethoxylation without noticeable catechol production. On both
supports, nitridation by ammonolysis increased the activity by a factor of 1.1 relative to nitridation
by nitrogenhydrogen. On SBA-15, ammonolysis preferentially produced the -Mo
2
N phase while the
N
2
/H
2
mixture produced the -Mo
2
N
0.78
phase. The incorporation of Co led to a marginal improvement
in exposed Mo species but generally had a diminishing effect on HDO activity. The lack of catechol pro-
duction using the SBA-15 silica support is important in minimizing coking reactions and also opens up
possibilities for utilizing silica supports with highly controlled pore sizes to possibly inuence product
distribution in HDO of more diverse feed streams derived from biomass conversion processes.
2012 Elsevier B.V. All rights reserved.
1. Introduction
The removal of oxygen from biomass-derived oils through
catalytic hydrodeoxygenation (HDO) is receiving considerable
attention because of its potential as feedstock for the production
of fuels and value-added chemicals. Most of the initial studies on
HDO reactions have been conducted over metal suldes supported
on alumina [1,2]. However, recently there have been signicant
efforts on the development of catalysts, based on newor modied
supports and newactive phases, withminimal hydrogenconsump-
tionandhighselectivitytowarddirect oxygenremoval [1]. Guaiacol
(2-methoxyphenol) has commonly been used as a model com-
pound for the HDOstudies because it is known to exist signicantly

Corresponding author. Tel.: +56 41 2207236; fax: +56 41 2245374.

Corresponding author at: Department of Chemical and Biological Engineering,


University of Maine, Orono, ME 04469, United States. Tel.: +1 207 581 2291;
fax: +1 207 581 2323.
E-mail addresses: nescalona@udec.cl (N. Escalona),
WDeSisto@umche.maine.edu (W.J. DeSisto).
in bio-oils, because of its propensity for coke formation, and also
because of its intransigence to deoxygenation [2]. Several of the
studies have reported that the HDO activity and selective trans-
formation of guaiacol to phenol is distinctively inuenced by the
nature of the support [3,4]. Catalysts supported on alumina (Al
2
O
3
)
displayed higher activity compared with alternative supports such
as silica and carbon owing to higher dispersion of the active phase
[3]. However, alumina-supported catalysts suffer from coke for-
mation which limits the lifetime of the catalyst [3]. The benets
of using silica and carbon supports lie in the minimal coke forma-
tionandgreater selectivitytowardphenol productionover catechol
[3].
Interests in exploring non-sulded catalysts for HDO have
increased due to issues related to contamination of the feed by
the sulding agent [5,6]. Novel active phases such as metal nitrides
have beenshownto be aneffective catalyst for HDS [7] and HDN[8]
reactions. However, only a limitednumber of studies have reported
their performance for HDO reactions. Recently, Monnier et al. [9]
reported activity on -Al
2
O
3
-supported Mo, W, and V nitride cat-
alysts for HDO of oleic acid and canola oil. In a recent study, we
demonstrated high activity and rapid demethoxylation of guaiacol
0926-860X/$ see front matter 2012 Elsevier B.V. All rights reserved.
http://dx.doi.org/10.1016/j.apcata.2012.05.039
52 I. Tyrone Ghampson et al. / Applied Catalysis A: General 435436 (2012) 5160
to phenol using unsupported Mo nitride catalysts [10]. It is the pur-
pose of this paper to extend our investigations from bulk nitrides
to alumina- and silica-supported nitrides for HDO of guaiacol.
Mesoporous silica materials have generatedinterests as catalyst
supports in heterogeneous catalysis owing to their very high spe-
cic surface areas, controllable pore diameters, narrow pore size
distributions and large pore volumes. These properties have made
them a more desirable alternative to conventional silica supports,
enablingthecontrol of catalyst particlesizeandinuencingproduct
selectivity through transport effects [11,12]. In particular, SBA-15
is of general interest because of its high structure regularity, low-
cost and non-toxicity [13]. Studies to extend our understanding of
the SBA-15 pore structure and its relationship to transport proper-
ties [14], and the hydrothermal stability of SBA-15 [15] have been
examined to generate fundamental information relevant to the use
of this material as a support in developments of catalysts for bio-
fuel production. We focus here on demonstrating the reactivity of
SBA-15 silica supported molybdenumnitride catalysts for the HDO
of a biomass-derived compound.
Here, we report on the synthesis and characterization of
molybdenum nitride dispersed on SBA-15 mesoporous silica
and -alumina. The molybdenum nitride was synthesized by
impregnation, oxidation and thermal conversion to the nitride.
Thermal conversion was achieved by two different procedures:
ammonolysis and reduction/nitridation using hydrogen/nitrogen
mixtures. The resulting materials were characterized using nitro-
genadsorptiondesorption(surface area, porosity), XRD, elemental
analysis andXPS, andtheir activity was comparedwithcommercial
sulded NiMo/Al
2
O
3
catalysts for the HDO of guaiacol.
2. Experimental
2.1. Catalyst preparation
SBA-15 mesoporous silica was synthesized following a reported
procedure [16]. In a typical synthesis, 6g of Pluronic P123 ethylene
oxide/propylene oxide block copolymer block copolymer (BASF,
M
avg
=5800, EO
20
PO
70
EO
20
) was dissolved with stirring in 45g
of deionized water and 180g of 2mol L
1
HCl. Then, 12.75g of
tetraethyl orthosilicate (Aldrich, TEOS, 99%) was added to the
solution with stirring at 40

C for 24h. The homogeneous solgel


mixture was then transferred into a polymer ask, sealed and
heated at 100

C for 48h. After ltration and washing with water,


the white solid product recovered was air-dried at room temper-
ature for 24h. The sample was then calcined in air with a heating
rate of 1

Cmin
1
to 500

C and held for 10h. The alumina support


was obtained commercially (Alfa Aesar, -Al
2
O
3
, 1/8

pellets) and
used without any further purication.
The supported molybdenumoxide precursors were prepared by
incipient wetness impregnation using aqueous solutions of ammo-
nium heptamolybdate (Fischer Scientic, (NH
4
)
6
Mo
7
O
24
4H
2
O,
A.C.S. grade). The mixture was then left overnight at ambient tem-
perature. After impregnation, the samples were dried for 12h
at 110

C and 120

C for SBA-15- and alumina-supported sam-


ples respectively, and subsequently calcined in a ow of dry air
with a heating rate of 1

Cmin
1
at 500

C and held for 3h. The


bimetallic oxide precursors were prepared by impregnating the
Mo oxide supported materials with an aqueous solution of cobalt
(II) nitrate hexahydrate (Acros Organics, Co(NO
3
)
2
6H
2
O, 99%), fol-
lowed by the same drying-calcination procedure described above.
The supported oxides were prepared to obtain nominal loading
of 10wt% Mo metal content for monometallic samples, 10wt%
Mo metal and 2.4wt% Co metal content for the bimetallic sam-
ples. All oxide precursors were sieved to obtain a 180450m
particle size. Supported molybdenum nitrides were prepared by
temperature-programmed reaction of the oxidic precursors with
NH
3
(Matheson, 99.99%), or N
2
/H
2
mixtures (N
2
, BOC Gases, Grade
5; H
2
, Matheson, 99.99%). The synthesis procedure involved ow-
ing 300mL min
1
of either NH
3
or N
2
/H
2
(N
2
/H
2
=5/1 (v/v)) over
the 2.5g of the supported Mo oxide precursor while the temper-
ature was increased from room temperature to 300

C in 30min,
then from 300

C to 500

C within 5.6h, and from 500

C to 700

C
within 1.7h. The temperature was maintained at 700

C for 2h, and


then cooled to room temperature under NH
3
ow for the nitrides
prepared via ammonolysis or cooled under nitrogen ow for the
nitrides prepared using the N
2
/H
2
mixture. Finally, the sample was
passivated in 1% O
2
/N
2
for 12h at room temperature. For nota-
tion, nitrides prepared using ammonia (method 1) have sufx A,
while nitrides preparedusingnitrogenhydrogenmixture (method
2) have sufx NH.
Molybdenum, cobalt andnitrogencontents inthe catalysts were
performed by the Analytical Laboratory of the Department of Plant,
Soil and Environmental Sciences at the University of Maine, and by
GalbraithLaboratory. ICP-AES was used for the metal content while
a combustion method was used for nitrogen.
2.2. Catalyst characterization
Nitrogen sorption isotherms were obtained at 77K using a
Micromeritics ASAP-2020 instrument to evaluate the BET spe-
cic surface area (S
BET
), total pore volume (TPV) and average
pore diameter (d
pore
). Prior to the measurements, the samples
were outgassed under vacuum following conditions common for
these materials: SBA-15-supported materials were outgassed at
200

C for 8h, and alumina-supported materials were outgassed


at 250

C for 2h. S
BET
was calculated using the adsorption branch
in the range of 0.05P/P
0
0.25 and the TPV was recorded at
P/P
0
=0.995. The primary pore diameter was estimated from the
maximum in the BJH pore size distribution. The micropore vol-
ume was estimatedfromthe
s
-plot methodusing-Al
2
O
3
(Fischer
Scientic, powder certied, S
BET
=0.9m
2
g
1
,
ref,0.4
=0.28cm
3
g
1
STP) for alumina-supported nitrides and LiChrospher Si-1000 silica
gel (S
BET
=26.4m
2
g
1
,
ref,0.4
=9.12cm
3
g
1
STP) [17] for SBA-15-
supported nitrides as reference adsorbents.
X-ray diffraction (XRD) patterns of powdered samples were
recorded on a PANalytical XPert PRO X-ray diffractometer
equipped with a graphite monochrometer and Cu K radiation
(45kV, 40mA) in a parallel beam optical geometry. The standard
scan parameters were 1585

2 with a step size of 0.02

and a
counting time of 10s per step. Identication of the phases was
achieved by reference to JCPDS diffraction le data.
X-ray photoelectron spectra of reduced catalysts were obtained
on a VG Escalab 200R electron spectrometer using a Mg K
(1253.6eV) photon source. The passivated catalysts were activated
ex situ with H
2
at 450

C for 6h. After reduction, the samples were


cooled to room temperature, ushed with nitrogen and stored in
asks containing isooctane (Merck, 99.8%), then transferred to the
pre-treatment chamber of the spectrometer. The binding energies
(BE) were referenced to the C 1s level of the carbon support at
284.9eV. Intensities of the peaks were calculated fromthe respec-
tive peak areas after background subtraction and spectrum tting
byacombinationof Gaussian/Lorentzianfunctions. Relativesurface
atomic ratios were determined fromthe corresponding peak areas,
corrected with tabulated sensitivity factors [18], with a precision
of 7%.
The acid strength and acid site concentration of some selected
catalysts were measured using a potentiometric method [19],
whereby a suspension of the material in acetonitrile was titrated
with n-butylamine. The variation in electric potential was regis-
tered on a Denver Instrument UltraBasic pH/mV meter.
I. Tyrone Ghampson et al. / Applied Catalysis A: General 435436 (2012) 5160 53
2.3. Catalytic activity measurements
Guaiacol HDO activity measurements were carried out in a
300mL stainless steel batch autoclave (Parr Model 4841) at 300

C
and under a hydrogen pressure of 5MPa. Prior to catalytic testing,
the passivated samples were activated ex situ under H
2
at 450

C
for 6h. As a basis for comparison, commercial Ni-Mo/Al
2
O
3
cata-
lyst (Procatalyse HR 346) was pre-sulded using a 10vol% H
2
S in
H
2
mixture at 350

C for 3h. Approximately 0.25g of freshly pre-


treated catalyst (nitride catalysts or the commercial catalyst) was
addedtothe liquidfeedcontaining 2.53mL guaiacol (0.232mol L
1
,
Merck, 99%), 80mL decalin (Merck, 99.5%) as solvent, and 700L
of hexadecane (Merck, 99%). Hexadecane was used as an internal
standardfor quantitativeGCanalysis. Thesealedreactor was evacu-
ated with nitrogen (AGAChile, Grade 5) for 30min with continuous
stirring; after a leak check the reactor was heated to 300

C under
N
2
which was subsequently replaced with H
2
and then pressurized
to 5MPa. This pressure was maintained for the entire duration of
the experiment by adding H
2
to the reactor whenever necessary.
Liquid samples were periodically withdrawn during the course of
the reaction after purging the sampling line with a small amount of
the reactant mixture. The samples were analyzed by a Perkin Elmer
(Clarus 400) gas chromatograph equipped with a ame ionization
detector (FID) and a CP-Sil 5 CB column (Agilent). The product
distributions were identied by their column retention time in
comparison with available standards. The initial concentration of
guaiacol was taken as 100% in order to ignore slight conversion
before isothermal condition was achieved. The catalytic activity
was expressed by the initial reaction rate.
3. Results and discussion
3.1. Catalyst properties
Fig. 1 shows the X-ray diffraction proles of the supports and
catalysts used in this study. Examination of the XRD patterns
of Mo and CoMo nitrides supported on alumina revealed only
diffraction peaks associated with the supports. Broad diffraction
peaks characteristic of -Al
2
O
3
(JSPDS ref no: 010-0425) may have
masked the nitride peaks. The XRD patterns of MoN/SBA-15-A and
MoN/SBA-15-NH catalysts showed broad peaks for crystalline -
Mo
2
N (2 =37.13, 43.41, and 63.03) and -Mo
2
N
0.78
(2 =37.51,
43.11, 62.89, and 75.45) respectively. The estimated average diam-
eter of the Mo nitride crystals from the diffraction peaks of the
MoN/SBA-15-A and MoN/SBA-15-NH catalysts were 2 and 3nm
respectively. The XRD pattern collected for the CoMoN/SBA-15-
A catalyst revealed the formation of Mo
2
N but no evidence of
Co
3
Mo
3
N phase. The presence of Mo
2
N as the only phase observed
in this catalyst may suggest that the particle size of the Co
3
Mo
3
N
phasewas small andbelowthedetectionlimit. It is evident fromthe
XRD results that the crystal structure of supported Mo nitride par-
ticles are closely related to their nitridation and purging treatment
during the formationof the particles: catalysts preparedby method
1 resulted in -Mo
2
N particles and those prepared by method 2
resulted in -Mo
2
N
0.78
.
Elemental analyses of passivated, supported Mo and CoMo
nitrides are listed in Table 1. The nitrogen contents for nitride sam-
ples from method 1 were higher compared to samples prepared
from method 2 in both supports. The lower nitrogen content for
the method 2 samples may be due to purging the samples in ow-
ing nitrogen after nitridation which removed weakly bonded NH
x
species [20]. The atomic N/Mo ratios of the nitride catalysts are
presented in Table 1. While the theoretical N/Mo is 0.5 and 0.39 for
-Mo
2
N and -Mo
2
N
0.78
respectively, it is plausible that excess N
could reside in interstitial sites and defects like grain boundaries.
Fig. 1. XRD patterns of (a) alumina-supported samples and (b) SBA-15-supported
samples.
These ndings, in addition to the absence of diffraction peaks for
crystalline Mo nitride phase on the alumina-supported catalyst,
suggest that small crystallites of Mo nitrides below the detection
limit were formed on the catalysts. This is consistent with other
report in the literature for Mo
2
N/Al
2
O
3
catalysts [21].
Nitrogen sorption analyses were performed to determine the
difference in support morphology between -Al
2
O
3
and SBA-15
materials. The N
2
adsorptiondesorption isotherms and pore size
distribution (PSD) of alumina and SBA-15 materials are shown in
Fig. 2. The isotherm and PSD for the as-prepared SBA-15 support
were consistent with previously reported results for this material
[16,22]. AccordingtoIUPACclassications, theN
2
isotherms (Fig. 2a
and b) belong to a type IV isothermwhich is typical of mesoporous
materials [23]. The isotherms for the SBA-15 materials in Fig. 2b
Table 1
Chemical composition of passivated nitride catalysts.
Catalyst Mo (wt%) Co (wt%) N (wt%) N/Mo
atomic
ratio
MoN/Al
2
O
3
-A 8.39 1.76 1.44
MoN/Al
2
O
3
-NH 8.74 0.49 0.38
CoMoN/Al
2
O
3
-A 8.33 1.87 1.66 1.54
MoN/SBA15-A 7.22 2.26 2.15
MoN/SBA15-NH 6.67 0.92 0.95
CoMoN/SBA15-A 7.34 1.95 1.42 1.32
54 I. Tyrone Ghampson et al. / Applied Catalysis A: General 435436 (2012) 5160
Fig. 2. Morphology of samples under study: N
2
adsorptiondesorption isotherms of (a) alumina-supported samples and (b) SBA-15-supported samples; BJH pore size
distributions of (c) alumina-supported samples and (d) SBA-15-supported samples.
show a sharp inection in the relative pressure (P/P
0
) range from
0.7 to 0.8, indicative of the presence of uniform pore size distri-
bution [24]. The isotherms of -Al
2
O
3
-based materials revealed a
sharp inection in the P/P
0
range from0.8 to 1.0, suggestive of wide
pore size distribution. Fig. 2a and b also shows that the quantity
of nitrogen adsorbed and the P/P
0
position of the inection point
on the isotherm of the support decreased after impregnation and
nitridation.
The BJH pore size distributions (PSD), as derived from the des-
orption branch of the N
2
isotherm of the materials under study,
are shown in Fig. 2c and d. As seen in Fig. 2c and d, the SBA-15
mesoporous silica support yielded a narrow pore size distribu-
tion centered at 8.3nm, while the alumina support yielded a
broad pore size distribution centered at 9.3nm. In addition, PSD
curves of metal nitrides supported on SBA-15 mesoporous silica
revealed a bimodal pore distribution system with peaks at 4.5nm
and 7.2nm, indicating the presence of small complementary pores
and ordered mesoporous pores, respectively, as observed by oth-
ers [12,25]. The bimodal PSD and the broadening of the higher
mesopore peak suggest that the metal nitride species were not
uniformly distributed in the mesopores of the SBA-15 support. A

s
-plot analysis revealed the presence of micropores for SBA-15
support (2.4% microporosity) and more than a 73% decrease in
microporosity after impregnation and thermal treatment: this sug-
gests microporeblockingbythemetal nitridespecies, leadingtothe
near-eliminationof micropores andthe appearance of complemen-
tary pores caused by inter-particle porosity. These complementary
pores of size <4nm have been discussed to provide connectivity
between the primary mesopores of SBA-15 [26,27]. A closer look at
the PSD of the SBA-15-supported catalysts suggests greater com-
plementary pores for the MoN/SBA-15-A catalyst as compared to
the MoN/SBA-15-NH catalyst despite the marginally higher Mo
content of the former. The XRD data of the two catalysts suggests
that Mo nitride crystallites formed in the MoN/SBA-15-NHcatalyst
were larger than those formed in the MoN/SBA-15-Acatalyst. Thus,
it is possiblethat access for N
2
adsorption/desorptioncouldbemore
limited by the relatively larger crystallites of the MoN/SBA-15-NH
catalyst, leading to lesser generation of the complementary pores.
This is in good agreement with the difference in the N
2
physisorp-
tion data presented in Table 2.
In contrast to the pore size distribution results of the SBA-
15-supported metal nitrides, the pore size distributions of the
metal nitride/Al
2
O
3
catalysts were unimodal (which is similar to
the supports PSD) and decreased slightly to smaller mesopores
in comparison to the PSD of the Al
2
O
3
support. Furthermore, the
micropore volumes of the catalysts estimated from the
s
-plot
I. Tyrone Ghampson et al. / Applied Catalysis A: General 435436 (2012) 5160 55
Table 2
Adsorption properties of supports and passivated supported nitride catalysts.
Sample SBET (m
2
g
1
) TPV (cm
3
g
1
) V (cm
3
g
1
) dpore (nm)
-Al
2
O
3
207 0.62 0.03 9.3
MoN/Al
2
O
3
-A 191 0.51 0.02 8.6
MoN/Al
2
O
3
-NH 183 0.50 0.02 8.9
CoMoN/Al
2
O
3
-A 182 0.49 0.02 8.9
SBA-15 818 1.25 0.03 8.3
MoN/SBA15-A 418 0.75 0.008 7.3
MoN/SBA15-NH 397 0.70 0.006 7.2
CoMoN/SBA15-A 387 0.62 0.007 7.5
method showed a 33% decrease in microporosity. These results
indicate a coating of the inner surfaces of the mesopores by the
metal nitride species, causing the pore entrance of a fraction of
micropores to close.
Table 3lists a summary of XPS results of the surface composition
and oxidation states of the components in the reduced, passivated
Mo nitride catalysts. The binding energies of the Mo 3d
5/2
, N 1s, Co
2p
3/2
, Si 2p and Al 2p core levels and the surface atomic ratios are
presented in Table 3. The Mo 3d and N1s XPS spectra of the nitride
catalysts are illustrated in Fig. 3. The Mo 3d
5/2
spectra presented
three peaks centered at 229.0, 230.5 and 232.6eV: peak position of
229.00.2eVis typical of Mo
+
(2< <4) assigned to Mo
2
Nspecies
[28]; binding energies of 230.50.1eV and 232.60.3eV are typ-
ical of Mo
4+
and Mo
6+
respectively in molybdenum oxynitrides
[29]. This result indicates that Mo
2
N and Mo oxynitrides coex-
ist on the surface of all the catalysts. The surface distribution of
the Mo oxidation states in the Mo 3d (compiled in parentheses in
Table 3) shows that molybdenumoxynitride was the dominant sur-
face phase, while Mo
2
N particles were distributed from 9 to 30%.
This result indicates that despite reductionof the passivatednitride
catalysts at 450

C, the sample surfaces were mainly oxynitrides


rather thannitrides. Table3alsoshows that higher amount of Mo
2
N
(Mo
+
) were formedonthe surface of the SBA-15silica support than
on the -Al
2
O
3
support, indicating that more nitrogen-decient
patches of Mo were on the surface of the former. This behavior is
attributed to weaker metal-support interactions of the SBA-15 sil-
ica support whichleads to relatively easier reducibility of the MoO
3
precursor. The XPS results also show that the Mo
+
content on the
method 2 catalyst surfaces was higher than the method 1 catalyst.
This result is however unclear. The Co 2p
3/2
binding energy for the
supported-CoMo nitrided samples of 778.40.1eV is within the
range of the reported binding energy of Co
0
for cobalt-nitride
species [30], while the binding energy of 781.50.2eV compares
well with Co
3+
cation of CoMo oxynitrides [30]. The amount of
Co
3+
species in the bimetallic nitrides was higher than the amount
of zero-valent cobalt. Thus, the most abundant surface metal
species in the bimetallic nitrides were the oxynitrides. The N 1s
core-level spectra (shown in Fig. 3) made three contributions: the
binding energy=394.40.3eVcould be attributed to the Me O N
bond (Me:Mo, Co, etc.) [31]; the binding energy=396.70.3eV is
close to Mo N bond [30]; the binding energy near 398.50.3eV
has been reported to be assigned to nitrogen atoms trapped in the
grain boundary of Mo nitrides [32]. Finally, the binding energies
of 103.4eV obtained for SBA-15-supported catalysts and 74.5eV
obtained for alumina-supported catalysts were identical to the Si
2p and Al 2p of the parent SiO
2
[33] and Al
2
O
3
respectively [34].
This suggests that the alumina and SBA-15 silica supports were not
nitrided under the synthesis condition.
The XPS Mo 3d/Al 2p (or Si 2p), N 1s/Al 2p (or Si 2p), and Co/Al
2p (or Si 2p) atomic ratios for the reduced, passivated catalysts are
shown in Table 3. The Mo/Al surface atomic ratio was greater than
the Mo/Si atomic ratio in all the catalysts. This indicates that SBA-
15-supported catalysts possess their molybdenumspecies located
inside the inner silica porous structure while the Al
2
O
3
-supported
catalysts had their molybdenum species located on the external
support surface of the catalyst. This is in good agreement with
their N
2
adsorptiondesorption isotherms of the passivated cat-
alysts (shown in Fig. 1) which shows a more pronounced decrease
in the quantity of N
2
adsorbed and a shift of the hysteresis loop to
lower P/P
0
values of the SBA-15-supported catalysts, indicating the
Table 3
XPS binding energies (eV) and surface atomic ratios of reduced, passivated nitride catalysts.
Catalyst Mo3d
5/2
N 1s Co 2p
3/2
Si 2p or Al 2p Mo/Si(Al) N/Si(Al) Co/Si(Al)
MoN/Al
2
O
3
-A
228.9 (9) 394.7 (17)

74.5 0.099 0.236


230.5 (22) 396.7 (36)
232.7 (69) 398.7 (47)
MoN/Al
2
O
3
-NH
229.0 (15) 394.5 (19)

74.5 0.133 0.349


230.5 (25) 396.3 (34)
232.6 (60) 398.4 (47)
CoMoN/Al
2
O
3
-A
228.9 (19) 394.5 (29)
230.5 (30) 396.5 (35) 778.5(18) 74.5 0.108 0.386 0.038
232.6 (51) 398.3 (36) 781.5(82)
MoN/SBA15-A
229.0 (25) 394.4 (26)

103.4 0.033 0.084


230.5 (29) 396.5 (35)
232.9 (46) 398.7 (39)
MoN/SBA15-NH
228.9 (27) 394.4 (28)

103.4 0.035 0.086


230.4 (31) 396.4 (38)
232.6 (42) 398.5 (34)
CoMoN/SBA15-A
228.9 (30) 394.4 (33)
230.5 (28) 396.5 (38) 778.4(19) 103.4 0.034 0.108 0.013
232.6 (42) 398.6 (29) 781.7(81)
56 I. Tyrone Ghampson et al. / Applied Catalysis A: General 435436 (2012) 5160
Fig. 3. XPS (a) Mo 3d
5/2
and (b) N 1s spectra for the reduced, passivated nitride catalysts.
presenceof particles intheporous structure. Table3alsoshows that
catalysts prepared by reduction/nitridation using a N
2
/H
2
mixture
produced more exposed Mo and N species on the surface of the
support than catalysts prepared by ammonolysis. However, this
behavior is inconsistent with the estimates of dispersion based on
particle sizes determined by XRD of SBA-15-silica supported cata-
lysts. These results suggest that the nitride species depositedonthe
surface were different fromthose deposited inside the pores of the
support. In other words, the SBA-15 silica support promoted the
formation of nitrogen-decient patches of Mo on the nitride sur-
face, whiletheformationof nitrides withhigher amount of nitrogen
was preferentially located inside the silica pore.
3.2. Reactivity
3.2.1. Activity of supported metal nitride catalysts: Al
2
O
3
vs.
SBA-15
In the present study, we evaluated and compared the catalytic
properties of Mo nitrides supported on SBA-15 mesoporous silica
and Mo nitrides supported on conventional alumina. Fig. 4 shows
the evolution of reactants and products during the HDO of guaiacol
on supported Mo nitride catalysts: the product yields are reported
in g/100g of guaiacol in the feed. Guaiacol transformation followed
the reaction scheme (Fig. 5) proposed by Bui et al. [35]. The scheme
mainly involves two stages: the rst stage involves methoxy group
removal onguaiacol to formphenol by either demethylation(DME)
and dehydroxylation, or by direct demethoxylation (DMO). The
second stage involves parallel pathways for phenol deoxygenation
either through direct hydrogenolysis (DDO) to produce benzene or
through hydrogenation (HYD) of the benzene ring prior to oxy-
gen removal to form cyclohexene. Methylation of the aromatic
ring forms heavy compounds such as di- and tetra-methyl phenols
and dimethyl catechols [35]. Fig. 3 shows that these heavy com-
pounds were observedinsignicant quantity withbothAl
2
O
3

and
SBA-15-supported catalysts. Fig. 4 also indicates that there were
clear differences between alumina- and SBA-15-supported cata-
lysts in terms of the changes in products concentration with time.
Over the alumina-supportedcatalysts, catechol was the major early
I. Tyrone Ghampson et al. / Applied Catalysis A: General 435436 (2012) 5160 57
Fig. 4. Variation of the transformation of guaiacol and the yield of products with time for (a) MoN/Al
2
O
3
-A, (b) MoN/Al
2
O
3
-NH, (c) CoMoN/Al
2
O
3
-A, (d) MoN/SBA-15-A, (e)
MoN/SBA-15-NH, and (f) CoMoN/SBA-15-A catalysts.
products while phenol surpassed catechol at longer reaction time.
The SBA-15-supported catalysts produced more phenol than cat-
echol at both lower and higher conversions. These results are
consistent with previously published work by Centeno et al. [3]
on metal sulde catalysts. On the basis of the products evolu-
tion in Fig. 3, alumina-supported catalysts proceeded through
both the DME and DMO pathways, while the mesoporous silica-
supported catalysts proceeded mainly through DMO pathway.
Under batch reaction conditions the formation of methane and
methanol, byproducts of DMO and DME respectively, could
not be separated by the column used and hence were not
observed. Methylcatechol was observed only with the alumina-
supported catalyst, consistent with other reported studies [4,36].
This product was formed probably through methyl-substitution
of catechol [4] or through transalkylation of guaiacol [36]. It
has recently been reported that catechol formation and its sub-
sequent methylation can be attributed to the acidic properties
of the support [4]: an -Al
2
O
3
support possessing more acid
sites than TiO
2
and ZrO
2
supports were found to favor DME
and methylation routes. The acid strength of MoN/Al
2
O
3
-A and
MoN/SBA-15-A catalysts, estimated from potentiometric titration
of the catalyst in acetonitrile with n-butylamine [19], were iden-
tical with the initial electrode potential of 120mV and 115mV,
respectively; however, the MoN/Al
2
O
3
-A catalyst possessed four
times a higher density of acid sites (8.2meq m
2
) than the
MoN/SBA-15-A catalyst (2.1meq m
2
). This suggests that the
propensity for DME and methylation routes of the alumina-
supported catalysts were inuenced by their higher total acidity.
Another possible reason could be due to the nature of sites
at the surface of the respective catalysts. The presence of
Lewis acid sites may be responsible for alumina promoting
catechol formation. Furthermore, the active sites on alumina
are different to those on SBA-15, as suggested by XRD and
XPS. Trace amounts of deoxygenated products such as benzene,
58 I. Tyrone Ghampson et al. / Applied Catalysis A: General 435436 (2012) 5160
Fig. 5. Guaiacol transformation pathways proposed by Bui et al. [35].
cyclohexene and cyclohexane (collectively denoted HDO products
in Fig. 4) were also observed. Other possible products from gua-
iacol conversion including anisole, o- and m-cresol, veratrole, etc.
[36] were not detected.
The activities of the catalysts expressed as the reaction rates, as
well as the calculated intrinsic activities are presented in Table 4.
In addition, the total guaiacol conversion after 4h of reaction is
reported in Table 4. The reaction rates were calculated from the
initial reaction during the transformation of guaiacol shown in
Fig. 4. Blank reactions with the Al
2
O
3
and SBA-15 silica supports
alone showed an appreciable conversion of guaiacol to catechol
by the former while no signicant conversion was observed for
the latter. The reaction rates of the Al
2
O
3
-supported catalysts were
more than two times higher than the SBA-15-supported catalysts.
The differences in catalytic activity between Mo nitride supported
on alumina and SBA-15 could be related to the density of acid
sites of the supports. Thus, the higher reaction rates displayed
by the alumina-supported nitride catalysts could be explained by
their faster conversionof guaiacol to catechol, methylcatechols and
heavy compounds as a result of their higher total acidity [5]. The
Lewis acidic nature of the alumina support have previously been
found to be prone to substantial coke formation through strong
interaction with guaiacol, forming doubly anchored phenates [37].
The intrinsic activities, calculated using experimental reaction rate
data normalized by the molybdenum content of the catalyst, are
presented in Table 4. The trend in the initial intrinsic activities
was similar to that of the initial reaction rates since the catalysts
contained similar Mo contents.
Although the Mo nitride catalysts supported on SBA-15
mesoporous silica showed lower activities, they have particu-
lar advantages over the alumina-supported nitride and sulde
catalysts in terms of the higher specicity for phenol over cate-
chol. This has implications with regards to lower consumption of
hydrogen and less formation of coke in HDO applications. In addi-
tion, the ability to nely tune pore sizes in ordered mesoporous
silica supports offers other opportunities to inuence activity and
selectivity by controlling catalyst dispersion and the diffusion of
reactants and products fromthe active catalyst site.
Fig. 6. Phenol and catechol yields calculated at 10% guaiacol conversion.
The reaction rates obtained with the supported catalysts were
higher than the reported activities of unsupported molybdenum
nitride catalysts [10] (except for the CoMoN/SBA-15-A). This result
suggests that although the surface area of supported nitride cat-
alysts consists of 10% of the active nitride phase in comparison
to unsupported nitride catalysts whose surface area are made up
of only the active nitride phase, the use of a support in this reac-
tionenhancedactivitythroughgreater accessibilityof activesites to
reactants as a result of the supports porosity. Inaddition, improved
catalyst dispersion and possible electronic effects of supported cat-
alysts may have played a role in activity enhancement.
Thephenol/catechol selectivity(g/100gguaiacol converted) cal-
culated at 10% guaiacol conversion is presented in Table 4 and
illustrated in Fig. 6: SBA-15-supported catalysts produced signif-
icantly more phenol than catechol, while the alumina-supported
catalysts produced more catechol than phenol. These results indi-
cate a higher selectivity of the nitrides supported on alumina
I. Tyrone Ghampson et al. / Applied Catalysis A: General 435436 (2012) 5160 59
Table 4
Total conversion, reaction rates and phenol/catechol selectivity of catalysts under study.
Catalyst Conversion
a
(%) Reaction rate
(10
6
mol g
1
catalyst
s
1
)
Intrinsic activity
(10
4
mol Mo at
1
s
1
)
Product distribution
b
(%)
Phenol Catechol
MoN/Al
2
O
3
-A 66 14.0 160.6 1 26
MoN/Al
2
O
3
-NH 62 12.7 139.2 2 26
CoMoN/Al
2
O
3
-A 58 10.8 124.7 5 22
NiMoS/Al
2
O
3
63 7.4 76.3 1 11
MoN/SBA15-A 44 6.4 84.7 26 6
MoN/SBA15-NH 40 5.2 74.7 22 9
CoMoN/SBA15-A 24 0.5 6.8 12 6
a
Total conversion after 4h of reaction.
b
Phenol and catechol selectivity calculated at 10% guaiacol conversion.
toward the demethylation pathway compared to that of nitrides
supported on SBA-15 which had a higher capacity for aromatic
carbonoxygen hydrogenolysis. The results conrmthat both sup-
port modied the active sites of the nitrides. Hydrogenolysis
of the methyl-oxygen bond, as well as hydrogenolysis of the
C
aromatic
OCH
3
bond, has been suggested to take place on both
the support surface and on the nitride species [38]. However, since
the bare SBA-15 silica support was catalytically inert for HDOreac-
tion the higher phenol production displayed by these catalysts may
be attributed primarily to the DMO sites on the Mo nitride and
oxynitride.
3.2.2. Activity of differently synthesized Mo nitride catalysts:
method 1 vs. method 2
It can be observed in Table 4 that the nitride catalysts prepared
using method 1 displayed higher reaction rates than the nitride
catalysts prepared using method 2. In general, it was concluded
that the higher bulk N/Mo ratio (shown in Table 1) and the for-
mation of -Mo
2
N phase (deduced from XRD) by the method 1
synthesis procedure led to a more favorable conversion of guaia-
col than the formation of -Mo
2
N
0.78
by the method 2 procedure.
This is consistent with the results of unsupported Mo nitrides,
which revealed higher guaiacol conversion for catalysts that con-
tainpredominantly-Mo
2
Nincomparisontocatalysts that contain
-Mo
2
N
0.78
particles [10]. There is no observable relationship
between the activity and the distribution of surface Mo nitride
and oxynitride species, suggesting that there might be multiple
active phases with differing activities for guaiacol conversion in
the molybdenumoxygennitrogen system. A study on the atomic
scale knowledge of the active phase of Mo nitride and oxynitride
for HDO catalysis is warranted.
Fig. 6 shows differences in phenol/catechol yields for the SBA-
15-supported nitride catalysts prepared using different nitridation
procedures. The higher phenol yield obtained for the MoN/SBA-
15-A catalyst could also be due the presence of the -Mo
2
N
crystalline phase which has a greater capacity to directly cleave
the C
aromatic
OCH
3
bond [10]. Also from Fig. 6, it can be observed
that MoN/Al
2
O
3
-A and MoN/Al
2
O
3
-NH catalysts displayed similar
phenol/catechol production. This behavior is not clear. However, it
could be due to the dominant effect of the Lewis acidity of the alu-
mina support (forming catechol) which slightly modied the active
sites.
The addition of Co to the supported Mo nitrides surprisingly
did not enhance guaiacol conversion rate. In fact, the reaction rate
was 1.1 times lower for the CoMoN/Al
2
O
3
-A catalyst as compared
to the MoN/Al
2
O
3
-A catalyst, while the CoMoN/SBA-15-A catalyst
was signicantly less active as comparedto the MoN/SBA-15-Acat-
alysts. This could be explained by the incomplete formation of the
Co
3
Mo
3
N phase (from XRD and XPS results) which typically has a
higher C-X (X:S or N) hydrogenolysis rate than monometallic Mo
nitride catalyst [39,40].
Fig. 7. Variation of the transformation of guaiacol and the yield of products with
time for NiMoS/Al
2
O
3
catalyst.
3.2.3. Comparison of nitrides to commercial sulded NiMo/Al
2
O
3
catalyst
The activity of a commercial sulded NiMo/Al
2
O
3
catalyst was
alsotestedfor theHDOof guaiacol andillustratedinFig. 7: theprod-
uct distribution as a function of time was similar to those obtained
with the alumina-supported nitride catalysts. This conrms the
strong inuence of the acidic properties of the alumina support.
The activity per gramof catalyst given in Table 4 obtained with the
commercial reference catalyst was two times lower as compared to
the Mo nitrides supported on alumina. Furthermore, the alumina-
supported nitrides showed more thantwo times higher activity per
Mo atoms than the sulded NiMo/Al
2
O
3
catalyst. This preliminary
comparison of results is encouraging for the application of nitrides
for HDO catalysis.
The phenol/catechol ratio as shown in Fig. 6 with the sul-
ded NiMo/Al
2
O
3
catalyst was comparable to the Mo nitrides
supported on alumina. This result further proves the overall capac-
ity of alumina-supported catalysts for demethylation of guaiacol,
regardless of the active phase [41]. This is in good agreement with
other reported ndings which indicated that Lewis acid sites of
the alumina support were mainly responsible for the conversion of
guaiacol to catechol [5,41]. Further information regarding leaching
and post-catalyst evaluation will be required for a more complete
evaluation of catalyst stability.
4. Conclusion
We have prepared alumina- and SBA-15-supported Mo nitride
catalysts and showed their reactivity for the HDO of guaiacol in
terms of reaction rates and phenol/catechol yields. Catalysts were
synthesized by nitridation using two procedures: thermal conver-
sion in ammonia and thermal conversion in nitrogen/hydrogen
mixtures. Nitridation using ammonia resulted in higher activities
60 I. Tyrone Ghampson et al. / Applied Catalysis A: General 435436 (2012) 5160
in both supports and correlated with the formation of -Mo
2
N
and a higher N/Mo ratio in the catalyst, in good agreement with
previously published results with unsupported nitride catalysts.
Nitridation using nitrogen/hydrogen resulted in the formation of
-Mo
2
N
0.78
. Consistent with ndings for MoS
2
-based catalysts on
alumina, the alumina supported nitrides resulted in signicant
conversion of guaiacol to catechol. The silica supported catalysts
resulted in minimal catechol production, and maximum phenol
production. For the catalysts studied here, the major factor inu-
encing activity was the active phase whereas the major factor
affecting phenol production over catechol production was the sup-
port which modied the nature of the active sites of nitrides. The
alumina-supported nitride catalysts were more active than a con-
ventional sulded NiMo/Al
2
O
3
catalyst and were generally more
active than previously reported unsupported nitride catalysts. The
addition of cobalt did not have a promoting effect on HDO activity.
These results are encouraging for the application of ordered meso-
porous silicas as supports for molybdenumnitride based catalysts
in HDO applications.
Acknowledgments
The authors acknowledge the nancial support of DOE Epscor
Grant #DE-FG02-07ER46373 and nancial support from CONICYT
Chile, projects PFB-27, PIA-ACT-130 and FONDECYT No. 1100512
grants. I. Tyrone Ghampson is indebted to NSF Career Award
0547103 for sponsoring a trip to the University of Concepcin. The
authors alsogratefullyacknowledgethetechnical assistanceof Nick
Hill and Manuel Veliz.
References
[1] J. Zakzeski, P.C.A. Bruijnincx, A.L. Jongerius, B.M. Weckhuysen, Chem. Rev. 110
(2010) 35523599.
[2] E. Furimsky, Appl. Catal. A 199 (2000) 147190.
[3] A. Centeno, E. Laurent, B. Delmon, J. Catal. 154 (1995) 288298.
[4] V.N. Bui, D. Laurenti, P. Delichre, C. Geantet, Appl. Catal. B101 (2011) 246255.
[5] E. Laurent, B. Delmon, Appl. Catal. A 109 (1994) 97115.
[6] M. Ferrari, R. Maggi, B. Delmon, P. Grange, J. Catal. 198 (2001) 4755.
[7] G.M. Dolce, P.E. Savage, L.T. Thompson, Energy Fuels 11 (1997) 668675.
[8] M. Nagai, Y. Goto, A. Irisawa, S. Omi, J. Catal. 191 (2000) 128137.
[9] J. Monnier, H. Sulimma, A. Dalai, G. Caravaggio, Appl. Catal. A 382 (2010)
176180.
[10] I.T. Ghampson, C. Sepulveda, R. Garcia, B.G. Frederick, M.C. Wheeler, N. Escalona,
W.J. DeSisto, Appl. Catal. A 413414 (2012) 7884.
[11] I.T. Ghampson, C. Newman, L. Kong, E. Pier, K.D. Hurley, R.A. Pollock, B.R. Walsh,
B. Goundie, J. Wright, M.C. Wheeler, R.W. Meulenberg, W.J. DeSisto, B.G. Fred-
erick, R.N. Austin, Appl. Catal. A 388 (2010) 5767.
[12] R. Nava, B. Pawelec, P. Casta no, M.C. lvarez-Galvn, C.V. Loricera, J.L.G. Fierro,
Appl. Catal. B 92 (2009) 154167.
[13] K. Soni, K.C. Mouli, A.K. Dalai, J. Adjaye, Catal. Lett. 136 (2010) 116125.
[14] R.A. Pollock, B.R. Walsh, J. Fry, I.T. Ghampson, Y.B. Melnichenko, H. Kaiser,
R. Pynn, W.J. DeSisto, M.C. Wheeler, B.G. Frederick, Chem. Mater. 23 (2011)
38283840.
[15] A. Galarneau, M. Nader, F. Guenneau, F. Di Renzo, A. Gedeon, J. Phys. Chem. C
111 (2007) 82688277.
[16] Z.H. Luan, M. Hartmann, D.Y. Zhao, W.Z. Zhou, L. Kevan, Chem. Mater. 11 (1999)
16211627.
[17] M. Jaroniec, M. Kruk, J.P. Olivier, Langmuir 15 (1999) 54105413.
[18] C.D. Wagner, L.E. Davis, M.V. Zeller, J.A. Taylor, R.H. Raymond, L.H. Gale, Surf.
Interface Anal. 3 (1981) 211225.
[19] R. Cid, G. Pecchi, Appl. Catal. 14 (1985) 1521.
[20] M. Nagai, Y. Goto, O. Uchino, S. Omi, Catal. Today 45 (1998) 335340.
[21] H.K. Park, J.K. Lee, J.K. Yoo, E.S. Ko, D.S. Kim, K.L. Kim, Appl. Catal. A 150 (1997)
2135.
[22] D.Y. Zhao, J.Y. Sun, Q.Z. Li, G.D. Stucky, Chem. Mater. 12 (2000) 275279.
[23] S.J. Gregg, K.S.W. Sing, Adsorption, Surface Area and Porosity, Academic Press,
London, 1982.
[24] O. Gonzlez, H. Prez, P. Navarro, L.C. Almeida, J.G. Pacheco, M. Montes, Catal.
Today 148 (2009) 140147.
[25] R. Nava, B. Pawelec, J. Morales, R.A. Ortega, J.L.G. Fierro, Micropor. Mesopor.
Mater. 118 (2009) 189201.
[26] M. Kruk, M. Jaroniec, C.H. Ko, R. Ryoo, Chem. Mater. 12 (2000) 19611968.
[27] A. Galarneau, H. Cambon, F. Di Renzo, R. Ryoo, M. Choir, F. Fibula, NewJ. Chem.
27 (2003) 7379.
[28] C. Shi, A.M. Zhu, X.F. Yang, C.T. Au, Appl. Catal. A 276 (2004) 223230.
[29] S.W. Yang, C. Li, J. Xu, Q. In, J. Phys. Chem. B 102 (1998) 69866993.
[30] K. Hade, M. Nagai, S. Omi, J. Phys. Chem. B 105 (2001) 40844093.
[31] T. Nagata, G. Koblmuller, O. Brewage, C.S. Gallant, J.S. Speck, Appl. Phys. Lett.
95 (2009).
[32] M. Nagai, J. Takada, S. Omi, J. Phys. Chem. B 103 (1999) 1018010188.
[33] J.S. Girard on, E. Quintet, A. Removal-Constant, P.A. Chernavskii, L. Gingerbread,
A.Y. Kodak, J. Catal. 248 (2007) 143157.
[34] K. Hade, M. Nagai, S. Omi, J. Phys. Chem. B 104 (2000) 20902098.
[35] V.N. Bui, G. Tousling, D. Laurenti, C. Microdots, C. Geantet, Catal. Today 143
(2009) 172178.
[36] T. Nimmanwudipong, R.C. Runabout, D.E. Block, B.C. Gates, Energy Fuels 25
(2011) 34173427.
[37] A. Popover, E. Kondratieva, J.M. Gospel, L. Marie, P. Basin, J.-P. Gibson, A. Trivet,
F. Maguey., J. Phys. Chem. C 114 (2010) 1566115670.
[38] G. de la Puente, A. Gil, J.J. Pis, P. Grange, Langmuir 15 (1999) 58005806.
[39] J.W. Logan, J.L. Heiser, K.R. McCrea, B.D. Gates, M.E. Bussell, Catal. Lett. 56(1998)
165171.
[40] Y. Li, Y. Zhang, R. Raval, C. Li, R. Zhai, Q. In, Catal. Lett. 48 (1997) 239245.
[41] E. Laurent, B. Delmon, Appl. Catal. A 109 (1994) 7796.

Das könnte Ihnen auch gefallen